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

High Verdet constant of Tb2O3-doped TiO-B2O3-Al2O3-Na2O magneto-optical glass

Open Access Open Access

Abstract

Tb-doped magneto-optical (MO) glass is widely used in fiber optics, optical isolators, and modulators. However, only the paramagnetic Tb3+ ions exhibit significant MO effects, whereas the diamagnetism Tb4+ ions suppress the MO effects. Therefore, the valence state control of Tb ions is very critical to optimize MO performance. Here, a reduction strategy was introduced to adjust the Tb valence in glass to achieve the high MO effect. The TiO, which has low valence Ti2+ ions and good reducibility, was used to suppress the oxidation of Tb3+ to Tb4+ ions. In the TiO-B2O3-Al2O3-Na2O glass, 10 mol% TiO can increase the Verdet constant at 650 nm by 19%. With the further increase in Tb2O3 concentration, the Verdet constant reaches a high value of 117 rad/(T·m) at 650 nm, which is close to the Verdet constant of TGG crystal (121 rad/(T·m)). This work provides a new approach to increase the Verdet constant of MO glass.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Magneto-optical (MO) glass refers to a class of magneto-optical materials with paramagnetism or diamagnetism. MO glass is an isotropy material without magnetic saturation. Because of its excellent MO performance, good transmittance in the visible and near-infrared, and ability to be made into a variety of complex optical components and large-size products and pulled into optical fibers, it has been used in MO devices such as MO modulators and MO isolators for aviation, military, guidance, satellite control, etc. [14] These materials are also known as Faraday materials, with the Faraday effect being defined by the rotation angle (θ=VBl) of a linearly polarized light beam under the influence of an external magnetic field (B) along a known optical path (l) [57]. The MO property of the material is typically described by the Verdet constant (V), which is generally proportional to the concentration of paramagnetic ions (such as some transition metals and rare earth (RE) ions) [5,8]. Therefore, the Verdet constant can be increased by introducing ions with high paramagnetism and high magnetic moment.

Among RE ions, Tb3+ ions exhibit strong magnetic susceptibilities (J = 6, g = 1.46) and moments (µeff = 9.5–9.72) [7,912], making it widely studied in the MO glass. However, Tb ions exist in two valence states(+3, + 4), and the presence of Tb4+ ions is detrimental to MO performance [13]. The magnetic dipole transition and electric dipole transition of the 4f-5d electron of Tb3+ ions are the main sources of their MO effect. Due to the unfilled 4f electron layer of RE elements, unpaired 4f electrons generate uncompensated magnetic moments, leading to strong magnetism. Currently, Tb3+ ions in MO glass are mainly introduced through Tb4O7, so not only does the raw material but also the oxidation reaction of Tb3+ to Tb4+ during the melting process can bring Tb4+ ions into the glass. Therefore, how to increase the ratio of the Tb3+/Tb4+ in the glass is an urgent issue that needs to be addressed. Yin et al. discovered that by substituting B2O3 for SiO2, the ratio of Tb3+/Tb4+ increases, resulting in improved MO performance [8]. Imaizumi et al. found that the Mn2+ (the new magnetic couplings) can improve the MO performance [14].

Considering the oxidizing properties of Ti2+, an appropriate amount of TiO can suppress the oxidation of Tb3+ to Tb4+. Therefore, TiO as a glass matrix can adjust the MO performance. In this work, the TiO, which has good reducibility, was introduced to suppress the oxidation of the Tb3+ ions. A series of Tb2O3-doped TiO-B2O3-Al2O3-Na2O glasses were prepared using the melting-quenching method. By introducing TiO, the mechanism of network structure and valence state changes in the glass were studied. Finally, a high concentration of 40 mol% Tb2O3 was successfully introduced in the glass, resulting in a Verdet constant as high as 117 rad/(T·m) at 650 nm.

