Next Article in Journal
Experimental Investigation of Water Vapor Concentration on Fracture Properties of Asphalt Concrete
Next Article in Special Issue
Enhancing Ablation Resistance of TaB2-Based Ultra-High Temperature Ceramics by Mixing Fine TaC Particles and Dispersed Multi-Walled Carbon Nanotubes
Previous Article in Journal
Prioritizing Sustainable Denim Fabric through Integrated Decision-Making Framework
Previous Article in Special Issue
UVC Up-Conversion and Vis-NIR Luminescence Examined in SrO-CaO-MgO-SiO2 Glasses Doped with Pr3+
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structure and Spectral Properties of Er3+-Doped Bismuth Telluride Near-Infrared Laser Glasses

1
College of Science, Changchun University, Changchun 130022, China
2
School of Materials Science and Engineering, Changchun University of Science and Technology, Changchun 130022, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(13), 3292; https://doi.org/10.3390/ma17133292
Submission received: 13 May 2024 / Revised: 14 June 2024 / Accepted: 14 June 2024 / Published: 3 July 2024

Abstract

:
TeO2-Bi2O3-B2O3-ZnO laser glasses doped with Er3+ were synthesized through an optimized melt-quenching method. The absorption spectra at 808 nm LD pumping were studied. Various spectral tests and data analyses indicate that the maximum fluorescence emission intensity can be obtained when the Er3+ doping concentration reaches 2%. In this case, the emission cross-section can reach up to 9.12 × 10−21 cm2 and the gain coefficient at 1.55 μm is 6.17 cm−1. Simultaneously, the sample has a lower phonon energy in the high-frequency band at 1077 cm−1, which reduces the probability of non-radiative relaxation. The calculated energy transfer coefficient CD-A is 13.8 × 10−40 cm6/s, reflecting the high cross-relaxation probability of Er3+ in the sample, which promotes the luminescence of 1.55 μm and favors the emission in the near-infrared region. The comprehensive results demonstrate that the prepared Er3+-doped bismuth telluride laser glass can be used as a promising and high-quality gain material for near-infrared lasers.

1. Introduction

It is well-known that Er3+ emits luminescence when effectively excited by laser diode pumping. This luminescence occurs with a jump at the 4I13/24I15/2 energy level. The center of the spectrum is located at 1550 nm. Rare-earth-ion-doped laser materials have achieved much attraction due to their extensive applications, including display devices, medical imaging, fiber-optic sensing, photonic crystal fibers, fiber-optic amplifiers (EDFAs), optical temperature sensors (OTSs), solid-state lasers, etc. [1,2,3,4,5,6]. Although it has long been of wide interest to researchers, it remains a hot topic to this day. As research has progressed, the types of glass matrices selected have become more varied. The main purpose of these matrices has changed: more expansion is obtained in improving the intensity as well as the efficiency of the Er3+ radiative excursion. Various glass matrices influence the electron layer structure of rare earth ions. Thus, discovering excellent matrix glass is also an essential aspect in achieving near-infrared intensity enhancement. Multi-component glass fibers present several benefits when compared to SiO2 glass. These benefits include lower phonon energy, good thermal stability, larger full-width at half-maximum (FWHM) values, and greater absorption/emission cross-sections. Examples of such glasses include fluoride glass, tellurite glass, and germanate glass. Furthermore, matrix glasses that contain heavy oxides, including TeO2, Bi2O3, and others, show promise as material species for near-infrared, fiber-optic sensing, and non-linear optic fields [7,8,9]. Tellurite glass is a suitable matrix glass material for near-infrared laser materials due to its many network structural units, adjustable glass compositional interval, low phonon energies (700 cm−1), good thermal stability, high refractive indices (~1.8–2.2), and ability to accommodate high rare earth doping concentrations and higher laser damage thresholds [10].
Moreover, it is crucial to include glass-forming agents, and boron oxide (B2O3) is commonly preferred. B2O3 glasses offer significant advantages over other types of glasses, including high mechanical properties, chemical and thermal stability, and a low melting point [11]. However, the network structure generates high phonon energies, which increase the non-radiative emission of tellurite glass. An appropriate amount of heavy metal oxides can be added to reduce the phonon energy and increase the energy transfer efficiency [1]. Recent research findings have indicated that adding ZnO to tellurite glass can increase its density, while reducing both the softening point and glass transition temperature of the binary glass [12,13]. At present, the majority of studies concentrate on reporting the luminescence properties of heavy metal oxide glass matrices that contain rare earth ion-doped compounds such as ZnO, Bi2O3, and so on. However, the weak luminescence of Te-activated glass in the bands used for near-infrared communication (e.g., O-L band) means that there has been less exploration into Te-activated glass components, and the combination of Er3+-doped tellurite glass with ZnO, B2O3, and Bi2O3 oxides remains a topic that has not yet received extensive attention.
In this work, Er3+-doped bismuth tellurite glasses with the composition TeO2-Bi2O3-B2O3-ZnO was prepared by conventional melt-quenching method. The thermal stability of tellurite glass and the luminescence properties of rare earth ions were investigated experimentally and theoretically. The Judd–Ofelt intensity parameters, spontaneous leapfrog chances, fluorescence branching ratios, and radiative lifetimes were calculated. The maximum absorption cross-section recorded was 3.0 × 10−21 cm2, while the emission cross-section was 9.12 × 10−21 cm2. Subsequently, gain coefficients were calculated. Finally, the energy transfer mechanism of Er3+ was analyzed.

