Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Synergistic luminescence of Tb3+ and Pr3+ ions in Sr2Y(MoO4)4 for high-resolution optical thermometry

Yosra Bahrouniab, Ikhlas Kachoua, Kamel Saidiac and Mohamed Dammak*a
aApplied Physics Laboratory, Faculty of Sciences of Sfax, Department of Physics, University of Sfax, Sfax, Tunisia. E-mail: madidammak@yahoo.fr; mohamed.dammak@fss.usf.tn
bUniversity of La Laguna, Physics Department, MALTA-Consolider Group, IMN and IUdEA, P. O. Box 456, E-38206 San Cristóbal de La Laguna, Santa Cruz de Tenerife, Spain
cUniversity of Sfax, Department of Physics, Sfax Preparatory Engineering Institute, 1172 - 3000Sfax, Tunisia

Received 30th June 2025 , Accepted 22nd September 2025

First published on 26th September 2025


Abstract

Phosphors of the SrY2(MoO4)4 series co-doped with Pr3+ and Tb3+ ions were synthesized and investigated for optical temperature sensing applications. Their structure, morphology, and photoluminescent properties were thoroughly characterized using X-ray diffraction (XRD), field emission scanning electron microscopy (FESEM), UV-vis absorption spectroscopy, and photoluminescence (PL) analysis. The intense dual-mode color-tunable emission revealed prominent transition bands corresponding to 5D47Fj (j = 6, 5, 4, 3) for Tb3+ and 3P03H6 and 3P03F2 for Pr3+ ions under UV excitation at 450 nm. Chromaticity parameters for both doped and co-doped phosphors were also investigated, demonstrating emissions concentrated in the green and white regions. These properties highlight their potential for applications in solid-state lighting. The synthesized phosphors were further evaluated for their potential in optical temperature sensing based on the fluorescence intensity ratio (FIR) principle. The SrY2(MoO4)4: 0.1 Pr3+/0.4 Tb3+ phosphors demonstrated strong red emission with a color purity of 85% and achieved a high relative sensitivity of 1.9% K−1 for the 488 nm and 648 nm emission ratio, exceeding several reported molybdate-based thermometers, confirming their potential as multifunctional materials for efficient red component generation in solid-state lighting and precise, non-contact optical temperature sensing.


Introduction

Temperature is a fundamental physical parameter crucial for assessing heat and cold, significantly impacting many scientific and commercial applications.1,2 In recent years, there has been an increasing interest in the development of more precise, easy, and economical temperature measuring methods. Conventional techniques are often classified into contact and non-contact procedures. Contact methods necessitate physical touch between the sensor and the item, resulting in possible delays and difficulties in high-temperature settings, whereas non-contact approaches mitigate these constraints and are progressively favored in many domains.3,4

For non-contact techniques, optical thermometry has garnered significant attention due to its high precision, low error margin, and cost efficiency.5 Methods such as emission intensity, fluorescence intensity ratio (FIR), decay lifetime, and emission band shift have been extensively investigated.6–8 FIR technology, in particular, has emerged as the most popular due to its robust anti-interference capability, accuracy, and operational simplicity.9,10 This method typically requires two luminescent centers with distinct temperature-dependent responses.11 As such, phosphors doped with lanthanide ions and transition metals are frequently used in optical temperature sensing, leveraging the impact of temperature variations on their luminescence properties.

Optical temperature sensing using FIR generally involves thermally coupled levels (TCL).12,13 By employing FIR-based sensors, dependence on external conditions is minimized, enhancing accuracy and resolution. For this purpose, lanthanide ions such as Tb3+, Pr3+, and Ho3+ have been widely studied. Selecting a suitable host material is also critical; low phonon energy is preferred to reduce non-radiative transitions.14,15 Molybdate-based matrices have been widely used owing to their exceptional properties, including high chemical stability, brilliance, high melting point, low toxicity, and long persistence without radioactive radiation.16–18 In this study, the focus is on SrY2 (MoO4)4: 0.1 Pr3+/0.4 Tb3+ (SYMO: 0.1 Pr3+/0.4 Tb3+) phosphors, synthesized using the solid-state method. Pr3+ and Tb3+ are well-known for their efficient luminescent properties, with Pr3+ emitting green-blue and red light from its 3P0 and 1D2 energy levels to the ground state, and Tb3+ predominantly emitting green light from the 5D47Fx transitions (x = 3, 4, 5, 6).19,20 The luminescence properties of the 3P0 and 1D2 excited states, which decay to the 3H4 ground state, are strongly affected by host lattice parameters such as phonon energy and crystal structure. In certain compounds, however, the 3P0 emission can be entirely suppressed.21–23

This study introduces a novel approach to enhance the luminescence efficiency of Sr2Y(MoO4)4 (SYMO) phosphors via co-doping with Pr3+ and Tb3+ ions. For the first time, this co-doped system exhibits tunable emissions spanning the green to near-white spectral regions, paving the way for advanced color modulation in photonic materials. Structural and optical characterizations confirm the successful integration of both activator ions into the SYMO host lattice, resulting in efficient and thermally stable luminescence. The optimized SYMO: 0.1 Pr3+/0.4 Tb3+ composition demonstrates promising dual-functionality in photoluminescence and high-resolution optical thermometry, positioning it as a strong candidate for applications in smart lighting, white LEDs, and non-contact temperature sensing technologies.