2. Experimental

The experiment involved preparing a series of MO glasses using the melt-quenching method, and the compositions of the glasses are shown in Table 1. The raw materials comprised Tb4O7 (99.999%, Aladdin), H3BO3 (99.999%, Aladdin), Al2O3 (99.999%, Aladdin), TiO (99.999%, Aladdin), and Na2CO3 (99.999%, Aladdin). According to the pre-designed composition, the ingredients were thoroughly mixed for approximately 30 min and then transferred to alumina crucibles for melting at 1450 °C for 90 min in a high temperature lifting furnace. The glass was rapidly poured onto a preheated copper plate to prevent crystallization. Subsequently, the glass disks were transferred to an annealing furnace at 550 °C and naturally cooled to room temperature to eliminate the internal stress. The prepared MO glasses were then polished and characterized (Supplement 1, Fig. S1).

Tables Icon

Table 1. Compositions of the MO glasses

The glass structure is characterized by Fourier transform infrared spectroscopy (FTIR, INVENIO S) and Raman spectroscopy. For infrared spectroscopy testing, the glass samples are ground into powder and then mixed with dried spectroscopic-grade KBr powder in a ratio of 1:100 and pressed into disks at 10 MPa using a hydraulic press. Raman spectra were recorded on a Raman spectrometer (LABHRev-UV). The excitation wavelength is 532 nm. For the Differential Scanning Calorimetry (STA449F1), 30 mg powder samples were used to test from 25 °C to 1000°C. The density of the glass is measured using the Archimedes’ principle method. XRD (D8 Discover, Bruker, Karlsruhe, Germany) measurements of the powder sample within the 2θ range of 5°-70°. The samples were measured by X-ray photoelectron spectroscopy (XPS, ESCALAB 250Xi) with Mg-Kα as the X-ray source. The binding energies derived from the XPS measurements were calibrated concerning the C1s peak. The prepared glass samples were subjected to optical transmittance measurements using an ultraviolet-visible-near infrared (UV-Vis-NIR) spectrophotometer (UH4150, Hitachi, Japan). The photoluminescence (PL) spectra were recorded using an assembled optical system (QM/TM/NIR, PTI). Faraday rotation angles was recorded by commercial measurement setup (FD-FZ-C, Fudan Tianxin, Shanghai) under different magnetic field intensities. The light source is a 650 nm semiconductor laser.

3. Results and discussion

3.1 Structural analysis

The FTIR spectra of borate glass in Fig. 1(a) and 1(b) show the main absorption peaks between 400 and 1600 cm-1. The spectra can be divided into three regions: (1) 1200–1600 cm-1 corresponds to the asymmetric stretching vibration of the B-O bond in triangular [BO3] units. (2) 950–1200 cm-1 corresponds to the asymmetric stretching vibration of the B-O bond in tetrahedral [BO4] units. (3) 500–950 cm-1 corresponds to the vibration of Al-O and Ti-O bonds in [AlO4] and [TiO4] units, respectively [1520]. The detailed distribution of corresponding peaks can be found in Supplement 1, Table S1. As shown in XRD patterns (Fig. S2), no crystal precipitated in all the prepared glasses, confirming the amorphous nature of these glasses.

 figure: Fig. 1.

Fig. 1. FTIR spectra of MO glasses with different concentrations of (a) TiO and (b) Tb2O3. (c) peak fitting of Raman spectra in 600-1100 cm-1. Quantitative analysis of [TiO4]/[TiO6] and [BO3]/[BO4] area fraction ratio with different concentrations of (d) TiO and (e) Tb2O3.

Download Full Size | PDF

As shown in Fig. 1(a), the absorption peak of the [BO4] units between 950 and 1200 cm-1 weakened. This is due to the reduction of free oxygen atoms in the network structure as B2O3 is replaced by TiO, leading to the transformation from [BO4] units to [BO3] units. Additionally, the peak at 680 cm-1 for [TiO6] is enhanced, indicating the transformation from [TiO4] units to [TiO6] units with the increase of TiO[18,21]. Moreover, the absorption peak of the [BO4] unit in this range weakens with the increase of Tb2O3 as shown in Fig. 1(b). Since Tb ions do not participate in forming the glass network structure, it will deteriorate the structure with the increase of Tb2O3. The peak at 680 cm-1 for [TiO6] also increased due to the transformation from [TiO4] units to [TiO6] units.