2. Materials and Methods

Er3+-doped tellurite glasses with the molar composition of 80TeO2-5Bi2O3-5B2O3-10ZnO-xEr2O3 (where x = 1, 1.5, 2, 2.5, 3 mol%, which were labeled as TBBZ1 to TBBZ5) were prepared by the melt quenching method with raw materials including TeO2, B2O3, ZnO, and Er2O3, all of which have a purity of 99.99%. A precisely measured 10 g of the raw material was mixed well, ground thoroughly in an onyx mortar, melted in an alumina oxide crucible at 1000 °C for about 30 min under the protection of an O2/N2 atmosphere. The glass samples were annealed at 370 °C for 3 h in order to eliminate thermal stress, and then slowly reduced to room temperature at a cooling rate of 10 °C/h. Finally, the glass samples were cut and polished to a thickness of 2 mm for subsequent optical and spectral measurements.
Distilled water was used as the impregnation solution and the density of the TBBZ3 glass was measured to be 4.7841 g/cm3 using the Archimedes’ method. Refractive index at 1550 nm measured with SPA 4000TM prism coupler at room temperature was 1.81. The glass transition temperature (Tg) and crystallization temperature (Tx) were measured by differential scanning calorimeter (DSC) at a heating rate of 10 °C/min under an N2 atmosphere. In addition, the absorption spectra were tested by a PerkinElmer Lambda 900 UV/VIS/NIR spectrophotometer and the emission spectra were measured with a Triax 320 type spectrometer (Jobin Yvon Corp., Stow, MA, USA) under 808 nm laser diode excitation. Raman and infrared spectra were measured using a Renishaw inVia Raman microscope and a Spectrum Two type spectrometer (PerkinElmer, Waltham, MA, USA).

3. Results

3.1. Thermal Stability

Thermal stability has an impact on fiber formation performance, the glass transition temperature (Tg), the crystallization temperature (Tx), and their difference ΔT = Tx − Tg is an important evaluation index in the thermal stability analysis of glasses. Stretching of fiber-optic cables is preferred when the temperature difference (ΔT) exceeds 100 °C [14]. Figure 1 shows the DSC curves of TBBZ glass with values of Tg, Tx, and ΔT of 396 °C, 524 °C, and 128 °C, respectively. Its ΔT is larger than that of other Er3+-doped tellurite glass samples (TeO2-ZnO-Na2O, ΔT = 127.2 °C) [15], germanate glass (Bi2O3-GeO2-Ga2O3-Na2O, ΔT = 121 °C) [16], and germanium tellurite glass (GeO2-TeO2-Na2O, ΔT = 106 °C) [17]. The further improvement in thermal stability may be attributed to the new evolution of the original glass network structure due to the introduction of B2O3. Due to the reduced atomic radius of boron, the repulsion between non-bonding electrons is somewhat diminished, facilitating a tighter bonding between atoms. This, together with TeO2, modulates the population of complex oxygen ions, thereby enhancing the thermal stability of the poly glasses presented in this paper [18]. As a result, the tellurite glass produced in this study shows favorable thermal stability, which is advantageous for the formation of fibers.

3.2. XRD and Transmission Spectroscopy

To investigate the structure of tellurite glass, the sample material was characterized and analyzed. The XRD spectra of the TBBZ glass samples are given in Figure 2. As can be seen from the figure, there are two broad bands at 2θ equal to 30°, as well as near 50° without sharp crystal peaks, indicating that the prepared glass is completely amorphous in nature and does not give rise to the phenomenon of precipitation of crystals, presenting an amorphous state.
In addition, the Er3+-doped TeO2-Bi2O3-B2O3-ZnO glass has a high transmittance, as shown in Figure 3. The transmittance of the glass in the near-infrared band can reach 90% with high transmittance, which is an important reference value for improving the optical properties of tellurite glass. The overall transmittance decreases slightly with increasing Er3+-doping concentration, which is due to absorption by the Er3+ ions. The spectral positions and shapes of the glasses at different rare earth ion doping concentrations are approximately the same, approximately due to the fact that the Er3+ ions do not aggregate in the local ligand field but are uniformly distributed in the tellurite glass network, which is similar to previous reports [19].

3.3. Raman Spectrum and FT-IR Spectra

In order to investigate the physical structure of TBBZ glass, an analytical discussion of the glass network was performed. Typically, Te4+ ions have a larger radius and generate a looser glass network. Whereas B3+ ions has a smaller radius, the incorporation of B3+ ions results in the formation of BO3 triangles and BO4 tetrahedra within the glass. These groups combine to form a variety of three-dimensional topological cages comprising borate or diborate rings, metaborate rings, and others (see Figure 4). Such a structure can assist in modulating the degree of polymerization of the tellurite glass topological cage structure, promoting the NIR luminescent centers to be properly spaced from each other and suppressing non-radiative relaxation; the network modifier (ZnO) provides free oxygen in the glass structure and induces the transformation of the borate structural unit from BO3 to BO4, which improves the polymerization of the glass network and further promotes the formation of the [TeO4] topological cage. Figure 4 demonstrates the involvement of boron oxide in the modulation of the topological cage structure of the Er3+-doped tellurite glass network in this paper. Erbium and bismuth, as activators, are located outside the topological cage and have an insignificant influence on the composition of the glass network.
Figure 5 shows the Raman spectra of the matrix glass measured using the Gaussian fitting method. There are seven distinct spectral bands near 146 cm−1, 264 cm−1, 371 cm−1, 462 cm−1, 625 cm−1, 780 cm−1, and 1077 cm−1. Among them, the Gaussian peak at 146 cm−1 is due to the Bi3+ vibration in [BiO6] [20], and the Gaussian peaks at 264 cm−1, 371 cm−1, and 462 cm−1 are due to the bending vibration or symmetric stretching of Te-O-Te in the [TeO4] (tbps), [TeO3+δ], and [TeO3] (tps) moieties [21] and the [BiO3], and Bi-O-Bi symmetric stretching [22] in [BiO6] is caused by the joint action. Unlike the peak around 650 cm−1 in normal tellurite glasses, the introduction of B2O3 shifts the peak there to 780 cm−1, probably because the addition of B2O3 breaks the Te-O-Te bond, resulting in the transformation of [TeO4] into [TeO3] and [TeO3+δ] units [22]. With the addition of B2O3, the formation of a new peak at 1077 cm−1 is associated with the B-O and B-O-B symmetric stretching vibrations in [BO3] and [BO4] [23]. Therefore, the maximum phonon energy of the matrix glass is 1077 cm−1 and the low phonon energy reduces the probability of non-radiative relaxation and favors Er3+ emission in the NIR region.
In addition, a Gaussian decomposition of the infrared spectral lines of the TBBZ glass was performed in order to determine the molecular vibrations in the TBBZ glass network. The infrared spectrogram at 330–1200 cm−1 is given in Figure 6. Four distinct spectral bands are present at approximately 341 cm−1, 572 cm−1, 1023 cm−1, and 1157 cm−1. The spectral band at 341 cm−1 corresponds to the symmetric stretching or bending vibrations of Te-O-Te in [TeO4] (tbps), [TeO3+1], and [TeO3] (tps) [24]. The 572 cm−1 peak could potentially originate from the bending vibration of B-O-B in [BO3] and the stretching vibration of Te-O in [TeO4] [25]. The bands at 1023 cm−1 and 1157 cm−1 are attributable to B-O antisymmetric stretching vibrations in [BO3] [26,27].