Experimental section

Sample preparation

The Pr3+/Tb3+ codoped SYMO phosphors were prepared through a traditional solid state method. First, high purity SrCO3, MoO3, Y2O3, Pr6O11 and Tb4O7 were thoroughly grounded in an agate mortar and placed into an alumina crucible. The mixture was pre-calcined at 550 °C for 5 hours to remove volatiles and initiate the reaction. The samples were sintered at 1000 °C for 5 hours in a flowing CO2 atmosphere. This atmosphere was chosen to help control the oxidation states of Pr3+ and Tb3+ ions. Although CO2 is generally considered mildly oxidizing, at high temperatures it can also act in a slightly reducing manner, which helps stabilize the desired oxidation states during the sintering process.

Characterization

The materials were characterized by the use of chosen analytical techniques. The phase purity and crystal structure of samples were identified through powder X-ray diffraction (XRD) patterns on PANalytical X'Pert Pro diffractometer in the Bragg–Brentano geometry. The sample morphologies were examined using a scanning electron microscope (JEOL, JSM 6510LV). Excitation and emission spectra were acquired using a Jobin Yvon Fluoromax-4 spectrofluorometer (Horiba) equipped with a 150 W xenon lamp as the excitation source. To investigate the temperature dependence, photoluminescence spectra were collected with a Horiba Jobin Yvon HR 320 spectrometer, with the sample mounted inside a cryostat.

Results and discussion

Structural study and morphology

Fig. S1 presents the XRD pattern of the synthesized SYMO sample together with the reference data from the JCPDS card no. 82-2369. The observed diffraction peaks match well with those of the pure SYMO phase, confirming good agreement with the standard pattern (JCPDS no. 82-2369).24 All observed diffraction peaks can be indexed to the tetragonal scheelite-type structure of Sr2Y(MoO4)4, As show in Fig. S1, which has a scheelite-type structure in space group I41/a, indicating that no other impurities are included in the as-obtained sample. Moreover, a Rietveld refinement analysis was performed on the Pr3+/Tb3+ doped SYMO samples to improve the clarity of the crystal structure. Fig. 1a shows the results of the refinement. The observed and calculated profiles are in remarkable agreement, indicating the refinement's high accuracy. A comprehensive summary of the structural properties of the samples is given in Table S1. From the figure, we observed that the difference profile of the SYMO was almost flat, indicating a very good fit between observed and calculated profiles. The reliability factor was less than 2, indicating good fitting results and further confirming the phase purity.
image file: d5ra04652a-f1.tif
Fig. 1 (a) Rietveld refinement of XRD patterns. (b) SEM images for SYMO codoped 0.1 Pr3+/0.4 Tb3+ sample.

Fig. 1b illustrates the morphology of SYMO: 0.1 Pr3+/0.4 Tb3+ phosphors, revealing a rod-like one-dimensional structure indicative of anisotropic crystal growth. The rods exhibit an average diameter of about 10 μm and axial lengths extending to several hundred micrometers. These observations suggest that the Pr3+/Tb3+ co-doping concentration has little effect on the crystal growth behavior.

Optical characterization

UV-visible. Based on the UV-vis reflectance, the observed emission features may be attributed to the presence of energy levels introduced by Pr3+ and Tb3+ ions within the host band gap, which facilitate radiative transitions under excitation.25,26

The reflectance spectra of pure SYMO, SYMO: 0.4 Tb3+, SYMO: 0.1 Pr3+, and SYMO: 0.1 Pr3+/0.4 Tb3+ samples, recorded over the 200–900 nm range, are shown in Fig. 2.


image file: d5ra04652a-f2.tif
Fig. 2 UV-vis diffuse reflectance spectra of SYMO, SYMO: 0.1 Pr3+, SYMO: 0.4 Tb3+ and SYMO: 0.1 Pr3+/0.4 Tb3+.

A pronounced absorption band appears in the UV region (200–330 nm), which originates from the Mo–O charge transfer transition, corresponding to electron excitation from the O 2p states to the Mo 4d levels.27 Moreover, the appearance of bands centered at 446, 487 and 495 nm is monitored and is attributed to the appearance of the transitions: 3H43P2, 3H43P1 and 3H43P0 which related to the Pr3+ ions.

In the case of Pr3+, the introduction of 4f states near the valence band can lead to the formation of localized defect levels, effectively narrowing the optical band gap. Additionally, Pr3+ can introduce slight lattice distortions due to its ionic radius, which perturbs the Mo–O bonding environment and modifies the electronic band structure. Conversely, Tb3+ doping may slightly widen the band gap due to its higher electronegativity and relatively lower interaction with the host valence band states, leading to a modest blue-shift in the absorption edge. These effects reflect the different roles of 4f orbitals and local structural influences introduced by each dopant.