Raman spectra can recognize the Ti-O vibration more clearly. As shown in Supplement 1, Fig. S3, the peaks in samples are mainly located in three bands: 350 cm-1, 600-850 cm-1, and 800-1000 cm-1. The corresponding peak assignments can be found in Table S2. For the 20T0Ti sample, there is no peak between 600 and 850 cm-1 due to the lack of TiO doping in the glass [22]. As shown in Fig. 1(c), peak fitting was conducted on the Raman spectra between 600 and 1100 cm-1 to facilitate a more precise quantitative analysis. As shown in Fig. 1(d) and 1(e), the area ratio of [TiO4] to [TiO6] continuously decreases with the increase of TiO. The peak of [TiO4] at 820 cm-1 can be considered as the former of the glass network, while the peak of [TiO6] at 718 cm-1 can be considered as the modifier of the glass network. This implies a transformation from [TiO4] to [TiO6] structural units, resulting in a decrease in the glass-forming ability (Fig. S4). The increase of the area ratio of [BO3] to [BO4] indicates that [BO4] is transforming into [BO3] structural unit. With the increase of Tb2O3, the area ratio of [TiO4] to [TiO6] continuously decreases, indicating a transformation from [TiO4] to [TiO6]. Furthermore, the area ratio of [BO3] to [BO4] increases, indicating a transformation from [BO4] to [BO3]. It is confirmed that the increase of Tb2O3 makes the structure of the glass network worse.

3.2 Valance states of Ti and Tb in MO glass

The glass contains a mixture of variable valence elements Ti and Tb. Since the MO property mainly arises from the Tb3+ ions, it is essential to understand the mechanism of Tb valence variation. Through XPS spectra of the Tb 3d and Ti 2p orbitals, the valence states of Tb and Ti can be analyzed.

Figure 2(a) shows the XPS fitting peak of the Tb element, in which the peaks can be divided into two parts. The binding energies at 1239, 1263, and 1274 eV correspond to Tb3+, while the binding energies at 1241, 1243, 1251, 1276, and 1278 eV correspond to Tb4+ ions [3,23]. As shown in Fig. 2(b), the binding energies at 463, 458, and 457 eV correspond to the +4, + 3, and +2 valence states of Ti ions, respectively [2426]. For the 20T0Ti glass, no peak in the 450-470 cm-1 can be seen. The XPS peak positions and their corresponding Ti and Tb ions valences are summarized in Supplement 1, Tables S3 and S4.

 figure: Fig. 2.

Fig. 2. XPS fitting peak of the (a) Tb element and (b) Ti element. The Calculated valence state of Ti and Tb ions, as well as Ti and Tb ions area fraction with different concentration of (c) TiO and (d) Tb2O3.

Download Full Size | PDF

As shown in Fig. 2(c), when the Tb2O3 concentration is fixed, the calculated valence state distributions of Tb ions decrease first and then increase with the increase in TiO concentration, and the calculated valence state distributions of Ti are the opposite. This is because more electrons enter the glass network structure with the increase of TiO, leading to simultaneous processes of Ti2+→Ti3+→Ti4+ and Tb4+→Tb3+. However, there is a critical point when the TiO concentration reaches 10 mol% (20T10Ti). The two processes mentioned above are inhibited when the TiO concentration increases to 15 mol%. Therefore, it is concluded that under the same Tb2O3 concentration, the ratio of Tb3+/Tb4+ increases first and then decreases with the increase of TiO. As shown in Fig. 2(d), when the TiO concentration is fixed, the valence state distribution of Ti and Tb ions increases with the increase in Tb2O3 concentration. The optical basicity of Tb2O3 and B2O3 are 0.954 and 0.425, respectively. With the increase of Tb2O3 concentration, the optical basicity of the glass increases. The high optical basicity of Tb2O3 makes it easy for Tb3+ to lose electrons and oxidize to Tb4+. As a result, the proportion of Ti2+ ions and Tb3+ ions decrease.