3.4. Absorption Spectrum and Judd–Ofelt Analysis

The absorption spectra of Er3+-doped tellurite glass samples are shown in Figure 7 from 400 to 2000 nm. The intensity of the absorption peaks of the glass samples was significantly enhanced when the Er3+ doping concentration was increased. Observation of the Er3+ absorption spectrum revealed the presence of seven absorption peaks on the spectral line, and the absorption bands located at 489, 529, 552, 655, 798, 975, and 1533 nm corresponded to the electronic leaps from the ground state (4I15/2) to the excited states 4F7/2, 2H11/2, 4S3/2, 4F9/2, 4I9/2, 4I11/2, and 4I13/2, respectively. The absorption band of Er3+ at less than 300 nm could not be measured due to the rapid increase in the electronic absorption fringes [28]. This indicates the uniform binding of Er3+ into the tellurite glass network. Er3+ exhibits a clear absorption band at approximately 808 nm (4I15/24I9/2), indicating the potential for commercial 808 nm laser diodes to excite singly doped Er3+ in tellurite glass fibers.
To examine the luminescent properties of tellurite glasses, the Judd–Ofelt theory [29,30] was used to analyze both the local structure and the luminescent properties of trivalent rare earth ions. Based on the absorption spectra, the J-O intensity parameter Ωλ(λ=2,4,6) of the TBBZ glass can be calculated by the least squares method. Table 1 displays the J-O parameters of various tellurite matrix glasses doped with different concentrations of Er3+. The parameters of J-O intensity (Ω2, Ω4, Ω6) exhibit a declining trend as the concentration of Er3+ doping increases. It should be noted that the calculated glass strength parameters Ω2 > Ω4 > Ω6 are similar to the phenomenon previously found in tellurite glasses [31]. In general, Ω2 correlates with the ligand symmetry of the immediate surroundings and is highly responsive to the covalent bonding between rare earth ions and anionic ligands. Larger values of Ω2 indicate lower ligand symmetry and a more compact glass network structure. The Ω2 values of TBBZ glass in Table 1 exceed those of other matrix glasses; this could be attributed to the disruption of the glass network through doping with oxides such as B2O3 and ZnO, leading to the improved stability of the TBBZ glass. While the value of Ω2 decreases with continuous Er3+ doping, the reason for this phenomenon may be the conversion of [TeO3](tp) to [TeO4](tbp) units and the subsequent decrease in ligand bond covalency of Er3+ [32]. Ω4 is related to the viscosity of the glass and decreases with the increasing Er3+ doping concentration. Ω6 is proportional to the body glass stiffness, and as Ω6 increases, its corresponding mechanical properties decrease. Meanwhile, a reduction in the covalency between rare earth ions and oxide ions appears to markedly decrease the value of Ω6 [32].
After obtaining the intensity parameter Ωλ(λ=2,4,6), it was used to calculate the glass TBBZ3’s spontaneous leap chance Arad, fluorescence branching ratio β, and radiation lifetime τ. As shown in Table 2, the spontaneous jump chance of Er3+ in TBBZ3 at 4I11/24I13/2 is about ~25 s−1, which is higher than that of fluorotelluride [36] (21 s−1), ZBLAN (18 s−1), and fluoroaluminate [37] (19 s−1). The higher spontaneous transition probability Arad is easier to achieve laser emission and can be effectively used in the field of 2.7 μm lasers [38]. The radiative lifetime at 4I13/24I15/2 is 3.52 ms, which is higher than the reported TeO2-ZnO-ZnF2 tellurite glass [39] (3.23 ms) and bismuthate glass [40] (2.94 ms). The above studies show that the prepared TBBZ3 glass has a strong potential for development in the application of lasers in the near-infrared band.