To evaluate the band gap energy, the reflectance spectra were transformed using the Kubelka–Munk function (eqn (1)). Subsequently, the F(R) values were analyzed with the Tauc relation (eqn (2)) in order to estimate the energy separation between the valence band and the conduction band of the synthesized samples, as expressed below:28,29

 
image file: d5ra04652a-t1.tif(1)
 
(F(R))n = A(Eg) (2)

The estimated band gap (Eg) values of the samples in this study are 3.5 eV for SYMO, 3.6 eV for SYMO: 0.4 Tb3+, 3.3 eV for SYMO: 0.1 Pr3+, and 3.4 eV for SYMO: 0.1 Pr3+/0.4 Tb3+, as determined by extrapolating the linear region of the plots shown in Fig. S3. These values are consistent with those reported in the literature.30

Photoluminescence properties. Fig. 3a displays the PL excitation spectra of the SYMO: 0.4 Tb3+ monitored at an emission wavelength of 534 nm. The spectrum shows that the broad band around a wavelength of about 296 nm is attributed to the ligand-to-metal transfer band, which corresponds to the ligand transfer from the O2− ion to the Mo6+ ion in the [MoO4]2− groups, and the sharp peaks at 351 nm (7F65D2), 359 nm (7F65G5), 369 nm (7F65G6), 378 nm (7F65D3) and 486 nm (7F65D5) are attributed to the transitions of Tb3+ ions.1,31 Upon excitation at 375 nm, as shown in Fig. 3b in the range of 420–520 nm, four PL components (488, 543, 584 and 624 nm) come from the electronic transitions of Tb3+ between a excited state (5D4) and ground states (7F3, 7F4, 7F5 and 7F6, respectively).32,33
image file: d5ra04652a-f3.tif
Fig. 3 Excitation and emission spectra (a and b), of the SYMO: 0.4 Tb3+ phosphor monitored at 534 nm, and excited at 375 nm, respectively. (c and d), of SYMO: 0.1 Pr3+ phosphor monitored at 620 nm, and excited at 450 nm, respectively.

Fig. 3c presents the PL excitation spectrum of the SYMO host monitored at an emission wavelength of 620 nm. A wide excitation band is observed between 250 and 320 nm, together with several sharp peaks in the 450–490 nm region. The broad feature centered at 285 nm is attributed to charge transfer from O2− to Mo6+ ions.34,35 Moreover the sharp peaks are ascribed to 3H43P2 (451 nm), 3H43P1 (472 nm) and 3H43P0 (487 nm).36,37 Upon 450 nm excitation, the studied material shows numerous sharp lines in the 460−760 nm range, the PL spectrum reveals, the emissions peaking at 530 nm, 556 nm, 625 nm, 649 nm corresponding to the 3P03H4, 3P13H5, 3P03H6 and 3P03F2 respectively, while two emissions at 619 nm and 687 nm are attributed to the 1D23H4 and 1D23H5 transitions of Pr3+ ions.38,39 In particular, the blue and orange-red emissions at 556 and 625 nm originate from two thermally coupled levels of 3P0 and 3P1 as shown in Fig. 3b.

Investigation of luminescence in SYMO: 0.1 Pr3+/0.4 Tb3+ phosphors. Fig. 4a depicts the excitation spectrum of SYMO: 0.1 Pr3+/0.4 Tb3+ microcrystals recorded at 534 nm emission. Strong and broad excitation bands appear in the 230–355 nm region, which are assigned to charge–transfer transitions arising from electron transfer between the O2− 2p orbitals and the Mo6+ 5d orbitals in the MoO4 units of the host lattice. In all spectra shown in Fig. 4a, sharp peaks centred at 355 nm (7F65D2: Tb3+), 375 nm (7F65D3: Tb3+), 472 nm (3H43P1: Pr3+) and 486 nm (7F65D5: Tb3+, 3H43P0: Pr3+) are observed, which were in accordance with those of the doped-only samples. Under 450 nm excitation, Fig. 4b presents the photoluminescence (PL) spectra of SYMO co-doped with Pr3+ and Tb3+ ions at room temperature. The material exhibits several sharp emission lines corresponding to bands centered at 488 nm (Pr3+, Tb3+), 530 nm (Pr3+), 544 nm (Tb3+), 620 nm (Tb3+, Pr3+), 648 nm (Pr3+) and 689 nm (Pr3+), within the 460–760 nm range.40–42
image file: d5ra04652a-f4.tif
Fig. 4 Excitation and emission spectra of the SYMO: 0.1 Pr3+/0.4 Tb3+ phosphors monitored at λem = 534 nm, and λex = 450 nm respectively.
CIE chromaticity coordinates. The CIE chromaticity coordinates (x, y) of the synthesized phosphors under 450 nm UV excitation were calculated and are presented in the CIE chromaticity diagram shown in Fig. 5.43 The color coordinates and corresponding correlated color temperature (CCT) values of the phosphors are summarized in Table S2. The CCT values were determined from the CIE coordinates using McCamy's formula:14,44
 
CCT = −437n3 + 3601n2 − 6861n + 5514.3 (3)

image file: d5ra04652a-t2.tif
Here, xe = 0.3320 and ye = 0.1858 represent the epicenter coordinates of the color.

image file: d5ra04652a-f5.tif
Fig. 5 CIE 1931 chromaticity diagram of SYMO phosphors codoped with (a) 0.4 Tb3+, (b) 0.1 Pr3+ and (c) 0.1 Pr3+/0.4 Tb3+ excited at λex = 450 nm.

