Next Article in Journal
Poly(sodium acrylate)-Modified Magnetite Nanoparticles for Separation of Heavy Metals from Aqueous Solutions
Previous Article in Journal
Excipients Used for Modified Nasal Drug Delivery: A Mini-Review of the Recent Advances
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Luminescence of Mn4+ in a Zero-Dimensional Organic–Inorganic Hybrid Phosphor [N(CH3)4]2ZrF6 for Dual-Mode Temperature Sensing

School of Chemistry and Chemical Engineering, Weifang University, Weifang 261061, China
*
Author to whom correspondence should be addressed.
Materials 2022, 15(19), 6543; https://doi.org/10.3390/ma15196543
Submission received: 25 August 2022 / Revised: 15 September 2022 / Accepted: 17 September 2022 / Published: 21 September 2022
(This article belongs to the Section Materials Chemistry)

Abstract

:
Searching for new low-dimensional organic–inorganic hybrid phosphors is of great significance due to their unique optical properties and wide applications in the optoelectronic field. In this work, we report a Mn4+ doped zero-dimensional organic–inorganic hybrid phosphor [N(CH3)4]2ZrF6, which was synthesized by a wet chemical method. The crystal structure, thermal stability, and optical properties were systemically investigated by means of XRD, SEM, TG-DTA, FTIR, DRS, emission spectra, excitation spectra, as well as decay curves. Narrow red emission with high color purity can be observed from [N(CH3)4]2ZrF6:Mn4+ phosphor, which maintains effective emission intensity even at room temperature, indicating its potential practical application in WLEDs. In the temperature range of 13–295 K, anti-Stokes and Stokes sidebands of Mn4+ ions exhibit different temperature responses. By applying the emission intensity ratio of anti-Stokes vs. Stokes sidebands as temperature readout, an optical thermometer with a maximum absolute sensitivity of 2.13% K−1 and relative sensitivity of 2.47% K−1 can be obtained. Meanwhile, the lifetime Mn4+ ions can also be used for temperature sensing with a maximum relative sensitivity of 0.41% K−1, demonstrating its potential application in optical thermometry.

1. Introduction

In recent years, luminescent materials have attracted special attention for their applications in optoelectronics, especially in the fields of temperature sensing [1,2,3]. Traditional thermometers are based on the expansion properties of liquids or metals, requiring contact with objects and heat exchange, thus leading to errors between the measured temperature and the actual temperature [4,5]. Therefore, temperature sensors based on luminescent materials are in high demand due to their high accuracy and non-invasive properties. Optical thermometry can be performed using a number of techniques, such as emission intensity, decay lifetime, band width, peak shift, and fluorescence intensity ratio (FIR). Among them, FIR and decay lifetime techniques have a self-reference character to remove the disturbance of excitation power or detector stability. The traditional FIR technique is based on the intensity ratio between two emissions originating from two thermally coupled energy levels of rare earth ions in up-conversion processes, such as Er3+ (2H11/2/4S3/2), Ho3+ (5S2/5F4), Tm3+ (3F2,3/3H4), and so on [6,7,8]. However, such phosphors have some defects, such as large overlap of emission peaks of thermally coupled energy levels, especially in the high temperature range, which reduces the measurement accuracy. Furthermore, the absorption of rare earths in the near infrared region is weak, and the quantum efficiency is low [9,10,11,12,13,14,15]. Therefore, it is necessary to develop a new optical thermometer with the advantages of high peak resolution, low cost, and effective emission.
Mn4+ doped fluoride phosphors have attracted much attention due to their low production cost and wide absorption from the ultraviolet to blue region [16,17]. The Mn4+ ion has been regarded as one of the most promising activators for narrow red emission, because its emission locates at about 630 nm in fluoride crystal fields. Meanwhile, the narrow emission of Mn4+ consists of some vibronic transitions due to an 2Eg4A2g transition, which can be regarded as thermally coupled vibronic levels. Hence, Mn4+ doped fluoride phosphor is an excellent candidate for applications in fluorescence thermometers [18,19,20,21,22,23,24]. Recently, zero-dimensional organic inorganic phosphors have attracted unparalleled interest owing to their unique crystallographic structures with fascinating optical characteristics [25]. The organic ingredients expand the atomic distances between isolated luminescent centers, expanding the 0D family and leading to unprecedented luminescence properties. Although tunable broadband emission can be obtained from 0D metal halides, there are few reports on narrow band emission in 0D systems. Recently, T. Jüstel et al. reported narrow red emission in organic–inorganic [C(NH2)3]2GeF6:Mn4+ phosphor [26]. However, the narrow red emission of the Mn4+ in this material could only be observed at cryogenic temperature due to the multi-phonon relaxation. P. Cai et al. reported an organic–inorganic [N(CH3)4]2TiF6:Mn4+ phosphor, whose narrow red emission of Mn4+ can be observed even at room temperature [27]. Inspired by this, we synthesized organic–inorganic [N(CH3)4]2ZrF6:Mn4+ phosphor directionally. Accordingly, luminescence properties of 0D metal halides mainly depend on the intrinsic photophysical properties of their inorganic building blocks. Therefore, the substitution of Ti by Zr would exhibit excellent photoelectric properties.
In the present work, we adopt a wet chemical method to obtain a zero-dimensional organic–inorganic hybrid phosphor, [N(CH3)4]2ZrF6:Mn4+, which has been synthesized by combining hydrothermal and solid state methods. The photoluminescence property as well as the dynamic decay of [N(CH3)4]2ZrF6:Mn4+ phosphor have been investigated in a wide temperature range from 13 to 295 K.