3.5. Fluorescence Spectrum and Lifetime

The topological cage structure demonstrated in Figure 4 can effectively regulate the ligand field around rare earth ions and thus the luminescence behavior. Figure 8a,b shows the fluorescence spectra of TBBZ1-5 in the range of 1400–1700 nm and 2500–3200 nm, respectively, both produced by 808 nm LD pump excitation. The emission peak at the central wavelength of 1550 nm in Figure 8a is due to Er3+ during the energy transfer from 4I13/2 to 4I15/2. The emission peak at the central wavelength of 2750 nm in Figure 8b is caused by the jumping of Er3+ at the 4I11/24I13/2 energy level. From the observation, it can be seen that the fluorescence intensity of TBBZ glass at both 1550 nm and 2750 nm shows a repeated decreasing and increasing trend as the Er3+ concentration is doped from 1 mol% to 3 mol%, and the fluorescence intensity reaches the peak at 2 mol% concentration. The phenomenon in which a reduction in intensity occurs is known as fluorescence concentration quenching, which may be due to aggregation between Er3+ ions caused by an increase in Er2O3 content, which leads to enhanced interactions and increased probability of radiationless leaps, resulting in a decrease in the luminescence efficiency, and a similar trend has been found in tellurite glasses [31]. The insets in Figure 8a,b show the full-width values of the fluorescence half-peaks for each Er3+ doping concentration at 1.55 μm and 2.7 μm, respectively. The effective fluorescence half-peak full-width (Δλeff) can be calculated from the following equation:
Δ λ eff = I λ   d λ I max
where I(λ) is the fluorescence intensity at the wavelength λ and Imax is the maximum intensity of the fluorescence emission. From the figure, it can be seen that increasing the Er3+ doping concentration in the range of the 1.55 μm and 2.7 μm bands shows a repeated decreasing and increasing trend in the FWHM in line with the change in fluorescence intensity. Both peak at a doping concentration of 2%, 87.56 nm compared to 184.8 nm. The larger half-peak full-width value is crucial for optimizing the bandwidth characteristics of the glass.
The fluorescence decay of Er3+ also provides important information on luminescence and energy transfer processes, Figure 9 shows the fluorescence decay curves at 1.55 μm for the Er3+:4I13/2 energy level in the TBBZ3 glass sample under pulsed excitation at 808 nm LD, with the lifetimes averaged. It is well-known that the energy level lifetime of rare earth ions refers to the length of time experienced when the fluorescence intensity decays to the highest intensity of e−1. Fitting by exponential decay function, the fluorescence lifetime (τmea) of TBBZ3 glass was measured to be 3.2 ms, which is much longer than that of bismuthate glass (2.05 ms) [9], zinc fluoride glass (2.05 ms) [41], and fluorotelluride glass (1.36 ms) [42]. Increased longevity is advantageous for the population accumulation towards achieving lasers.
The quantum efficiency η of the 4I13/2 energy level of Er3+ can be calculated by reference to Equation (2):
η = τ mea τ rad × 100 %
where τmea is the measured fluorescence lifetime and τrad is the calculated radiation lifetime. The quantum efficiency of the Er3+:4I13/2 energy level in TBBZ3 is 90%, which is higher than that of the Er3+-doped fluoride glass (21.29%) [19], the high quantum efficiency guarantees ample gain for the 1.55 μm laser, which is a promising contender for 1.55 μm laser realization.

3.6. Absorption and Emission Cross-Sections and Gain Coefficients

To characterize the spectral properties of TBBZ glass in the NIR region, the absorption and emission cross-sections at 1.55 μm were calculated and analyzed. The absorption cross-section σabs can be calculated by the Beer–Lambert law [43]:
σ abs λ = 2.303 lg I 0 / I N 0 L
where lg(I0/I) is the absorption intensity; I0 is the incident light intensity; I is the transmitted light intensity; N0 is the rare earth ion concentration; and L is the sample thickness.
The emission cross-section σemi can be calculated by the Füchtbauer–Ladenburg [14] equation:
σ emi = λ 4 A rad 8 n 2 c × λ I λ λ I λ d λ
where λ is the wavelength; Arad is the spontaneous jump chance; c is the speed of light; n is the refractive index; and I(λ) is the fluorescence intensity at wavelength λ. Figure 10 shows the absorption and emission cross-sections produced by Er3+ during energy transfer at two energy levels, 4I13/2 and 4I15/2, with peaks σabs = 4.49 × 10−21 cm2 and σemi = 9.12 × 10−21 cm2 at 1532 nm and 1556 nm, respectively. This is higher than the previously reported Er3+-doped tellurite glass [39] (8.7 × 10−21 cm2) and fluoroaluminate glass [44] (5.30 × 10−21 cm2). The high absorption–emission cross-section is conducive to the enhancement of the pump absorption efficiency and the effective emission of Er3+ at 1.55 μm, which further proves that the TBBZ3 glass is an excellent dielectric material for applications in the near- and mid-infrared.
The gain coefficient G(λ) [44] can be approximated based on the previously calculated absorption σabs and emission cross-section σemi:
G λ = N P σ emi λ 1 P σ abs λ
where N denotes the concentration of rare earth ions and P is the population inversion factor. Figure 11 gives the gain coefficients calculated for the TBBZ3 glass at 1.55 μm from the range 0–1 in steps of 0.2. Positive gain at 1550 nm is achieved when the inversion factor P reaches 0.4 or higher, and the low inversion factor indicates that the Er3+-doped tellurite glass has a low pumping threshold at 1.55 μm. The gain coefficient with inversion factor P = 1 peaks near 1556 nm with a maximum gain coefficient Gmax = 6.17 cm−1, which is superior to fluoride glass [44] (1.15 cm−1), as well as other tellurite glasses [45] (0.81 cm−1). Due to its high gain performance, TBBZ3 glass has the potential to be used as a gain medium material in the field of rare-earth-doped fiber lasers within the 1.55 μm band.