All samples exhibit CCT values under 3200 K, indicating a warm appearance, as values above 3200 K are usually classified as cool light sources.45 Fig. 5a shows that the CIE coordinates of SYMO: 0.1Pr3+ phosphors fall within the red region. In Fig. 5b, under 375 nm excitation, the CIE coordinates of SYMO: 0.4 Tb3+ phosphors are located in the green region. Fig. 5c illustrates that the CIE coordinates of SYMO: 0.1 Pr3+/0.4 Tb3+ phosphors lie in the blue region. These coordinates provide a key measure of how the emitted light is perceived and are essential for evaluating the overall emission characteristics of the phosphors.

Thermal stability. Thermal stability is a crucial property for lighting applications, particularly for phosphors used in white LEDs (w-LEDs).46 High-power LED chips can reach temperatures up to 380 K, at which point luminescence intensity typically decreases significantly. To evaluate the thermal stability of SYMO: 0.1 Pr3+/0.4 Tb3+, PL spectra were recorded under 450 nm excitation at different temperatures, as shown in Fig. S4a. The PL emission profile remains largely unchanged with increasing temperature; however, the intensity of the transitions gradually decreases from 300 K to 380 K due to thermal quenching. Notably, the temperature response differs between the (MoO4)4− band and the Pr3+/Tb3+ bands: the (MoO4)4− band shows a sharp drop in intensity, whereas the Pr3+/Tb3+ emissions decrease more gradually, as illustrated in the histogram of Fig. S4b. Even at 380 K, the luminescence intensity retains 45% of its initial value.

Thermal quenching is primarily caused by non-radiative relaxation, with the number of excited electrons increasing as the temperature rises. This thermal activation process can be quantitatively described using the Arrhenius equation.

 
image file: d5ra04652a-t3.tif(4)

In the equation, I0 represents the luminescence intensity at room temperature, I is the intensity at elevated temperatures, K denotes the Boltzmann constant, and CCC is a constant. Fig. S4c shows the plot of (I0/I − 1) versus 1/KBT, which follows a linear relationship with a slope corresponding to the activation energy, ΔE. From this fitting, the thermal activation energy of SYMO: 0.1 Pr3+/0.4 Tb3+ is estimated to be 0.4 eV, confirming the high thermal stability of this phosphor. This value is notably higher than those reported for other phosphors, such as Lu2Mo3O12: Tb3+ (0.29 eV),47 and CaYAlO4: Tb3+ (0.39 eV)48 further indicating the superior thermal stability of SYMO: 0.1 Pr3+/0.4 Tb3+.

Temperature sensing properties

In further evaluation of the potential of the investigated materials for optical temperature sensing, the high-temperature PL spectra of SYMO: 0.1 Pr3+/0.4 Tb3+ were measured. It can be seen from Fig. 7 that the total luminescence intensity gradually decreases with the increase of the sample temperature under the excitation of 450 nm, in the range of 215–498 K, indicating that a temperature-dependent luminescence quenching takes place which related to the enhancement probability of the non-radiative processes.49 Interestingly, the 3P0 and 1D2 related emissions of Pr3+ show different temperature responses.50,51

The differing temperature responses of Tb3+ and Pr3+ ions suggest that these phosphors could be used in optical thermometry, with the FIR method serving to assess their performance in real temperature measurements. As illustrated in Fig. 6, the temperature-dependent luminescence property of Pr3+/Tb3+ ions in SYMO can be well interpreted by the above kinetic, process since the measured plots of FIR versus T are well fitted by equation:52,53

 
image file: d5ra04652a-t4.tif(5)
where A and B are fitting constants, ΔE is the effective energy difference, and K is the Boltzmann constant. Notably, the SYMO: 0.1 Pr3+/0.4 Tb3+ phosphor is suitable for thermal sensing within the temperature range of 215–400 K. The fluorescence intensity ratio (FIR) of the 3P03H4 transition (544 nm) to the 3P03F2 transition (648 nm) can be used for temperature calibration, as these two levels are exclusively associated with Pr3+ transitions. Additionally, the FIR of the 5D37F3 transition (Tb3+) to the 3P03F2 transition (Pr3+) is also evaluated. The integrated intensities for FIR values: FIR = I544 nm/I648 nm, FIR = I488 nm/I648 nm, and FIR = I531 nm/I648 nm as a function of temperature are illustrated in Fig. 7.


image file: d5ra04652a-f6.tif
Fig. 6 Temperature-dependent PL spectra of the SYMO phosphor codoped 0.1 Pr3+/0.4 Tb3+ excited at 450 nm.

image file: d5ra04652a-f7.tif
Fig. 7 FIR vs. temperature variations for SYMO: 0.1 Pr3+/0.4 Tb3+ corresponding to the transition intensity ratios: (a) 488/688 nm, (b) 544/648 nm, and (c) 531/688 nm.