2. Materials and Methods

2.1. Sample Preparation

[N(CH3)4]2ZrF6: Mn4+ powder samples were successfully synthesized by a wet chemical method [27]. K2MnF6 powders were used for the Mn4+ source and prepared according to the published literature [10,28]. Pure [N(CH3)4]2ZrF6 powder was prepared by hydrothermal method [29]. Practically, 0.493 g ZrO2, 0.219 g [N(CH3)4]Cl, 10.0 mL DMF, and 1.25 mL HF (48%) were mixed and stirred for 30 min in a Teflon reactor. Then, the mixture solution was heated at 180 °C for 48 h in an oven. Natural cooling to room temperature, the crystalline sample was collected by filtration. Washed with DMF for three times, the product was dried at 100 °C for 12 h. Mn4+ doped [N(CH3)4]2ZrF6 was obtained by facile solid-state method. Appropriate [N(CH3)4]2ZrF6 and K2MnF6 in stoichiometric (1% mol) were well weighted and ground thoroughly in an agate mortar. Then, the mixture was heated in an oven at 80 °C for 24 h.

2.2. Characterization

The crystalline information of the obtained products was identified by means of X-ray diffraction (XRD, Philips X’Pert MPD, Almelo, The Netherlands). The morphology was evaluated by JSM-6700F field emission scanning electron microscopy (SEM). Thermal analyses were performed on a Shimadzu (Japan) thermal analyzer with a heating rate of 10 °C/min from room temperature to 800 °C in N2 atmosphere (30 mL/min). Fourier transformed infrared (FTIR) spectrum was measured by using a FT/IR-4100 (JASCO) spectrometer in ATR mode with sample powder. The quantum yield (QY) of the obtained sample was measured by the Edinburgh Instruments FLS1000 equipped with an integrating sphere attachment. The concentration of Mn4+ in the sample was analyzed by inductively coupled plasma–mass spectrometry (ICP-MS) on a Perkin Elmer (NexION 300D). The diffuse reflectance spectrum (DRS) was recorded by a V-670 (JASCO) spectrophotometer. The photoluminescence excitation (PLE) spectrum was recorded by a QuantaMaster 300 (PTI) high power Xenon flash lamp. The photoluminescence (PL) spectra and decay lifetimes were recorded by using a 355 nm pulsed Nd:YAG laser (Spectron Laser Sys. SL802G) with a pulse width of 5 ns and a repetition rate of 10 Hz. The luminescence is collected by the 75 cm monochromator (ActonResearch Corp. Pro-750, Acton, MA, USA) and multiplied by the PMT (Hamamatsu R928, Hamamatsu Electronic Press Co., Ltd., Shizuoka, Japan). The temperature was controlled by a liquid helium flow cryostat.