3.7. Energy Transfer Mechanism and Microscopic Parameter

Figure 12 shows the energy level and energy transfer process diagram of Er3+ in TBBZ glass during 808 nm LD pump excitation. The energy of Er3+ at 4I15/2 is excited to 4I9/2 by ground state absorption (GSA). A portion of the Er3+ energy is transferred to the neighboring 4I9/2 energy level by means of energy transfer (EM). Due to the narrow energy gap between 4I9/2 and 4I11/2, Er3+ emits at 2678 nm from the 4I9/2 level to 4I11/2 by means of non-radiative transitions to 4I13/2 and subsequently to 4I13/2, and emits at 1533 nm when it returns from the 4I13/2 level to 4I15/2. The fluorescence intensity of TBBZ glass at 1.53 μm rises with increasing Er3+ doping concentration. This increase can be attributed to the cross-chattering process (CR) between 4I9/24I13/2 and 4I15/24I13/2.
To further investigate the energy transfer process, Dexter’s theory was used to calculate the microscopic parameters of the energy transfer between Er3+ based on the absorption and emission cross-sections of rare earth ions. The microscopic energy transfer probability between rare earth ions is expressed by the following equation [14]:
W D - A = C D - A R 6
where D and A denote donor and acceptor ions, respectively, and R is the distance between the donor and acceptor. The energy transfer constant is defined as follows [14]:
C D - A = R C 6 τ D
In this equation, RC is the critical radius of the interaction and τD is the intrinsic lifetime. When examining phonon involvement, the constant for energy transfer can be determined using the subsequent equation [14]:
C D - A = 6 c g low D 2 π 4 n 2 g up D 0 e 2 n ¯ + 1 S 0 S 0 m m ! n ¯ + 1 m σ emis D λ m + σ abs A λ d λ
where c is the speed of light; n is the refractive index of the glass; and g low D and g up D are the simplicity of the lower and upper energy levels of the donor, respectively. ħω0 is the maximum phonon energy (1077 cm−1 in this paper), n = (1/eħω0/KT − 1) is the average phonon mode occupancy at temperature T, m is the number of phonons involved in the energy transfer, S0 is the Huang Kun factor, which has a value of 0.31 for Er3+ ions, and λ m + = (1/λmħω0) is the m phonon emission corresponding to the wavelength.
The energy transfer coefficients were analyzed by finding the TBBZ3 glass cross-retardation (CR) probability according to Equations (7) and (8). Figure 13 illustrates the emission and absorption cross-sections of Er3+:4I9/2 + 4I15/24I13/2 + 4I13/2, along with the m-phonon (m = 1) emission sidebands of the donor during the Er3+:4I9/24I13/2 transition. From the figure, it can be seen that the CR process of the prepared samples is a multi-phonon mechanism, with 0-phonon-assisted accounting for the majority (99.999%); this is a resonance process. Table 3 shows the phonon number, probability share, and microscopic parameters of the CR process in TBBZ3 glass. The process has an energy transfer coefficient CD-A of 13.8 × 10−40 cm6/s and a favorable cross-chattering process, which enhances the luminescence of 1.55 μm.

4. Discussion

Er3+-doped TeO2-Bi2O3-B2O3-ZnO glasses were prepared by melt quenching and characterized for their thermal stability, matrix structure, and spectral properties. The results show that the Er3+-doped bismuth tellurite glass has good thermal stability (ΔT of 128 °C) and low phonon energy (1077 cm−1). The intensity parameters Ω2,4,6 and other spectral parameters were obtained by Judd–Ofelt theoretical calculations. In this paper, TBBZ3 (Er3+ = 2%) glass was prepared to achieve a high spontaneous leaping chance Arad (25 s−1) at 4I11/24I13/2, as well as a high radiation lifetime τ (3.52 ms) at 4I13/24I15/2. The high-frequency band at 1077 cm−1 leads to a topological transition, giving rise to a deep topological cage structure of [BO3], [BO4]. In addition, the TBBZ3 glass achieved the highest fluorescence emission intensity and bandwidth values when pumped at 808 nm, with an emission cross-section (σemi) of 9.12 × 10−21 cm2. It also exhibited satisfactory gain performance at 1.55 μm (Gmax of 6.17 cm−1). Taken together, it shows that the Er3+-mono-doped TeO2-Bi2O3-B2O3-ZnO glass has good prospects for applications in laser materials in the near-infrared region, which can be further explored in the future with the goal of lowering the phonon energy and reducing the probability of non-radiative leaps.

5. Conclusions

In this paper, Er3+-doped TeO2-Bi2O3-B2O3-ZnO glasses were successfully prepared by melt quenching method. Testing revealed that the rare-earth-(Er3+)-doped bismuth tellurite glasses have more stable thermodynamic properties and lower maximum phonon energies. The calculated Ω2,4,6 using J-O theory indicates that the samples have superior optical properties, a higher spontaneous radiation probability, and a longer radiation lifetime (4I11/24I13/2). Furthermore, the sample exhibits a broad fluorescence emission bandwidth and an emission cross-section (9.12 × 10–21 cm2) under 808 nm pump light excitation, and it also exhibited satisfactory gain performance at 1.55 μm (Gmax of 6.17 cm−1). Consequently, the Er3+-doped TeO2-Bi2O3-B2O3-ZnO glass prepared in this study has the potential for significant applications in near-infrared materials.