In all cases, the experimental data were fitted using the exponential function in eqn (5), indicating that the temperature dependence of the FIR follows the Boltzmann distribution. All evaluated FIRs are suitable for thermal sensing, with one particular FIR exhibiting the highest thermometric performance.

Additionally, the absolute sensitivity (Sa) and relative sensitivity (Sr) are key parameters for evaluating the performance of optical thermometers and can be calculated using the following equations:54,55

 
image file: d5ra04652a-t5.tif(6)

By plotting Sr against temperature, we observed that a doping concentration of 40% Tb3+ combined with 10% Pr3+ optimizes thermometric performance. This configuration achieved a maximum Sr at approximately room temperature (RT) in the SYMO crystal, as shown in Fig. 8c corresponding to the FIR (488/648).


image file: d5ra04652a-f8.tif
Fig. 8 Variation of temperature sensitivity parameters (absolute and relative sensitivity) as a function of temperature for the SYMO: 0.1 Pr3+/0.4 Tb3+ sample, corresponding to the transition intensity ratios: (a) 488/688 nm, (b) 544/648 nm, and (c) 531/688 nm.

The variation in the intensity ratio between emissions at 544 nm and 648 nm yields a sensitivity Sr of 1.5% K−1. Considering the emissions at 488 nm and 648 nm, the Sr increases to 1.9% K−1, while for the lines at 531 nm and 648 nm, the Sr reaches 1.8% K−1, all evaluated at 328 K. The enhanced thermometric performance observed with these intensity ratios is attributed to the greater energy difference between the thermally coupled states.

Besides Sr, temperature resolution δT is one of the most important parameters for FIR thermometry and is denoted as:56,57

 
image file: d5ra04652a-t6.tif(7)
where δFIR and δFIR/FIR represent the uncertainty and the relative uncertainty of the FIR, respectively. In order to assess the δT of the PL lines at 215 K and 450 K, the FIRs between the integrated intensities of the FIR (544/648), FIR (531/648) and FIR (488/648) emission lines were calculated on the basis of the results. Fig. S5 displays the measured values of the uncertainties at different temperatures for the SYMO sample, and from the graph it is obvious that they are in the range of 0.25–0.5 K, which is an impressive result. A comparison of the sensitivity parameters with previously reported optical thermometry results is presented in Table 1. This analysis highlights the potential of the prepared sample for use in non-contact optical temperature sensing applications.

Table 1 Temperature measurement range, maximum value of Sr of different temperature indicators
Phosphors Temperature (K) Sr (% K−1) Reference
SrMoO4:0.01 Pr3+ 298–498 0.98 60
YVO4:0.005 Pr3+ 303–324 1.14 61
SrLu2O4: Pr3+ 78–490 1.63 62
LaMg0.4Nb0.5O3: Pr3+ 298–523 0.83 63
La2ZnTiO6: Tb3+ 293–573 0.734 64
CaWO4: Tb3+ 343–783 1.21 65
CaNb2O6:0.01Tb3+/0.5%Bi3+ 298–448 1.39 66
KBaY(MoO4)3: Tb3+/Eu3+ 303–503 1.10 67
SrY2(MoO4)4: Tb3+/Sm3+ 290–440 0.93 68
NaLu(WO4)2: Tb3+/Pr3+ 583–783 1.45 69
LuNbO4: Pr3+/Tb3+ 283–493 1.26 70
Y2Mo4O15: Pr3+/Yb3+ 298–508 1.24 71
SrY2(MoO4)2: Pr3+/Tb3+ 215–400 1.9 This work


To evaluate the accuracy of the temperature measurement method, the thermometric parameters (FIR values) were repeatedly measured while cycling the sample between low and high temperatures, as shown in Fig. S6. The repeatability (R) was calculated using the following equation:58,59

 
image file: d5ra04652a-t7.tif(8)
Here, Mi(T)c_represents the measurement parameter (FIR or band centroid) for the i-th cycle, and M(T)c is the average value of M(T) over seven cycles. The FIR values exhibit reversible changes with temperature, and all measured FIRs exceeded 97% across the tested temperature range, confirming the excellent repeatability and reliability of the employed thermometric methods.

Conclusion

In conclusion, the SYMO: 0.1 Pr3+/0.4 Tb3+ phosphor series was successfully synthesized via the solid-state method. X-ray diffraction confirmed the formation of phase-pure microcrystals with a scheelite-like tetragonal structure. The materials exhibited a wide optical band gap of 3.7 eV, which is consistent with strong host lattice absorption in the UV region and supports efficient excitation of the rare-earth ions under near-UV or blue light (450 nm). These Pr3+ activated SYMO phosphors exhibit excellent structural and optical properties, making them promising candidates for optical thermometry. The photoluminescence (PL) spectra revealed a strong contribution from the host matrix under 450 nm excitation, characterized by a broadband profile likely arising from overlapping emissions. Chromatic coordinates derived from the PL spectra place the emissions predominantly in the green and white regions, highlighting their suitability for luminescent applications.