3. Results

Figure 1a shows the XRD patterns of the obtained [N(CH3)4]2ZrF6:1%Mn4+ sample. All the peaks match well with the standard [N(CH3)4]2TiF6 card, and no trace of impurity phase can be found [10,29], indicating that the pure [N(CH3)4]2ZrF6 phase was obtained. By the Williamson–Hall plot, the crystalline grain size of 76 nm and micro strain of 0.00467 were obtained. The morphology of the obtained phosphor is displayed in Figure 1b. The particles are in the size of ten to twenty micrometers with a smooth surface, and most of them are clustered into larger particles.
For the purpose of confirming the thermal stability of [N(CH3)4]2ZrF6:1%Mn4+, TG-DTA curves were measured in N2 atmosphere, as shown in Figure 2. Four endothermic peaks can be observed. A broad endothermic peak at about 100 °C coincided with a 5% mass loss, which could be a loss of water adsorbed to the particle surface. The second endothermic peak occurs at 270 °C and is comparatively weak and broad, coinciding with a 10% mass loss. Since phase transition is a transition without weight loss, we tentatively assigned the peak at 270 °C as the thermal decomposition of a portion of the [N(CH3)4] group [26]. In addition, an obvious mass loss occurs at about 420 °C, coinciding with the endothermic peak, which corresponds to the decomposition of the sample and indicates a good thermal stability of [N(CH3)4]2ZrF6:1%Mn4+. The nature of the endothermic peak at 650 °C without mass loss is not known [27].
Figure 3 shows the FTIR spectrum of the synthesized [N(CH3)4]2ZrF6:1%Mn4+ powder. The absorption bands correspond to the vibrations of molecular groups. According to the literature, those absorption bands are marked in the spectrum. The C–N stretching vibrations can be observed at 956 and 1023 cm−1. The peaks located at 1472, 1598, 2830, and 3024 cm−1 can be ascribed to the C–H vibrations in the CH3 group [29,30]. The peak at 3460 cm−1 corresponds to the H2O molecular vibration [31].
Figure 4a shows the DRS of the pure and Mn4+ doped [N(CH3)4]2ZrF6 samples at room temperature [32,33]. In the spectrum of undoped powder, a band at about 200 nm can be found. Upon doping with Mn4+, two new absorption bands at 250 and 360 nm appear, which correspond to the F (2p) → Mn4+ (3d) charge transfer (CT) and 4A24T1 transitions, respectively. This indicates that Mn4+ ions have been introduced into the host lattice successfully. In addition, the concentration of Mn4+ in the sample was analyzed by ICP-MS, and the concentration of Mn4+ (Mn:Zr = 2494:586166 μg/L) is verified to be 0.71 mol% relative to Zr. By using the Kubelka–Munk equation, the energy gap of the phosphors can be evaluated based on the transformation of DRS, as shown in Figure 4b. From the figure, the energy gap of the pure sample is about 5.41eV, while the introduction of Mn4+ reduces the band gap to 4.87 eV, of which the energy level of Mn4+ is between conduction and the valence band.
Figure 5a shows the normalized room temperature PLE and PL spectra of the as-synthesized [N(CH3)4]2ZrF6:1%Mn4+ phosphor. In the PLE spectrum, three excitation bands can be observed in the wavelength range from 200 to 500 nm by monitoring at 630 nm emission. The band located at 253 nm is due to the CT band, and the latter two broad bands at 360 nm (4A24T1) and 475 nm (4A24T2) correspond to the spin-allowed transitions of Mn4+ ions. In the PL spectrum, the narrow red emission lines can be found in the range of 580–660 nm, owing to the spin-forbidden 2Eg4A2g transition of Mn4+ in the octahedral field. According to the research of Ok et al., [N(CH3)4]2TiF6 crystallizes in the centrosymmetric structure with zero-dimension, in which the TiF6 group is separated by organic cations [34]. According to S. Adachi’s summary, negligible ZPL can be found when Mn4+ is located in a higher crystal symmetry [1]. Thus, there is no obvious zero phonon line (ZPL) in the PL spectrum of [N(CH3)4]2ZrF6:Mn4+, which is similar to that of [N(CH3)4]2TiF6:Mn4+ [27]. However, phonon-activated sidebands can be found in either side of the ZPL. The emission peaks at 599, 609, and 613 nm can be assigned to the v3, v4, and v6 anti-Stokes vibrational modes, while the emission peaks at 630, 633, and 647 nm are due to the Stokes v6, v4, and v3 vibrational modes, respectively. Figure 5b shows the PLE and PL spectra with abscissa expressed in wavenumbers. Efficient red luminescence from [N(CH3)4]2ZrF6:Mn4+ phosphor could be observed by the UV light illumination, and the QY value of the sample is about 31%, as shown in Figure S1.
As is known, Mn4+ is a high field ion with 3d3 configuration, and its luminescence property is highly related to the crystal environment, such as crystal strength, site symmetry, covalency and nephelauxetic effect [35]. Generally, the energy level scheme of Mn4+ in the host can be expressed by the Tanabe–Sugano diagram [36], as shown in Figure 6. Dq stands for the local crystal field strength, which can be obtained from the mean peak energy (21,052 cm−1) of the 4A2g4T2g transition by using the following equation [37]:
D q = E ( A 4 2 g T 4 2 g ) 10
The Racah parameter B can be determined by the equation:
D q B = 15 ( x 8 ) x 2 10 x
where the parameter x can be defined as:
x = E ( A 4 2 g T 4 1 g ) E ( A 4 2 g T 4 2 g ) D q
On the basis of the peak energy (15,873 cm−1) of 2Eg4A2g transition obtained from the emission spectrum, the Racah parameter C can be evaluated from the expression:
E ( E 2 g A 4 2 g ) B = 3.05 C B + 7.9 1.8 B D q
Based on Equations (1)–(4), the crystal field parameters of Dq, B, C were determined to be 2105, 638 and 3660 cm−1, respectively.
To investigate the luminescence properties of [N(CH3)4]2TiF6:Mn4+ at low temperature, the temperature-dependent emission spectra were recorded in the temperature range from 13 to 295 K, as displayed in Figure 7. At 13 K, only ZPL and Stokes sidebands can be observed. With the increase in temperature, electrons have more opportunity to populate at a higher vibrational state. Therefore, anti-Stokes sidebands located at a higher energy side of ZPL start to appear at 100 K. With rising temperature, both of the emission intensities of anti-Stokes and Stokes sidebands increase. Figure 8a shows the integral emission intensities of anti-Stokes sidebands, Stokes sidebands, and total Mn4+ emission as a relation with temperature. It can be observed that all the emission intensities increase with increasing temperature from 100 to 295 K. According to the thermally coupled energy levels (TCL) theory of trivalent rare earth ions, the adjacent levels with energy gaps between 100–2000 cm−1 can be thermally populated. Therefore, the phonon-coupled anti-Stokes and Stokes sidebands could be considered as two adjacent levels. According to Boltzmann’s law, the emission intensity ratio (R) from thermally coupled energy levels can be expressed as:
R = A e x p ( E k T )
By transforming Equation (5), the emission intensity ratio (R) of anti-Stokes vs. Stokes sidebands of Mn4+ can be fitted by the formula:
L n R = A T + B
where A is a constant, and B is a correction term for the comprehensive population of thermally coupled energy levels [38]. Figure 8b shows the experimental data fitted by using Equation (6), and the values of A and B were obtained to be 247.2 and −0.028, respectively. Sensitivity is an important parameter to evaluate the suitability for thermometry. The absolute sensitivity (Sa) and relative sensitivity (Sr) can be defined as:
S a = dR dT = A T 2 e BT A T
S r = 1 R dR dT = A T 2
Figure 8c,d exhibit the calculated sensitivities by using Equations (7) and (8). The absolute sensitivity increases first with rising temperature, reaching a maximum of 2.13% K−1 at 119 K and then decreasing to 0.119% K−1 at 295 K. As for the relative sensitivity, it decreases with increasing temperature from 2.47% K−1 at 100 K to 0.28% K−1 at 295 K. The obtained sensitivities are comparable to recently reported phosphors, Refs. [39,40] indicating its potential application in the FIR technique based on optical temperature sensing.
However, the dynamic lifetime decay of the emission shows different temperature dependence from emission intensity. Figure 9a shows the decay curves of a Mn4+ doped [N(CH3)4]2ZrF6 phosphor in the temperature range from 13 to 295 K. With increasing temperature, the decay curves always keep single exponential characteristics, and lifetimes become shorter, indicating the enhancement of a non-radiative relaxation process with rising temperature. According to [1,27,41,42], the excited dynamic of Mn4+ in fluoride phosphors is commonly related to activation energy, spin-orbit interaction and vibration modes. Therefore, the kinetic process of the 2Eg4A2g emission was complicated in the whole measured temperature range, and it is difficult to establish a simple non-radiative transition theoretical model to describe the temperature-dependent decay behavior of Mn4+ doped phosphors. In this case, the average lifetime was adopted to evaluate the values, which can be calculated by the equation [43]:
I ( t ) = A 1 exp ( t τ 1 ) + A 2 exp ( t τ 2 )
τ a v e = A 1 × τ 1 2 + A 2 × τ 2 2 A 1 × τ 1 + A 2 × τ 2
where I(t) is the intensity at time t, A1 and A2 represent the pre-exponential factors, τ1 and τ2 stand for the fast decay lifetime and slow decay lifetime, respectively, and τave is the average lifetime. The calculated lifetime values at various temperatures are listed in Figure 9b. The values decrease from 7.649 to 3.634 ms as temperature increases from 13 to 295 K.
For practicality and conciseness, the temperature dependence of the decay lifetime can be well fitted based on the Mott−Seitz model, as follows [44]:
τ 0 τ T = 1 + A exp ( E a k B T )
where τ0 is the decay lifetime at 13 K, τT is the decay lifetime at temperature T, A′ is an empirical pre-exponential factor, Ea is the activation energy, and kB is the Boltzmann constant. Here, in order to be more practical, τ0 is regarded as a constant, and the fitting equation is transformed to:
1 τ T = B + C exp ( D T )
where B′, C′, and D′ are the empirical parameters related to τ0, A′ and Ea. After fitting the experimental data by using Equation (12) (Figure 9c), the values of B′, C′, and D′ are determined to be 132.7, 465.7, and −342.0, respectively. Therefore, the relative sensitivity (Sr) can be calculated from:
S r = | τ τ T | × 100 % = D T 2 τ C exp ( D T ) × 100 %
The obtained Sr is shown in Figure 9d, which increases firstly with rising temperature, reaching a maximum of 0.41% K−1 at 133 K, and then decreasing. Relying on the effective FIR technique in the temperature range from 100 to 295 K and lifetime thermometry in the temperature range from 13 to 295 K, the unique application potential of [N(CH3)4]2ZrF6:1%Mn4+ phosphor in the low temperature range can be forecasted.