3.3 Optical properties

The transmittance and absorption spectra of the polished MO glasses are shown in Fig. 3(a) and 3(b). The overall average transmittance of the glass exceeds 70%, and the maximum is more than 86.6%. A prominent absorption peak is observed at 483 nm, mainly attributed to the Tb3+: 7F65D4 transition. Adding TiO and Tb2O3 to the glass reduces the glass transmittance, thereby lowering its laser damage threshold. As shown in Supplement 1, Fig. S5, the band gap (Eg) of the glass decreases with the increase of TiO or Tb2O3 concentration, resulting in the gradual shift of the UV absorption edge towards longer wavelengths.

 figure: Fig. 3.

Fig. 3. (a) Transmission and (b) absorption spectra of MO glasses. (c) Excitation and (d) emission spectra of MO glasses. (e) Energy diagram with the electronic transition process of Tb3+ ions.

Download Full Size | PDF

As shown in Fig. 3(c), the excitation peaks of MO glass doped with different concentrations of Ti and Tb are mainly located at 305, 317, 339, 351, 369, and 377 nm between 300 and 400 nm, corresponding to the transitions of Tb3+ ions from ground state 7F6 to excited states 7H6, 5H7, 5L6,7,8, 5L9 + 5G4, 5L10 + 5G5, and 5D3 + 5G6, respectively. Due to the shielding effect of the 5s25p6 shell on the 4f electrons of Tb3+ ions, the excitation band remains unchanged. The maximum intensity of the excitation peak is at 377 nm [2730].

As shown in Fig. 3(d), the emission spectra of the glasses show four emission peaks at 489, 544, 585, and 621 nm under 377 nm excitation, originating from the Tb3+: 5D47FJ (J = 6, 5, 4, 3) transitions. Among these emissions, the highest emission is at 544 nm. With the increase of TiO, the decreased emission intensity indicates the increase in the concentration of Tb3+ ions. Concentration quenching is typically the result of nonradiative energy transfer between Tb3+ ions. As the concentration of Tb2O3 increases, the emission intensity further quenches to almost no luminescence. To provide a clear description of the electronic transition process in Tb3+ ions, a schematic of the electronic transition process for Tb3+ ions is presented in Fig. 3(e).

3.4 MO property of glass

The MO property is an essential parameter for measuring the performance of MO glass. The Verdet constant is commonly used to represent the MO property, expressed as θ=VBl. This work verifies the relationship between the valence state and MO property by comparing the effective concentration of Tb ions (NTb), the concentration of Tb3+ in MO glasses (CTb3+), and the Verdet constant of MO glasses at 650 nm.

3.4.1 Tb3+ concentration in MO glass

In Tb-doped paramagnetic glasses, the effective concentration of Tb ion (NTb) can effectively predict the MO properties [10]. The NTb refers to successfully incorporating Tb into the glass structure, as simulated with Formula S3 in Supplement 1. As shown in Fig. 4(a), NTb increases continuously with the increase of TiO. As shown in Fig. 5(a), the variation of the Verdet constant first increases and then decreases because the TiO leads to an initial increase followed by a decrease in the Tb3+/Tb4+ ratio. Therefore, the NTb cannot effectively predict the MO properties in this system. We introduced the concept of the concentration of Tb3+ (CTb3+) in MO glasses and simulated it with Formula S4 in Supplement 1 [5], Comparing Formula S3 and S4, it can be observed that using NTb alone does not accurately reflect the valence state distribution of Tb ions. NTb calculates the total concentration of Tb ions entering the glass network. The formula for CTb3+ includes the proportion of the Tb3+ in the total Tb ions (PTb3+), enabling the determination of the quantity of Tb3+ ions entering the glass network. When considering the existence of multiple valence states, it is evident that PTb3+ significantly increases with the increase in TiO concentration in the MO glass. As shown in Fig. 4(c) and 4(d), the variation trendency of CTb3+ is consistent with that of the Verdet constant. Hence, CTb3+ can effectively predict the MO property.

 figure: Fig. 4.

Fig. 4. The effective concentration of Tb ion (NTb) with different concentrations of (a) TiO and (b) Tb2O3. The concentration of Tb3+ (CTb3+) with different concentrations of (c) TiO and (d) Tb2O3

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Verdet constant of MO glasses with different concentrations of (a) TiO and (b) Tb2O3 at 650 nm, (c) Verdet constant of magneto-optical glass from recent literatures.