Author Contributions

Conceptualization, F.T. and G.X.; data curation, F.T., S.C., D.C. and Y.B.; formal analysis, F.T., G.X., Y.Z. and D.C.; funding acquisition, Y.M.; investigation, Y.M. and B.G.; project administration, Y.M.; resources, Y.M.; software, Y.M. and Y.Z.; supervision, F.T. and D.Z.; validation, G.X.; visualization, F.T. and Y.B.; writing—original draft, F.T. and G.X.; writing—review and editing, G.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Jilin Province Science and Technology Development plan, project (20220201151GX).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bhemarajam, J.; Swapna; Babu, M.M.; Prasad, P.S.; Prasad, M. Spectroscopic studies on Er3+ ions incorporated bismuth borolead lithium glasses for solid state lasers and fiber amplifiers. Opt. Mater. 2021, 113, 110818. [Google Scholar] [CrossRef]
  2. Chatterjee, D.K.; Rufaihah, A.J.; Zhang, Y. Upconversion fluorescence imaging of cells and small animals using lanthanide doped nanocrystals. Biomaterials 2008, 29, 937–943. [Google Scholar] [CrossRef] [PubMed]
  3. Shen, X.; Zhu, Y.; Zhou, Y.; Li, J. Broadband and flat near-infrared emission from Er3+/Tm3+ co-doped tellurite glass for amplifier applications. J. Opt. Soc. Am. B 2019, 37, 320–328. [Google Scholar] [CrossRef]
  4. Haouari, M.; Maaoui, A.; Saad, N.; Bulou, A. Optical temperature sensing using green emissions of Er3+ doped fluoro-tellurite glass. Sens. Actuators A 2017, 261, 235–242. [Google Scholar] [CrossRef]
  5. Basavapoornima, C.; Maheswari, T.; Depuru, S.R.; Jayasankar, C.K. Sensitizing effect of Yb3+ ions on photoluminescence properties of Er3+ ions in lead phosphate glasses: Optical fiber amplifiers. Opt. Mater. 2018, 86, 256–269. [Google Scholar] [CrossRef]
  6. Li, K.; Fan, H.; Zhang, G.; Bai, G.; Fan, S.; Zhang, J.; Hu, L. Highly biocompatible multi-walled carbon nanotube-chitosan nano-particle hybrids as protein carriers. J. Alloys Compd. 2011, 509, 3070–3073. [Google Scholar] [CrossRef]
  7. Rachkovskaya, G.E.; Zakharevich, G.B. Formation of Low-Temperature Seals and Coatings Based on Low-Melting Lead-Tellurite Glasses. Glass Ceram. 2005, 62, 192–194. [Google Scholar] [CrossRef]
  8. Oliveira, T.A.; Manzani, D.; Falcao-Filho, E.L.; Messaddeq, Y.; Boudebs, G.; Fedus, K.; de Araújo, C.B. Near-infrared nonlinearity of a multicomponent tellurium oxide glass at 800 and 1064 nm. Appl. Phys. B 2014, 116, 1–5. [Google Scholar] [CrossRef]
  9. Tan, F.; Xu, P.; Zhou, D.; Wang, L.; Yang, Q.; Song, X.; Han, K. 1.53 μm luminescent properties and energy transfer processes of Er3+/Yb3+ co-doped bismuth germanate glass laser material. J. Lumin. 2021, 239, 118300. [Google Scholar] [CrossRef]
  10. Xu, X.; Zhou, Y.; Zheng, S.; Yin, D.; Wang, X. Luminescence properties and energy transfer mechanism of Er3+/Tm3+ co-doped tellurite glasses. J. Alloys Compd. 2013, 556, 221–227. [Google Scholar] [CrossRef]
  11. Deopa, N.; Sahu, M.K.; Kaur, S.; Prasad, A.; Swapna, K.; Kumar, V.; Punia, R.; Rao, A.S. Enhanced visible green and 1.5 μm radiative emission of Er3+ions in Li2O-PbO-Al2O3-B2O3 glasses for photonic applications. J. Rare Earths 2021, 39, 520–525. [Google Scholar] [CrossRef]
  12. Sidek, H.; Rosmawati, S.; Azmi, B.Z.; Shaari, A.H. Effect of ZnO on the Thermal Properties of Tellurite Glass. Adv. Condens. Matter Phys. 2013, 2013, 709–712. [Google Scholar] [CrossRef]
  13. Silva, L.; Torquato, A.; Assis, I.; Dousti, M.R. Spectroscopic study of Er3+-doped zinc-tellurite glass and opaque glass-ceramic. Solid State Sci. 2020, 112, 106444. [Google Scholar] [CrossRef]
  14. Song, X.; Han, K.; Zhou, D.; Xu, P.; Xue, X.; Zhang, P. 2 μm emission properties and energy transfer processes in Tm3+ doped Bi2O3-GeO2-Na2O glass laser material. J. Lumin. 2020, 224, 117314. [Google Scholar] [CrossRef]
  15. Yin, D.; Zhou, Y.; Zheng, S.; Xu, X.; Wang, X.; Dai, S. Near-Infrared Spectroscopic Properties and Energy Transfer Mechanism of Er3+/Tm3+-Codoped TeO2-ZnO2-Na2O Glasses. Chin. E J. Lasers 2013, 40, 0906001. [Google Scholar]
  16. Yang, X.L.; Wang, W.C.; Zhang, Q.Y. BaF2 modified Cr3+/Ho3+ co-doped germanate glass for efficient 2.0 μm fiber lasers. J. Non-Cryst. Solids 2018, 482, 147–153. [Google Scholar] [CrossRef]
  17. Shi, D.-m.; Zhao, Y.-g. Structure and Spectroscopic Properties of Er3+-doped GeO2-TeO2-Na2O Glass. Chin. J. Lumin. 2018, 1352–1358. [Google Scholar]
  18. Song, C.; Xu, P.; Zhou, D.; Zhang, K.; Ma, B.; Cong, Y.; Bai, Y.; Cao, J.; Han, K. Effect of phonon energy on 2 μm emission in Tm3+/Yb3+ co-doped bismuth-tellurite glasses based on WO3 or B2O3 adjustment. J. Lumin. Interdiscip. J. Res. Excit. State Process. Condens. Matter 2023, 255, 119610. [Google Scholar] [CrossRef]
  19. Liu, Z.; She, J.; Peng, B. Spectroscopic properties of Er3+-doped fluoroindate glasses. J. Rare Earths 2022, 40, 1037–1042. [Google Scholar] [CrossRef]
  20. Chen, L.; Lai, Y. Thermal, structural, and metallization properties of 0.95[50B2O3-30BaO-(20−x)ZnO-xBi2O3]+0.05V2O5 glasses. J. Non-Cryst. Solids 2020, 556, 120552. [Google Scholar] [CrossRef]
  21. Jose, R.; Arai, Y.; Ohishi, Y. Raman scattering characteristics of the TBSN-based tellurite glass system as a new Raman gain medium. J. Opt. Soc. Am. B 2007, 24, 1517–1526. [Google Scholar] [CrossRef]
  22. Berwal, N.; Dhankhar, S.; Sharma, P.; Kundu, R.S.; Punia, R.; Kishore, N. Physical, structural and optical characterization of silicate modified bismuth-borate-tellurite glasses. J. Mol. Struct. 2017, 1127, 636–644. [Google Scholar] [CrossRef]
  23. Kotkova, K.; Ticha, H.; Tichy, L. Raman studies and optical properties of some (PbO)x(Bi2O3)0.2(B2O3)0.8x glasses. J. Raman Spectrosc. 2008, 39, 1219–1226. [Google Scholar] [CrossRef]
  24. Rada, S. Structure of TeO2·B2O3 glasses inferred from infrared spectroscopy and DFT calculations. J. Non-Cryst. Solids 2008, 354, 5491–5495. [Google Scholar] [CrossRef]
  25. Ardelean, I.; Cora, S.; Ciceo-Lucacel, R. Structural Investigations of CuO–B2O3–Bi2O3 Glasses by Means of EPR and FT-IR Spectroscopies. Mod. Phys. Lett. B 2004, 18, 803–810. [Google Scholar] [CrossRef]
  26. Lakshminarayana, G.; Baki, S.O.; Lira, A.; Kityk, I.V.; Mahdi, M.A. Structural, thermal, and optical absorption studies of Er3+, Tm3+, and Pr3+-doped borotellurite glasses. J. Non-Cryst. Solids 2017, 459, 150–159. [Google Scholar] [CrossRef]
  27. Saritha, D.; Markandeya, Y.; Salagram, M.; Vithal, M.; Singh, A.K.; Bhikshamaiah, G. Effect of Bi2O3 on physical, optical and structural studies of ZnO-Bi2O3-B2O3 glasses. J. Non-Cryst. Solids 2008, 354, 5573–5579. [Google Scholar] [CrossRef]
  28. Zhou, Y.; Yang, Y.; Huang, F.; Ren, J.; Yuan, S.; Chen, G. Characterization of new tellurite glasses and crystalline phases in the TeO2–PbO–Bi2O3–B2O3 system. J. Non-Cryst. Solids 2014, 386, 90–94. [Google Scholar] [CrossRef]
  29. Tao, J.; Chen, X.; Lin, W.; Zhou, G.; Zhao, G.; Chen, L. Calculation and analysis on J-O spectral parameters of tellurate luminescent glass doped Er3+. Chin. J. Electron. 2018, 35, 257–263. [Google Scholar]
  30. Zhang, J.; Xi, Z.; Wang, X.; Long, W.; Cui, M.; Zou, M. Spectrum Properties and Judd-Ofelt Analysis of PMN-PT: Er3+ Crystal. J. Synth. Cryst. 2020, 1023–1029+1056. [Google Scholar]
  31. Aouaini, F.; Maaoui, A.; Mohamed, N.; Alanazi, M.M.; El Maati, L.A. Visible to infrared down conversion of Er3+ doped tellurite glass for luminescent solar converters. J. Alloys Compd. 2022, 894, 162506. [Google Scholar] [CrossRef]
  32. Jlassi, I.; Elhouichet, H.; Ferid, M.; Barthou, C. JuddOfelt analysis and improvement of thermal and optical properties of tellurite glasses by adding P2O5. J. Lumin. 2010, 130, 2394–2401. [Google Scholar] [CrossRef]
  33. Fan, Y.; Hu, J.; Huang, Y.; Wang, D. 2 μm Luminescence Properties of Bismuth Tellurite Glass Co-doped with Ho3+\Yb3+\Tm3+\Ce3+. J. Mater. Sci. Eng. 2020, 6, 917–923. [Google Scholar]
  34. Pugliese, D.; Boetti, N.G.; Lousteau, J.; Ceci-Ginistrelli, E.; Bertone, E.; Geobaldo, F.; Milanese, D. Concentration quenching in an Er-doped phosphate glass for compact optical lasers and amplifiers. J. Alloys Compd. 2016, 657, 678–683. [Google Scholar] [CrossRef]
  35. Tian, Y.; Xu, R.; Hu, L.; Zhang, J. Spectroscopic properties and energy transfer process in Er3+ doped ZrF4-based fluoride glass for 2.7 mm laser materials. Opt. Mater.-Amst. 2011, 34, 308–312. [Google Scholar] [CrossRef]
  36. Li, X.; Liu, X.; Zhang, L.; Hu, L.; Zhang, J. Emission enhancement in Er3+/Pr3+-codoped germanate glasses and their use as a 2.7-mm laser material. Chin. Opt. Lett. 2013, 11, 12160101–12160103. [Google Scholar]
  37. Huang, F.; Zhang, Y.; Hu, L.; Chen, D. Judd-Ofelt analysis and energy transfer processes of Er3+ and Nd3+ doped fluoroaluminate glasses with low phosphate content. Opt. Mater. 2014, 38, 167–173. [Google Scholar] [CrossRef]
  38. Fang, Y.; Zhao, G.; Xu, J.; Zhang, N.; Ma, Z.; Hu, L. Energy transfer and 1.8 μm emission in Yb3+/Tm3+ co-doped bismuth germanate glass. Ceram. Int. 2014, 40, 6037–6043. [Google Scholar] [CrossRef]
  39. Miguel, A.; Morea, R.; Gonzalo, J.; Arriandiaga, M.A.; Fernández, J.; Balda, R. Near-infrared emission and upconversion in Er3+-doped TeO2-ZnO-ZnF2 glasses. J. Lumin. 2013, 140, 38–44. [Google Scholar] [CrossRef]
  40. Ouannes, K.; Soltani, M.; Boulon, G.; Alombert-Goget, G.; Guyot, Y.; Pillonnet, A.; Lebbou, K. Spectroscopic properties of Er3+-doped antimony oxide glass. J. Alloys Compd. Interdiscip. J. Mater. Sci. Solid State Chem. Phys. 2014, 603, 132–135. [Google Scholar] [CrossRef]
  41. Huang, F.; Guo, Y.; Tian, Y.; Xu, S.; Zhang, J. Intense 2.7 μm emission in Er3+ doped zinc fluoride glass. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2017, 179, 42–45. [Google Scholar] [CrossRef]
  42. Zhan, H.; Zhou, Z.; He, J.; Lin, A. Intense 2.7 μm emission of Er3+-doped water-free fluorotellurite glasses. Opt. Lett. 2012, 37, 3408–3410. [Google Scholar] [CrossRef]
  43. Yang, J.-X.; Zhao, X.; Zhang, J.-H.; Lin, H. 1.2 m Near Infrared Fluorescence Emission of Holmium Ions in Thermal Ion-exchanged Aluminum Germanate Waveguide Glasses. Chin. J. Lumin. 2018, 1519–1526. [Google Scholar]
  44. Klinkov, V.A.; Tsimerman, E.A.; Rokhmin, A.S.; Andreeva, V.D.; Tagil’tseva, N.O.; Babkina, A.N. 1.53 μm luminescent properties of Er3+-doped fluoroaluminate glasses. Opt. Mater. 2021, 121, 111585. [Google Scholar] [CrossRef]
  45. Sajna, M.S.; Thomas, S.; Jayakrishnan, C.; Joseph, C.; Biju, P.R.; Unnikrishnan, N.V. NIR emission studies and dielectric properties of Er3+-doped multicomponent tellurite glasses. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2016, 161, 130. [Google Scholar] [CrossRef] [PubMed]
Figure 1. DSC curve of TBBZ host glass.
Figure 1. DSC curve of TBBZ host glass.
Materials 17 03292 g001
Figure 2. The XRD pattern of TBBZ glass sample.
Figure 2. The XRD pattern of TBBZ glass sample.
Materials 17 03292 g002
Figure 3. Transmission spectra of TBBZ1-5 glass sample.
Figure 3. Transmission spectra of TBBZ1-5 glass sample.
Materials 17 03292 g003
Figure 4. Schematic diagram of topological cage structure evolution of TBBZ glass network.
Figure 4. Schematic diagram of topological cage structure evolution of TBBZ glass network.
Materials 17 03292 g004
Figure 5. The Raman spectra of TBBZ host glass.
Figure 5. The Raman spectra of TBBZ host glass.
Materials 17 03292 g005
Figure 6. FT-IR spectra of TBBZ host glass.
Figure 6. FT-IR spectra of TBBZ host glass.
Materials 17 03292 g006
Figure 7. The absorption spectra of Er3+-doped tellurite glasses.
Figure 7. The absorption spectra of Er3+-doped tellurite glasses.
Materials 17 03292 g007
Figure 8. (a) The 1.55 µm fluorescence spectra of TBBZ1-5 glasses; (b) 2.7 µm fluorescence spectra of TBBZ1-5 glasses.
Figure 8. (a) The 1.55 µm fluorescence spectra of TBBZ1-5 glasses; (b) 2.7 µm fluorescence spectra of TBBZ1-5 glasses.
Materials 17 03292 g008
Figure 9. Fluorescence decay curves of the Er3+:4I13/2 level in TBBZ3 glass.
Figure 9. Fluorescence decay curves of the Er3+:4I13/2 level in TBBZ3 glass.
Materials 17 03292 g009
Figure 10. Absorption and emission cross-sections of Er3+:4I13/24I15/2 transition in TBBZ3 glass.
Figure 10. Absorption and emission cross-sections of Er3+:4I13/24I15/2 transition in TBBZ3 glass.
Materials 17 03292 g010
Figure 11. Gain coefficient spectra of TBBZ3 glass.
Figure 11. Gain coefficient spectra of TBBZ3 glass.
Materials 17 03292 g011
Figure 12. Energy level transition diagram of Er3+.
Figure 12. Energy level transition diagram of Er3+.
Materials 17 03292 g012
Figure 13. Emission and absorption cross-section of Er3+ at 4I9/2 + 4I15/24I13/2 + 4I13/2 level and edge band diagram of m phonon (m = 1) emission in 4I9/24I13/2 transition.
Figure 13. Emission and absorption cross-section of Er3+ at 4I9/2 + 4I15/24I13/2 + 4I13/2 level and edge band diagram of m phonon (m = 1) emission in 4I9/24I13/2 transition.
Materials 17 03292 g013
Table 1. Judd–Ofelt strength parameters for erbium ions in the studied tellurite glass.
Table 1. Judd–Ofelt strength parameters for erbium ions in the studied tellurite glass.
GlassΩ2 (×10−20 cm2)Ω4 (×10−20 cm2)Ω6 (×10−20 cm2)Ref.
TBBZ14.311.441.40This work
TBBZ23.541.191.14This work
TBBZ33.131.051.0This work
TBBZ42.800.940.9This work
TBBZ52.570.860.82This work
HYTC0.62.841.561.82[33]
Tellurite1.881.981.39[34]
Fluoride3.081.461.69[35]
Table 2. Calculated spontaneous transition probability (Arad), total spontaneous transition probability (∑A), branching ratios (β), and radiative lifetime (τrad) radiative properties of TBBZ3 glass sample for various selected excited states of Er3+.
Table 2. Calculated spontaneous transition probability (Arad), total spontaneous transition probability (∑A), branching ratios (β), and radiative lifetime (τrad) radiative properties of TBBZ3 glass sample for various selected excited states of Er3+.
Transitionλ (nm)Arad (s−1)∑A (s−1)β (%)τrad (ms)
4I13/24I15/21533284.41284.41100.003.52
4I11/24I13/2267824.61612.254.021.63
4I15/2975587.65 95.98
4I9/24I11/243951.05922.910.111.08
4I13/2166465.37 7.08
4I15/2798856.49 92.80
Table 3. Micro-parameters of cross-relaxation processes of Er3+ in TBBZ3 glass.
Table 3. Micro-parameters of cross-relaxation processes of Er3+ in TBBZ3 glass.
GlassETPhonon Number (N) and
Contribution Radio (%)
Transfer Constant
(10−40 cm6/s)
TBBZE3CR01CD-A = 13.8
99.9990.001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tan, F.; Xie, G.; Ma, Y.; Zhang, Y.; Gao, B.; Cui, S.; Chen, D.; Ban, Y.; Zhou, D. Structure and Spectral Properties of Er3+-Doped Bismuth Telluride Near-Infrared Laser Glasses. Materials 2024, 17, 3292. https://doi.org/10.3390/ma17133292

AMA Style

Tan F, Xie G, Ma Y, Zhang Y, Gao B, Cui S, Chen D, Ban Y, Zhou D. Structure and Spectral Properties of Er3+-Doped Bismuth Telluride Near-Infrared Laser Glasses. Materials. 2024; 17(13):3292. https://doi.org/10.3390/ma17133292

Chicago/Turabian Style

Tan, Fang, Guoxing Xie, Yuqin Ma, Yunlong Zhang, Binhao Gao, Shunfa Cui, Dexiao Chen, Yumeng Ban, and Dechun Zhou. 2024. "Structure and Spectral Properties of Er3+-Doped Bismuth Telluride Near-Infrared Laser Glasses" Materials 17, no. 13: 3292. https://doi.org/10.3390/ma17133292

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