4. Conclusions

A novel Mn4+ doped zero-dimensional organic–inorganic hybrid [N(CH3)4]2ZrF6 phosphor was successfully synthesized by a wet chemical method for the first time. The photoluminescence from the [N(CH3)4]2ZrF6:Mn4+ phosphor exhibits sharp lines and can be observed even at room temperature. The phosphor can be effectively excited by NUV and blue light. Temperature-dependent emission spectra and decay curves were investigated to clarify the enhanced emission intensity and the shortened lifetime, which are due to the phonon-assistant emission and enhanced non-radiative transition during the temperature-raising process. The emission intensity ratio of anti-Stokes and Stokes sidebands can be used for temperature sensing with a maximum absolute sensitivity of 2.13% K−1 and relative sensitivity of 2.47% K−1. In addition, the decay lifetime of Mn4+ can be used as a thermometry, demonstrating its potential application in a dual-mode temperature sensor.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15196543/s1, Figure S1. The quantum yield of the [N(CH3)4]2ZrF6:1%Mn4+ sample. Figure S2. The fitting of temperature-dependent decay curves. The decay curves are recorded by monitoring at 630 nm with excitation at (a) 13 K, (b) 50 K, (c) 100 K, (d) 150 K, (e) 200 K, (f) 250 K, and (g) 295 K.