Download Full Size | PDF

3.4.2 Verdet constant of MO glass

As shown in Fig. 5(a), the addition of an appropriate amount of TiO is beneficial for the Verdet constant. The maximum Verdet constant at 650 nm is 59.4 rad/(T·m) for the 20T10Ti because the Tb3+/Tb4+ ratio is the highest at 10 mol% TiO, and CTb3+ is also at its maximum. As shown in Fig. 5(b), the Verdet constant increases continuously with the increase of Tb2O3 when the TiO concentration is fixed. The MO property of paramagnetic glass originates from Tb3+ ions. With the increase of Tb2O3, more Tb3+ ions enter the glass, which increases the MO property.

Currently, terbium gallium garnet (TGG) crystal is the main commercial MO material with a high Verdet constant of 121 rad/(T·m) at 650 nm [31]. In this work, the maximum Verdet constant of 117 rad/(T·m) at 650 nm is achieved when the Tb2O3 is 40 mol% (V40T10Ti). The MO crystals are limited in terms of long-distance light transmission and rotation. In contrast, magneto-optical glass can be quickly processed into any shape and size, and glass manufacturing is a mature technology, highlighting the potential for magneto-optical glass to replace TGG crystals. As shown in Fig. 5(c), this work significantly improves the Verdet constant compared to previously reported works [5,10,15].

4. Conclusion

Tb-doped TiO-B2O3-Al2O3-Na2O glasses were prepared using the melt-quenching method, and their structures were analyzed using Raman and infrared spectroscopy. The addition of an appropriate amount of TiO facilitates the simultaneous processes of Ti2+→Ti3+→Ti4+ and Tb4+→Tb3+, which is favorable for enhancing the Verdet constant. The variation trend of the effective concentration of Tb ions in this system cannot effectively predict the change in the Verdet constant, whereas the introduction of CTb3+ shows high consistency with the Verdet constant. When the glass doped with 20 mol% Tb2O3 and 10 mol% TiO (V20T20Ti), the Verdet constant reaches its maximum value of 59.4 rad/(T·m), which is 19% higher than that of V20T0Ti. With the increase of Tb2O3 concentration, more Tb3+ ions enter the glass network, leading to an increase in the Verdet constant. Consequently, a recorded Verdet constant of 117 rad/(T·m) at 650 nm was achieved with 40 mol% Tb2O3 doped MO glass. These results demonstrate that introducing TiO as a glass matrix can enhance the CTb3+ and obtain a higher Verdet constant.

Funding

National Natural Science Foundation of China (52372014, 62275206, U2231236).

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.

Supplemental document

See Supplement 1 for supporting content.

References

1. M. Djamal, L. Yuliantini, R. Hidayat, et al., “Glass medium doped rare earth for sensor material,” Mater. Today: Proc. 5(7), 15126–15130 (2018). [CrossRef]  

2. L. Feng, B. Lai, J. Wang, et al., “Spectroscopic properties of Er3+ in a oxyfluoride glass and upconversion and temperature sensor behaviour of Er3+/Yb3+-codoped oxyfluoride glass,” J. Lumin. 130(12), 2418–2423 (2010). [CrossRef]  

3. Y. Zhang, S. Murai, A. Maeno, et al., “Microstructure and Faraday effect of Tb2O3-Al2O3-SiO2-B2O3 glasses for fiber-based magneto-optical applications,” J. Am. Ceram. Soc. 105(2), 1198–1209 (2022). [CrossRef]  

4. H. Yin, Y. Gao, Y. Gong, et al., “Wavelength dependence of Tb3+ doped magneto-optical glass Verdet constant,” Ceram. Int. 44(9), 10929–10933 (2018). [CrossRef]  

5. X. Zhao, W. Li, Q. Xia, et al., “High Verdet constant glass for magnetic field sensors,” ACS Appl. Mater. Interfaces 14(51), 57028–57036 (2022). [CrossRef]  

6. H. Yang and Z. Zhu, “Magneto-optical glass mixed with Tb3+ ions: High Verdet constant and luminescence properties,” J. Lumin. 231, 117804 (2021). [CrossRef]  

7. K. J. Carothers, R. A. Norwood, and J. Pyun, “High Verdet constant materials for magneto-optical faraday rotation: a review,” Chem. Mater. 34(6), 2531–2544 (2022). [CrossRef]  

8. H. Yin, Y. Gao, H. Guo, et al., “Effect of B2O3 content and microstructure on Verdet constant of Tb2O3-doped GBSG magneto-optical glass,” J. Phys. Chem. C 122(29), 16894–16900 (2018). [CrossRef]  

9. L. Zhou and Z. Zhu, “Magneto-optical and luminescent properties of Tb3+ ions doped GBPZ magneto-optical glass,” J. Non-Cryst. Solids 574, 121165 (2021). [CrossRef]  

10. J. Ding, P. Man, Q. Chen, et al., “Influence of Tb3+ concentration on the optical properties and Verdet constant of magneto-optic ABS-PZZ glass,” Opt. Mater. 69, 202–206 (2017). [CrossRef]  

11. D. F. Franco, Y. Ledemi, W. Correr, et al., “Magneto-optical borogermanate glasses and fibers containing Tb3+,” Sci. Rep. 11(1), 9906–9921 (2021). [CrossRef]  

12. D. F. Franco, R. G. Fernandes, J. F. Felix, et al., “Fundamental studies of magneto-optical borogermanate glasses and derived optical fibers containing Tb3+,” J. Mater. Res. Technol. 11, 312–327 (2021). [CrossRef]  

13. M. Valeanu, M. Sofronie, A. C. Galca, et al., “The relationship between magnetism and magneto-optical effects in rare earth doped aluminophosphate glasses,” J. Phys. D: Appl. Phys. 49(7), 075001 (2016). [CrossRef]  

14. D. Imaizumi, T. Hayakawa, and M. Nogami, “Faraday rotation effects of Mn2+-modified Tb2O3-B2O3 glass in pulsed magnetic field,” J. Lightwave Technol. 20(4), 740–744 (2002). [CrossRef]  

15. H. Lin, B. Liu, L. Zhou, et al., “Enhancements of magneto-optical properties of GeO2-PbO-B2O3-SiO2-P2O5 glass doped Tb3+ ion,” Mater. Chem. Phys. 282, 125963 (2022). [CrossRef]  

16. W. A. Pisarski, J. Pisarska, and W. Ryba-Romanowski, “Structural role of rare earth ions in lead borate glasses evidenced by infrared spectroscopy: BO3$\leftrightarrows$BO4 conversion,” J. Mol. Struct. 744-747, 515–520 (2005). [CrossRef]  

17. S. Chen, S. Zhu, C. Liao, et al., “Influence of Boron anomaly on the NIR fluorescence concentration quenching of erbium ions in borate glasses,” J. Non-Cryst. Solids 576, 121274 (2022). [CrossRef]  

18. H. Doweidar, K. El-Egili, R. Ramadan, et al., “Structural units distribution, phase separation and properties of PbO-TiO2-B2O3 glasses,” J. Non-Cryst. Solids 466-467, 37–44 (2017). [CrossRef]  

19. H. An, P. Qian, Y. Ye, et al., “Structural and magneto-optical properties of PbO-Bi2O3-B2O3-Al(PO3)3 glasses,” J. Non-Cryst. Solids 538, 120021 (2020). [CrossRef]  

20. M. C. Dinca, B. A. Sava, A. C. Galca, et al., “Magneto-optical properties of borophosphate glasses co-doped with Tb3+ and Dy3+ ions,” J. Non-Cryst. Solids 568, 120967 (2021). [CrossRef]  

21. B. Zhang, F. He, X. Cao, et al., “The effect of TiO2 and B2O3 on sintering behavior and crystallization behavior of SrO-BaO-B2O3-SiO2 glass-ceramics,” Ceram. Int. 48(5), 7013–7023 (2022). [CrossRef]  