Thermometric characterization, conducted within the 215–400 K range using fluorescence intensity ratio (FIR) techniques, demonstrated the material's effectiveness for precise temperature sensing. Among the various emission pairs analyzed (488/648 nm, 544/648 nm, and 531/648 nm), the I488/I648 nm ratio, involving both Tb3+ and Pr3+ ions, exhibited the best thermometric performance. This pair achieved a maximum sensitivity (Sr) of 1.9% K−1 near room temperature (RT), underscoring the material's potential for practical applications in non-contact optical thermometry. Overall, the structural, optical, and thermometric enhancements observed in the SYMO: 0.1 Pr3+/0.4 Tb3+ phosphor makes it an excellent candidate for advanced optoelectronic applications, including precise temperature sensing across a wide range of environments. These findings underscore the significant potential of this material for practical use in cutting-edge technologies.

Conflicts of interest

There are no conflicts to declare.

Data availability

All data underlying the results are available as part of the article and no additional source data are required.

Supplementary information: supporting data and analyses, namely XRD patterns, UV-vis diffuse reflectance spectra, Tauc plots, temperature-dependent photoluminescence spectra, FIR-based thermometric performance (including temperature uncertainty and cyclic repeatability), as well as Rietveld refinement data, crystal structure parameters, and CIE chromaticity coordinates of the studied SYMO: Pr3+/Tb3+ phosphors. See DOI: https://doi.org/10.1039/d5ra04652a.

References

  1. L. Chen, K. He, G. Bai, H. Xie, X. Yang and S. Xu, J. Alloys Compd., 2020, 846, 156425H CrossRef.
  2. H. Y. Zhang, G. Li, S. Liu, H. Li, Y. Gong, Y. Liang and Y. Chen, Chem. Eng. J., 2020, 396, 125251 CrossRef CAS.
  3. D. Chen, S. Liu, Y. Zhou, Z. Wan, P. Huang and Z. Ji, J. Mater. Chem. C, 2016, 4, 9044–9051 RSC.
  4. H. Suo, X. Zhao, Z. Zhang, Y. Wang, J. Sun, M. Jin and C. Guo, Laser Photonics Rev., 2020, 15, 2000319 CrossRef.
  5. S. Senapati and K. K. Nanda, J. Mater. Chem. C, 2017, 5, 1074–1082 RSC.
  6. M. Behrendt, S. Mahlik, K. Szczodrowski, B. Kuklinski and M. Grinberg, Phys. Chem. Chem. Phys., 2016, 18, 22266–22275 RSC.
  7. M. D. Wisser, M. Chea, Y. Lin, D. M. Wu, W. L. Mao, A. Salleo and J. A. Dionne, Nano Lett., 2015, 15, 1891–1897 CrossRef CAS PubMed.
  8. Z. E. A. A. Taleb, K. Saïdi, M. Dammak, D. Przybylska and T. Grzyb, Dalton Trans., 2023, 52, 4954–4963 RSC.
  9. X. Li, M. Dong, F. Hu, Y. Qin, L. Zhao, X. Wei, Y. Chen, C. Duan and M. Yin, Ceram. Int., 2016, 42, 6094–6099 CrossRef CAS.
  10. A. Zhang, Z. Sun, G. Liu, Z. Fu, Z. Hao, J. Zhang and Y. Wei, J. Alloys Compd., 2017, 728, 476–483 CrossRef CAS.
  11. Z. Antic, M. Dramicanin, K. Prashanthi, D. Jovanovic, S. Kuzman and T. Thundat, Adv. Mater., 2016, 28, 7745–7752 CrossRef CAS PubMed.
  12. Y. Cheng, Y. Gao, H. Lin, F. Huang and Y. Wang, J. Mater. Chem. C, 2018, 6, 7462–7478 RSC.
  13. G. Bao, K. L. Wong, D. Jin and P. A. Tanner, Light Sci. Appl., 2018, 7, 96 CrossRef PubMed.
  14. K. Saidi, M. Yangui, C. H. Álvarez, M. Dammak, I. R. M. Benenzuela and M. Runowski, ACS Appl. Mater. Interfaces, 2024, 16, 19137–19149 CrossRef CAS PubMed.
  15. S. Tomar and C. Shivakumara, Methods Appl. Fluoresc., 2023, 11, 024001 CrossRef PubMed.
  16. G. Li, L. Li, M. Li, W. Bao, Y. Song, S. Gan, H. Zou and X. Xu, J. Alloys Compd., 2013, 550, 1–8 CrossRef CAS.
  17. Y. Wang, C. Lin, H. Zheng, D. Sun, L. Li and B. Chen, J. Alloys Compd., 2013, 559, 123–128 CrossRef CAS.
  18. S. F. Wang, K. K. Rao, Y. R. Wang, Y. F. Hsu, S. H. Chen and Y. C. Lu, J. Am. Ceram. Soc., 2009, 92, 1732–1738 CrossRef CAS.
  19. H. Qin, X. Gong, Y. Chen, J. Huang, Y. Lin, Z. Luo and Y. Huang, J. Lumin., 2019, 210, 52–57 CrossRef CAS.
  20. H. Boubekri, M. Diaf, K. Labbaci, L. Guerbous, T. Duvaut and J. P. Jouart, J. Alloys Compd., 2013, 575, 339–343 CrossRef CAS.
  21. W. T. Carnall, P. R. Fields and K. Rajnak, J. Chem. Phys., 1968, 49, 4424–4442 CrossRef CAS.
  22. H. H. Caspers, H. E. Rast and R. A. Buchanan, J. Chem. Phys., 1965, 43, 2124–2128 CrossRef CAS.
  23. J. Liu, Z. Shi, Q. Song, D. Li, N. Li, Y. Xue, J. Xu, J. Xu, Q. Wang and X. Xu, Opt. Mater., 2020, 108, 110219 CrossRef CAS.
  24. S. Cho, Solid State Sci., 2020, 101, 106155 CrossRef CAS.
  25. K. Saidi, M. Dammak, K. Soler-Carracedo and I. R. Martín, Dalton Trans., 2022, 51, 5108–5117 RSC.
  26. D. Babu, D. Jagadeesan and T. V. L. Thejaswini, Visible light responsive heterophase Titania monoliths for the fast and efficient photocatalytic decontamination of organic pollutants, Sci. Rep., 2024, 14, 27441 CrossRef CAS PubMed.
  27. M. C. Oliveira, J. Andrés, L. Gracia, M. S. M. P. de Oliveira, J. M. R. Mercury, E. Longo and I. C. Nogueira, Appl. Surf. Sci., 2019, 463, 907–917 CrossRef CAS.
  28. H. Zhang, Y. J. Liang, Y. L. Zhu, S. Q. Liu, K. Li, J. Yang and W. Lei, J. Alloys Compd., 2018, 767, 1030–1040 CrossRef CAS.
  29. P. Kubelka and F. Munk-Aussig, Z. Tech. Phys., 1931, 12, 593–601 Search PubMed.
  30. Y. Xiang, L. Yang, C. Liao, X. Xiang, X. Tang, H. Tang and J. Zhu, J. Adv. Ceram., 2023, 12, 848–860 CrossRef CAS.
  31. S. Chahar, R. Devi, M. Dalal, M. Bala, J. Dalal, P. Boora, V. B. Taxak, R. Lather and S. P. Khatkar, Ceram. Int., 2019, 45, 606–613 CrossRef CAS.
  32. F. F. Hu, H. L. Gong and G. A. Ashraf, et al., J. Alloys Compd., 2022, 909, 164776 CrossRef CAS.
  33. Y. K. Zheng, T. S. Yang and Y. F. Xiang, et al., J. Alloys Compd., 2022, 911, 165087 CrossRef CAS.
  34. L. Zhao, Z. Xicheng and J. Yuanru, Hydrothermal preparation and photoluminescent property of LiY(MoO4)2:Pr3+ red phosphors for white light-emitting diodes, J. Rare Earths, 2015, 33(1), 33 CrossRef.
  35. L. Tang, Q. Meng, W. Sun and S. Lü, Preparation and temperature sensing behavior of NaY(MoO4)2: Pr3+, Tb3+phosphors, J. Lumin., 2021, 230, 117728 CrossRef CAS.
  36. Y. Gao, F. Huang, H. Lin, J. Xu and Y. S. Wang, Sens. Actuators, B, 2017, 243, 137–143 CrossRef CAS.
  37. W. Tang, S. Wang, Z. Li, Y. Sun, L. Zheng, R. Zhang, B. Yang, W. Cao and M. Yu, Appl. Phys. Lett., 2016, 108, 061902 CrossRef.
  38. M. Sojka, M. Runowski, P. Woźny, L. D. Carlos, E. Zych and S. Lis, J. Mater. Chem. C, 2021, 9, 13818–13831 RSC.
  39. E. Cavalli, F. Angiuli, P. Boutinaud and R. Mahiou, J. Solid State Chem., 2012, 185, 136–142 CrossRef CAS.
  40. L. X. Lovisa, Y. L. R. L. Fernandes, L. M. P. Garcia, B. S. Barros, E. Longo, C. A. Paskocimas, M. R. D. Bomio and F. V. Motta, Opt. Mater., 2019, 96, 109332 CrossRef CAS.
  41. L. L. Tang, Q. Y. Meng, L. Bai, W. J. Sun and C. W. Wang, J. Lumin., 2022, 242, 118570 CrossRef CAS.
  42. D. S. Sun, L. L. Zhang, Z. D. Hao, H. Wu, H. J. Wu, Y. S. Luo, L. Yang, X. Zhang, F. Liu and J. H. Zhang, Dalton Trans., 2020, 49, 17779 RSC.
  43. P. Singh, S. Modanwal, H. Mishra and S. B. Rai, RSC Adv., 2023, 13, 22663–22674 RSC.
  44. C. S. McCamy, Color Res. Appl., 1992, 17, 142–144 CrossRef.
  45. I. Kachou, K. Saidi, C. Hernandez-Alvarez, M. Dammak and I. R. Martín, Mater. Adv., 2024, 5, 8280–8293,  10.1039/d4ma00699b.
  46. J. Qin, C. Hu, B. Lei, J. Li, Y. Liu, S. Ye and M. Pan, J. Mater. Sci. Technol., 2014, 30, 290–294 CrossRef CAS.
  47. N. Huang, G. Lu, B. Bai, Z. Chen, M. Zhang, Y. Li, C. Cao and A. Xie, Chemosensors, 2022, 10(12), 533 CrossRef CAS.
  48. Y. Wang, J. Cheng, Z. Wang, Y. Gong, C. Tu, J. Huang, Y. Sun and Y. Yu, Crystals, 2022, 12(12), 1729 CrossRef CAS.
  49. D. V. Lapaev, V. G. Nikiforov, V. S. Lobkov, A. A. Knyazev and Y. G. Galyametdinov, A photostable vitrified film based on a terbium(III) β-diketonate complex as a sensing element for reusable luminescent thermometers, J. Mater. Chem. C, 2018, 6, 9475–9481 RSC.
  50. D. V. Lapaev, V. G. Nikiforov, V. S. Lobkov, A. A. Knyazev and Y. G. Galyametdinov, J. Mater. Chem. C, 2018, 6, 9475–9481,  10.1039/C8TC01288A.
  51. M. Sojka, W. Piotrowski, L. Marciniak and E. Zych, J. Alloys Compd., 2024, 970, 172662,  DOI:10.1016/j.jallcom.2023.172662.
  52. F. Ayachi, K. Saidi, W. Chaabani and M. Dammak, J. Lumin., 2021, 240, 118451 CrossRef CAS.
  53. P. Haro-González, I. R. Martín, N. E. Capuj and F. Lahoz, Opt. Mater., 2010, 32, 1349–1351,  DOI:10.1016/j.optmat.2010.04.013.
  54. S. Jana, A. Mondal, J. Manam and S. Das, J. Alloys Compd., 2020, 821, 153342 CrossRef CAS.
  55. F. Chi, B. Jiang, Z. Zhao, Y. Chen, X. Wei, C. Duan, M. Yin and W. Xu, Sens. Actuators B-Chem., 2019, 296, 126640 CrossRef CAS.
  56. V. Kesarwani and V. K. Rai, J. Appl. Phys., 2022, 132, 113102 CrossRef CAS.
  57. Y. Bahrouni, I. Kachou, K. Saidi, T. Kallel, M. Dammak and I. Mediavilla-Jiménez, Mater. Adv., 2025, 6, 1307–1318 RSC.
  58. Z. Wang, M. Jia, M. Zhang, X. Jin, H. Xu and Z. Fu, Inorg. Chem., 2021, 60, 14944–14951 CrossRef CAS PubMed.
  59. M. Dai, Z. Fu, Z. Wang and H. Xu, Chem. Eng. J., 2023, 452, 139133 CrossRef CAS.
  60. L. Li, P. Yang, W. Xia, Y. Wang, F. Ling, Z. Cao, S. Jiang, G. Xiang, X. Zhou and Y. Wang, Ceram. Int., 2021, 47, 769–775 CrossRef CAS.
  61. H. Zhou, W. Gao, P. Cai, B. Zhang and S. Li, Solid State Sci., 2020, 104, 106283 CrossRef CAS.
  62. S. Wang, S. Ma, J. Wu, Z. Ye and X. Cheng, Chem. Eng. J., 2020, 393, 124564 CrossRef CAS.
  63. H. Zhang, Z. Y. Gao and G. G. Li, Chem. Eng. J., 2020, 380, 122491 CrossRef CAS.
  64. J. W. Li, J. W. Wang and R. S. Lei, J. Lumin., 2022, 249, 119047 CrossRef CAS.
  65. L. P. Li, F. Qin and L. Li, J. Lumin., 2019, 211, 258–263 CrossRef CAS.
  66. H. Zhang, H. Yang, G. Li, S. Liu, H. Li, Y. Gong, Y. Liang and Y. Chen, Chem. Eng. J., 2020, 396, 125251 CrossRef CAS.
  67. Y. Pan, N. Guo, L. Wang, J. Li, W. Lü and Y. Miao, ACS Appl. Electron. Mater., 2020, 2, 10–3426,  DOI:10.1021/acsaelm.0c00697.
  68. I. Kachou, M. Dammak, S. Auguste, F. Amiard and P. Daniel, Dalton Trans., 2023, 52, 18233–18246 RSC.
  69. Y. Wu, H. Suo, X. Zhao, Z. Zhou and C. Guo, Inorg. Chem. Front., 2018, 5, 2456–2464 RSC.
  70. Y. F. Wu, H. Suo and X. Q. Zhao, Inorg. Chem. Front., 2018, 5, 2456–2461 RSC.
  71. M. Fhoula, K. Saidi, C. Hernandez-Álvarez, K. S. Carracedo, M. Dammak and I. R. Martín, J. Alloys Compd., 2024, 979, 173537 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.