Author Contributions

Conceptualization, J.W. and M.S.; Data curation, J.L. and Y.W.; Funding acquisition, M.S.; Investigation, Y.W.; Methodology, J.W. and J.L.; Software, Y.W.; Supervision, J.W.; Validation, M.S.; Writing—original draft, J.W.; Writing—review and editing, M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Shandong Provincial Natural Science Foundation (grant number ZR2019QB011) and by the Project of Shandong Province Higher Educational Science and Technology Program (grant number J18KA081).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available from the corresponding authors upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Adachi, S. Review—Mn4+-Activated Red and Deep Red-Emitting Phosphors. ECS J. Solid State Sci. Technol. 2019, 9, 016001. [Google Scholar] [CrossRef]
  2. Cao, R.; Lv, X.; Ran, Y.; Xu, L.; Chen, T.; Guo, S.; Ao, H.; Yu, X. Rare-earth-free Li5La3Ta2O12:Mn4+ deep-red-emitting phosphor: Synthesis and photoluminescence properties. J. Am. Ceram. Soc. 2019, 102, 5910–5918. [Google Scholar] [CrossRef]
  3. Xue, J.; Noh, H.M.; Choi, B.C.; Park, S.H.; Du, P. 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. 2019, 382, 122861. [Google Scholar] [CrossRef]
  4. Zhou, Y.; Song, E.; Deng, T.; Wang, Y.; Xia, Z.; Zhang, Q. Surface Passivation toward Highly Stable Mn4+-Activated Red-Emitting Fluoride Phosphors and Enhanced Photostability for White LEDs. Adv. Mater. Interfaces 2019, 6, 1802006. [Google Scholar] [CrossRef]
  5. Qin, L.; Bi, S.; Cai, P.; Chen, C.; Wang, J.; Kim, S.I.; Huang, Y.; Seo, H.J. Preparation, characterization and luminescent properties of red-emitting phosphor: LiLa2NbO6 doped with Mn4+ ions. J. Alloys Compd. 2018, 755, 61–66. [Google Scholar] [CrossRef]
  6. Ming, H.; Liu, L.; He, S.; Peng, J.; Du, F.; Fu, J.; Yang, F.; Ye, X. An ultra-high yield of spherical K2NaScF6:Mn4+ red phosphor and its application in ultra-wide color gamut liquid crystal displays. J. Mater. Chem. C 2019, 7, 7237–7248. [Google Scholar] [CrossRef]
  7. Zhou, Y.; Song, E.; Deng, T.T.; Zhang, Q. Waterproof Narrow-Band Fluoride Red Phosphor K2TiF6:Mn4+ via Facile Superhydrophobic Surface Modification. ACS Appl. Mater. Interfaces 2018, 10, 880–889. [Google Scholar] [CrossRef]
  8. Adachi, S.; Takahashi, T. Direct synthesis and properties of K2SiF6:Mn4+ phosphor by wet chemical etching of Si wafer. J. Appl. Phys. 2008, 104, 023512. [Google Scholar] [CrossRef]
  9. Song, E.; Wang, J.; Shi, J.; Deng, T.; Zhang, Q. Highly Efficient and Thermally Stable K3AlF6:Mn4+ as a Red Phosphor for Ultra-High-Performance Warm White Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2017, 9, 8805–8812. [Google Scholar] [CrossRef]
  10. Jin, Y.; Fang, M.H.; Grinberg, M.; Mahlik, S.; Lesniewski, T.; Brik, M.G.; Luo, G.Y.; Lin, J.G.; Liu, R.S. Narrow Red Emission Band Fluoride Phosphor KNaSiF6:Mn4+ for Warm White Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2016, 8, 11194–11203. [Google Scholar] [CrossRef]
  11. Wu, W.L.; Fang, M.H.; Zhou, W.; Lesniewski, T.; Liu, R. High Color Rendering Index of Rb2GeF6:Mn4+ for Light-emitting Diodes. Chem. Mater. 2017, 29, 935–939. [Google Scholar] [CrossRef]
  12. Xi, L.; Pan, Y.; Chen, X.; Huang, S.; Wu, M. Optimized photoluminescence of red phosphor Na2SnF6:Mn4+ as red phosphor in the application in “warm” white LEDs. J. Am. Ceram. Soc. 2017, 100, 2005–2015. [Google Scholar] [CrossRef]
  13. Zhu, H.; Lin, C.C.; Luo, W.; Shu, S.; Liu, Z.; Liu, Y.; Kong, J.; Ma, E.; Cao, Y.; Liu, R.; et al. Highly efficient non-rare-earth red emitting phosphor for warm white light-emitting diodes. Nat. Commun. 2014, 5, 4312. [Google Scholar] [CrossRef] [PubMed]
  14. He, S.; Xu, F.; Han, T.; Lu, Z.; Wang, W.; Peng, J.; Du, F.; Yang, F.; Ye, X. A Mn4+-doped oxyfluoride phosphor with remarkable negative thermal quenching and high color stability for warm WLEDs. Chem. Eng. J. 2020, 392, 123657. [Google Scholar] [CrossRef]
  15. Cai, P.; Qin, L.; Chen, C.; Wang, J.; Seo, H.J. Luminescence, energy transfer and optical thermometry of a novel narrow red emitting phosphor: Cs2WO2F4:Mn4+. Dalton Trans. 2017, 46, 14331–14340. [Google Scholar] [CrossRef] [PubMed]
  16. Kesarwani., V.; Rai, V.K. Fluorescence intensity ratio technique and its reliability. Methods Appl. Fluoresc. 2022, 10, 034006. [Google Scholar] [CrossRef]
  17. Dong, L.; Zhang, L.; Jia, Y.; Xu, Y.; Yin, S.; You, H. ZnGa2-yAlyO4:Mn2+,Mn4+ Thermochromic Phosphors: Valence State Control and Optical Temperature Sensing. Inorg Chem. 2020, 59, 15969–15976. [Google Scholar] [CrossRef]
  18. Prasad, M.; Rai, V.K. Influence of transition metal ions: Broad band upconversion emission in thermally stable phosphors. J. Alloys Compd. 2020, 837, 155289. [Google Scholar] [CrossRef]
  19. Zhu, Y.; Li, C.; Deng, D.; Chen, B.; Yu, H.; Li, H.; Wang, L.; Shen, C.; Jing, X.; Xu, S. A high-sensitivity dual-mode optical thermometry based on one-step synthesis of Mn2+:BaAl12O19–Mn4+:SrAl12O19 solid solution phosphors. J. Alloys Compd. 2021, 853, 157262. [Google Scholar] [CrossRef]