22. Z. Wang and R. Xu, “Effects of TiO2 on the structural characteristics of CaO-SiO2-Al2O3-TiO2 glass in the same superheat state studied by Raman spectra,” Ceram. Int. 49(16), 26494–26504 (2023). [CrossRef]  

23. I. M. Nagpure, S. S. Pitale, E. Coetsee, et al., “Low voltage electron induced cathodoluminescence degradation and surface characterization of Sr3(PO4)2:Tb phosphor,” Appl. Surf. Sci. 257(23), 10147–10155 (2011). [CrossRef]  

24. Q. Chen, M. Zhang, H. Wang, et al., “Structures and magneto optical property of diamagnetic TiO2-TeO2-PbO-B2O3 glass,” J. Non-Cryst. Solids 468, 58–66 (2017). [CrossRef]  

25. S. H. Nam, S. J. Cho, C. K. Jung, et al., “Comparison of hydrophilic properties of TiO2 thin films prepared by sol-gel method and reactive magnetron sputtering system,” Thin Solid Films 519(20), 6944–6950 (2011). [CrossRef]  

26. H. Kang, C. Lee, S. C. Yoon, et al., “Layer-by-layer assembled multilayer TiOx for efficient electron acceptor in polymer hybrid solar cells,” Langmuir 26(22), 17589–17595 (2010). [CrossRef]  

27. C. Zuo, J. Huang, S. Liu, et al., “Luminescence and energy transfer of Tb3+-doped BaO-Gd2O3-Al2O3-B2O3-SiO2 glasses,” Spectrochim. Acta, Part A 187, 181–185 (2017). [CrossRef]  

28. P. Dharmaiah, C. S. D. Viswanath, C. Basavapoornima, et al., “Luminescence and energy transfer in Dy3+/Tb3+ co-doped transparent oxyfluorosilicate glass-ceramics for green emitting applications,” Mater. Res. Bull. 83, 507–514 (2016). [CrossRef]  

29. Z. X. Mo, H. W. Guo, P. Liu, et al., “Luminescence properties of magneto-optical glasses containing Tb3+ ions,” J. Alloys Compd. 658, 967–972 (2016). [CrossRef]  

30. Q. Wang, S. Song, N. Nie, et al., “Luminescent NaTb(SO4)2 nanoprobe for hydrogen peroxide based on switchable fluorescence of Tb4+/Tb3+ redox couple,” Ceram. Int. 47(13), 18942–18947 (2021). [CrossRef]  

31. B. Geist, R. Ronningen, A. Stolz, et al., “Radiation stability of visible and near-infrared optical and magneto-optical properties of terbium gallium garnet crystals,” Appl. Opt. 54(10), 2866–2869 (2015). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental Document

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 (5)

Fig. 1.
Fig. 1. FTIR spectra of MO glasses with different concentrations of (a) TiO and (b) Tb2O3. (c) peak fitting of Raman spectra in 600-1100 cm-1. Quantitative analysis of [TiO4]/[TiO6] and [BO3]/[BO4] area fraction ratio with different concentrations of (d) TiO and (e) Tb2O3.
Fig. 2.
Fig. 2. XPS fitting peak of the (a) Tb element and (b) Ti element. The Calculated valence state of Ti and Tb ions, as well as Ti and Tb ions area fraction with different concentration of (c) TiO and (d) Tb2O3.
Fig. 3.
Fig. 3. (a) Transmission and (b) absorption spectra of MO glasses. (c) Excitation and (d) emission spectra of MO glasses. (e) Energy diagram with the electronic transition process of Tb3+ ions.
Fig. 4.
Fig. 4. The effective concentration of Tb ion (NTb) with different concentrations of (a) TiO and (b) Tb2O3. The concentration of Tb3+ (CTb3+) with different concentrations of (c) TiO and (d) Tb2O3
Fig. 5.
Fig. 5. Verdet constant of MO glasses with different concentrations of (a) TiO and (b) Tb2O3 at 650 nm, (c) Verdet constant of magneto-optical glass from recent literatures.

Tables (1)

Tables Icon

Table 1. Compositions of the MO glasses

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.