  20. Xue, J.; Yu, Z.; Noh, H.M.; Lee, B.R.; Choi, B.C.; Park, S.H.; Jeong, J.H.; Du, P.; Song, M. Designing multi-mode optical thermometers via the thermochromic LaNbO4:Bi3+/Ln3+ (Ln = Eu, Tb, Dy, Sm) phosphors. Chem. Eng. J. 2021, 415, 128977. [Google Scholar] [CrossRef]
  21. Xue, J.; Li, F.; Liu, F.; Noh, H.M.; Lee, B.R.; Choi, B.C.; Park, S.H.; Jeong, J.H.; Du, P. Designing ultra-highly efficient Mn2+-activated Zn2GeO4 green-emitting persistent phosphors toward versatile applications. Mater. Today Chem. 2022, 23, 100693. [Google Scholar] [CrossRef]
  22. Xue, J.; Noh, H.M.; Park, S.H.; Lee, B.R.; Kim, J.H.; Jeong, J.H. NUV light induced visible emission in Er3+-activated NaSrLa(MoO4)O3 phosphors for green LEDs and thermometer. J. Am. Ceram. Soc. 2019, 103, 1174–1186. [Google Scholar] [CrossRef]
  23. Wang, X.; Liu, Q.; Bu, Y.; Liu, C.-S.; Liu, T.; Yan, X. Optical temperature sensing of rare-earth ion doped phosphors. RSC Adv. 2015, 5, 86219–86236. [Google Scholar] [CrossRef]
  24. Cai, P.; Wang, X.; Seo, H.J. Excitation power dependent optical temperature behaviors in Mn4+ doped oxyfluoride Na2WO2F4. Phys. Chem. Chem. Phys. 2018, 20, 2028–2035. [Google Scholar] [CrossRef] [PubMed]
  25. Li, M.; Xia, Z. Recent progress of zero-dimensional luminescent metal halides. Chem. Soc. Rev. 2021, 50, 2626–2662. [Google Scholar] [CrossRef] [PubMed]
  26. Baur, F.; Böhnisch, D.; Jüstel, T. Luminescence of Mn4+ in a Hexafluorogermanate with the Complex Organic Cation Guanidinium [C(NH2)3]2GeF6:Mn4+. ECS J. Solid State Sci. Technol. 2020, 9, 046003. [Google Scholar] [CrossRef]
  27. Cai, P.; Wang, S.; Xu, T.; Tang, Y.; Yuan, X.; Wan, M.; Ai, Q.; Si, J.; Yao, X.; Cao, Y.; et al. Mn4+ doped zero-dimensional organic-inorganic hybrid material with narrow-red emission. J. Lumin. 2020, 228, 117661. [Google Scholar] [CrossRef]
  28. Bode, H.; Jenssen, H.; Bandte, F. Über eine neue Darstellung des Kalium-hexafluoromanganats(IV). Angew. Chem. 1953, 65, 304–304. [Google Scholar] [CrossRef]
  29. Kim, E.A.; Dong, W.L.; Kang, M.O. Centrosymmetric [N(CH3)4]2TiF6 vs. noncentrosymmetric polar [C(NH2)3]2TiF6: A hydrogen-bonding effect on the out-of-center distortion of TiF6 octahedra. J. Solid State Chem. 2012, 195, 149–154. [Google Scholar] [CrossRef]
  30. Young, C.W.; Koehler, J.S.; Mckinney, D.S. Infrared Absorption Spectra of Tetramethyl Compounds. J. Am. Chem. Soc. 1947, 69, 1410–1415. [Google Scholar] [CrossRef]
  31. Voit, E.I.; Davidovich, R.L.; Udovenko, A.A.; Logvinova, V.B. The Structure and Vibrational Spectra of Gallium(III) Fluoride Dimeric Complex with Tetramethylammonium Cation. Opt. Spectrosc. 2019, 127, 984–990. [Google Scholar] [CrossRef]
  32. Brik, M.G.; Srivastava, A.M. Ab Initio Studies of the Structural, Electronic, and Optical Properties of K2SiF6 Single Crystals at Ambient and Elevated Hydrostatic Pressure. J. Electrochem. Soc. 2012, 159, J212. [Google Scholar] [CrossRef]
  33. Kasa, R.; Adachi, S. Red and Deep Red Emissions from Cubic K2SiF6:Mn4+ and Hexagonal K2MnF6 Synthesized in HF/KMnO4/KHF2/Si Solutions. J. Electrochem. Soc. 2012, 159, J89. [Google Scholar] [CrossRef]
  34. Ok, K.M. Toward the Rational Design of Novel Noncentrosymmetric Materials: Factors Influencing the Framework Structures Acc. Chem. Res. 2016, 49, 2774–2785. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, D.; Zhou, Y.; Xu, W.; Zhong, J.; Ji, Z.; Xiang, W. Enhanced luminescence of Mn4+:Y3Al5O12red phosphor via impurity doping. J. Mater. Chem. C 2016, 4, 1704–1712. [Google Scholar] [CrossRef]
  36. Zhong, J.; Chen, D.; Chen, X.; Wang, K.; Li, X.; Zhu, Y.; Ji, Z. Efficient rare-earth free red-emitting Ca2YSbO6:Mn4+, M(M = Li+, Na+, K+, Mg2+) phosphors for white light-emitting diodes. Dalton Trans. 2018, 47, 6528–6537. [Google Scholar] [CrossRef] [PubMed]
  37. Sijbom, H.F.; Verstraete, R.; Joos, J.J.; Poelman, D.; Smet, P.F. K2SiF6:Mn4+ as a red phosphor for displays and warm-white LEDs: A review of properties and perspectives. Opt. Mater. Express 2017, 7, 3332–3365. [Google Scholar] [CrossRef]
  38. Wang, X.; Wang, Y.; Marques-Hueso, J.; Yan, X. Improving Optical Temperature Sensing Performance of Er3+ Doped Y2O3 Microtubes via Co-doping and Controlling Excitation Power. Sci. Rep. 2017, 7, 758. [Google Scholar] [CrossRef]
  39. Fan, H.; Lu, Z.; Meng, Y.; Chen, P.; Zhou, L.; Zhao, J.; He, X. Optical temperature sensor with superior sensitivity based on Ca2LaSbO6: Mn4+, Eu3+ phosphor. Opt. Laser Technol. 2022, 148, 107804. [Google Scholar] [CrossRef]
  40. Fang, Y.; Zhang, Y.; Zhang, Y.; Hu, J. Achieving high thermal sensitivity from ratiometric CaGdAlO4:Mn4+, Tb3+ thermometers. Dalton Trans. 2021, 50, 13447–13458. [Google Scholar] [CrossRef]
  41. Senden, T.; Dijk-Moes, R.J.A.v.; Meijerink, A. Quenching of the red Mn4+ luminescence in Mn4+-doped fluoride LED phosphors. Light Sci. Appl. 2018, 007, 74–86. [Google Scholar] [CrossRef] [PubMed]
  42. Grinberg, M.; Lesniewski, T.; Mahlik, S.; Liu, R.S. 3d3 system Comparison of Mn4+ and Cr3+ in different lattices. Opt. Mater. 2017, 74, 93–100. [Google Scholar] [CrossRef]
  43. Du, P.; Luo, L.; Li, W.; Yan, F.; Xing, G. Multi-site occupancies and photoluminescence characteristics in developed Eu2+-activated Ba5SiO4Cl6 bifunctional platform: Towards manufacturable optical thermometer and indoor illumination. J. Alloys Compd. 2020, 826, 154233. [Google Scholar] [CrossRef]
  44. Chen, Y.; He, J.; Zhang, X.; Rong, M.; Xia, Z.; Wang, J.; Liu, Z.Q. Dual-Mode Optical Thermometry Design in Lu3Al5O12:Ce3+/Mn4+ Phosphor. Inorg. Chem. 2020, 59, 1383–1392. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD patterns of [N(CH3)4]2ZrF6:1%Mn4+ sample (top) at room temperature, and a standard [N(CH3)4]2TiF6 card (bottom) is given as a reference. (b) SEM image of [N(CH3)4]2ZrF6:1%Mn4+ sample.
Figure 1. (a) XRD patterns of [N(CH3)4]2ZrF6:1%Mn4+ sample (top) at room temperature, and a standard [N(CH3)4]2TiF6 card (bottom) is given as a reference. (b) SEM image of [N(CH3)4]2ZrF6:1%Mn4+ sample.
Materials 15 06543 g001
Figure 2. TG (blue) and DTA (red) curves of the [N(CH3)4]2ZrF6:1%Mn4+ sample recorded under N2 atmosphere.
Figure 2. TG (blue) and DTA (red) curves of the [N(CH3)4]2ZrF6:1%Mn4+ sample recorded under N2 atmosphere.
Materials 15 06543 g002
Figure 3. FTIR spectrum of [N(CH3)4]2ZrF6:1%Mn4+ sample at room temperature.
Figure 3. FTIR spectrum of [N(CH3)4]2ZrF6:1%Mn4+ sample at room temperature.
Materials 15 06543 g003
Figure 4. (a) Diffuse reflectance spectrum and (b) energy band gap of pure and Mn4+ doped [N(CH3)4]2ZrF6 samples at room temperature.
Figure 4. (a) Diffuse reflectance spectrum and (b) energy band gap of pure and Mn4+ doped [N(CH3)4]2ZrF6 samples at room temperature.
Materials 15 06543 g004
Figure 5. (a) PLE spectrum by monitoring at 630 nm, and PL spectrum with excitation of 355 nm of the [N(CH3)4]2ZrF6:1%Mn4+ sample at room temperature. (b) PLE and PL spectra with abscissa expressed in wavenumbers.
Figure 5. (a) PLE spectrum by monitoring at 630 nm, and PL spectrum with excitation of 355 nm of the [N(CH3)4]2ZrF6:1%Mn4+ sample at room temperature. (b) PLE and PL spectra with abscissa expressed in wavenumbers.
Materials 15 06543 g005
Figure 6. Tanabe–Sugano energy-level diagram of Mn4+ in [N(CH3)4]2ZrF6.
Figure 6. Tanabe–Sugano energy-level diagram of Mn4+ in [N(CH3)4]2ZrF6.
Materials 15 06543 g006
Figure 7. Temperature-dependent PL spectra of [N(CH3)4]2ZrF6:1%Mn4+ in the range from 13 to 295 K.
Figure 7. Temperature-dependent PL spectra of [N(CH3)4]2ZrF6:1%Mn4+ in the range from 13 to 295 K.
Materials 15 06543 g007
Figure 8. (a) Integral emission intensities of anti-Stokes sidebands (blue), Stokes sidebands (red), and total Mn4+ emission (black) as a relation with temperature from 13 to 295 K; (b) emission intensity ratio of anti-Stokes vs. Stokes sidebands (R) in a log relation with 1/T; (c) calculated absolute sensitivity; (d) calculated relative sensitivity.
Figure 8. (a) Integral emission intensities of anti-Stokes sidebands (blue), Stokes sidebands (red), and total Mn4+ emission (black) as a relation with temperature from 13 to 295 K; (b) emission intensity ratio of anti-Stokes vs. Stokes sidebands (R) in a log relation with 1/T; (c) calculated absolute sensitivity; (d) calculated relative sensitivity.
Materials 15 06543 g008
Figure 9. (a) Temperature-dependent decay curves in the temperature range from 13 to 295 K; (b) calculated lifetime values; (c) lifetime values as a relation with temperature and the fitting line; (d) calculated relative sensitivities.
Figure 9. (a) Temperature-dependent decay curves in the temperature range from 13 to 295 K; (b) calculated lifetime values; (c) lifetime values as a relation with temperature and the fitting line; (d) calculated relative sensitivities.
Materials 15 06543 g009aMaterials 15 06543 g009b
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, J.; Lu, J.; Wu, Y.; Song, M. Luminescence of Mn4+ in a Zero-Dimensional Organic–Inorganic Hybrid Phosphor [N(CH3)4]2ZrF6 for Dual-Mode Temperature Sensing. Materials 2022, 15, 6543. https://doi.org/10.3390/ma15196543

AMA Style

Wang J, Lu J, Wu Y, Song M. Luminescence of Mn4+ in a Zero-Dimensional Organic–Inorganic Hybrid Phosphor [N(CH3)4]2ZrF6 for Dual-Mode Temperature Sensing. Materials. 2022; 15(19):6543. https://doi.org/10.3390/ma15196543

Chicago/Turabian Style

Wang, Jing, Jitao Lu, Yahong Wu, and Mingjun Song. 2022. "Luminescence of Mn4+ in a Zero-Dimensional Organic–Inorganic Hybrid Phosphor [N(CH3)4]2ZrF6 for Dual-Mode Temperature Sensing" Materials 15, no. 19: 6543. https://doi.org/10.3390/ma15196543

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop