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

Improved 2 µm broadband luminescence in Tm3+/Ho3+ doping tellurite glass

Open Access Open Access

Abstract

Tm3+/Ho3+ doping tellurite glasses (TeO2-ZnO-La2O3) were prepared by applying melt-quenching technique, and the ∼2.0 µm band luminescence characteristics were examined. A broadband and relatively flat luminescence at 1600 to 2200 nm was observed in the tellurite glass co-doped by 1.0 mol% Tm2O3 and 0.085 mol% Ho2O3 under the excitation of 808 nm laser diode (LD), which is the result of spectral overlapping of 1.83 µm band of Tm3+ ions and 2.0 µm band of Ho3+ ions. Further, about 103% enhancement was acquired after the introduction of 0.1 mol% CeO2 and 7.5 mol% WO3 at the same time, which is primarily caused by the cross-relaxation between Tm3+ and Ce3+ ions together with the enhanced energy transfer from the Tm3+:3F4 level to Ho3+:5I7 level due to the increase in phonon energy. Spectral characteristics associated with the radiative transition of Ho3+ and Tm3+ ions on the basis of Judd-Ofelt theory, and the fluorescence decay behaviors after the addition of Ce3+ ions and WO3 component were analyzed to understand the broadband and luminescence enhancement. The findings in this work indicate that tellurite glass with optimal Tm3+-Ho3+-Ce3+ tri-doping combination and appropriate amount of WO3 is a prospective candidate for broadband optoelectronic devices operated in the infrared bands.

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

1. Introduction

Owing to the rich energy level structure, trivalent rare-earth ions can achieve luminescence with wavelengths ranging from ultraviolet (UV), visible to near/mid-infrared (NIR/MIR) bands under an appropriate optical excitation. Consequently, the study of luminescence characteristics of glass materials doped with rare earths has been the hotspot in developing optoelectronic devices like fiber-optic amplifiers, illumination displays and solid-state lasers. Meanwhile, because of the rich energy level structure, certain energy levels of various rare earth ions have comparable energy heights, and/or the energy gap between two energy levels of one ion is close to that of another ion. These characteristics make the co-doping of various rare earth ions become an efficient technical method to improve the luminescence performance of glass materials. With the co-doping of two rare earth ions for instance, it can make an enhancement in the luminescent intensity of the acceptor ion through the sensitization mechanism, or broaden the luminescent spectrum through the spectral overlapping of two adjacent luminescent bands from the co-doping ions. In addition, due to the selection rule of energy level transition, rare earth ions have strong absorption at some bands such as Er3+ ions at the 1480 nm band [1], Yb3+ ions at the 980 nm band [2], and Tm3+ ions at the 800 nm band [3], just corresponding to the working wavelengths of commercial laser diodes (LDs). Therefore, under the excitation of appropriate commercial lasers, co-doping with these rare earth ions can be expected to improve the pumping efficiency.

As a band between the MIR and the NIR regions, 2.0 µm band laser has wide and potential applications in environmental detection, biomedicine, remote sensing communication and other areas [46]. In the aspect of environmental monitoring, for instance, the main components of the atmosphere, like CO2, H2O, CO and N2O, have characteristic absorption peaks in the range of 2∼3 µm, so using this band laser can monitor the change of these gas contents in the environment. In the biomedicine, water molecules have a high absorption coefficient (∼100 cm-1) near the 2.0 µm band. When using 2.0 µm band laser to act on biological tissues, a large amount of heat will be generated so that the diseased tissues can be quickly removed. It has the advantages of high cutting accuracy, small postoperative wound, and good hemostasis. While in the remote sensing communication, laser in the 2.0 µm band possesses high penetration capability to the atmosphere because it is located in the one of atmospheric transmission window (1∼3 µm), so it can be directly applied to coherent Doppler wind radar and laser ranging; In addition, because of the inherent Rayleigh scattering effect of materials, if 2.0 µm band is used as the laser working wavelength instead of 1.55 µm band, the fiber loss caused by Rayleigh loss can be effectively reduced and the distance of relay free communication can be increased accordingly.

Currently, Ho3+ is among the most efficient rare earth ions in producing lasing emission at 2.0 µm band. Its 5I75I8 radiative transition with 2.0 µm band emission has a long fluorescence lifetime and great stimulated emission cross section. Nevertheless, so far, no commercial LD is available on the market that can excite Ho3+ directly to produce fluorescence at the 2.0 µm band. Amongst the trivalent rare earth ions, Tm3+ can be pumped by commercially 808 nm LD to emit strong 1.83 µm band lasing emission through the 3F43H6 radiative transition, and also the pump energy absorbed can be transferred to Ho3+:5I7 level from the Tm3+:3F4 level to make the latter produce fluorescence [7]. In recent years, there are many reports on the adopting of Tm3+/Ho3+ co-doping scheme to achieve ∼2.0 µm band broadband luminescence. For example, J. Kang et al. [8] prepared Tm3+/Ho3+ co-doped lanthanum aluminosilicate glass and photonic crystal fiber (PCF), and a broadband luminescence in the range of 1600 to 2200 nm was observed in the glass while an optical-to-optical efficiency 6.37% of this broadband output spectrum was obtained in the PCF. The broadband luminescence characteristics and laser performance were investigated. SK. Taherunnisa et al. [9] synthesized Tm3+/Ho3+ co-doped sodium-sulfo lead phosphate glasses, and the NIR and MIR photoluminescence spectra were recorded at 797 nm excitation. The studies of luminescence properties of Tm3+:1.8 µm and Ho3+:2.0 µm bands with doped concentrations suggested that Tm3+/Ho3+ co-doped phosphate glass is advantageous to generate 1600–2200 nm broadband luminescence. N.M. Ty et al. [10] reported the broadband flat NIR/MIR emissions in Tm3+–Ho3+ co-doped and Tm3+–Ho3+–Yb3+ tri-doped zinc silicate glasses under excitation of 808 and 980 nm LD respectively. Broadband relatively flat fluorescence emissions in the range of 1600 to 2200 nm in Tm3+–Ho3+ co-doped and Tm3+–Ho3+–Yb3+ tri-doped glasses were identified, and the influences of Ho3+ and Yb3+ concentrations on the broadband luminescence properties were investigated. Obviously, the realization of broadband luminescence is mainly attributed to the energy transfer from Tm3+:3F4 level to Ho3+:5I7 level. However, an energy difference exists between these two levels, so the energy transfer between them needs to be completed by phonon assistance (Tm3+:3F4 → Ho3+:5I7 + phonons), and the phonon energy of glass host has an important influence on the completion of energy transfer.

In this paper, trivalent Ce3+ ions and WO3 component with moderate phonon energy were introduced into Tm3+/Ho3+ co-doping tellurite glass, and the improved effect of Ce3+ tri-doping and WO3 introduction on the broadband luminescence of ∼2.0 µm band was reported. The broadband luminescence mechanism of Tm3+/Ho3+ co-doping, and the cross-relaxation process between Ce3+ and Tm3+ ions as well as the variation of host phonon energy induced by WO3 component which are responsible for the luminescence properties were studied. In comparison to traditional glasses like silica glass, silicate glass, aluminosilicate glass and borate glass, tellurite glass possesses some unique advantages [1113], such as high refractive index (∼2.0), large rare earth solubility (several time of silica glass), wide visible-to-infrared transmission window (∼0.4–5 µm) and low maximum phonon energy (∼750 cm-1). High refractive index can bring large emission cross-section and transition probability, large rare earth solubility can provide high gain per unit length, and low phonon energy can improve radiative quantum efficiency. All these make tellurite glass a potential host material for trivalent rare earth ions.

2. Experimental procedures

In this work, the melt-quenching technique was applied to prepare the doping tellurite glasses, and the detailed component amounts and corresponding sample names were listed in Table 1. To prepare glass sample, accurately weigh 15 g oxide powders as a batch according to the proportion in the table with a precision balance, pour it into the platinum crucible and stir it evenly. The platinum crucible was then heated to about 320 °C for 30 min with an aim of removing residual moisture from the powders. It was next transferred to a resistance furnace and melted in a dry atmosphere at 900 °C for 45 min, stirred by a quartz glass rod for 10 min in melting process to remove air bubbles generated and also to make molten liquid more homogeneous. Afterwards, the molten liquid was poured into a preheated silica mold and transferred instantly to a 350 °C annealing furnace for 2 hours for the release of thermal stress. Finally, the glass sample was polished into thin disc-shaped block with 1.6 mm thickness and 10 mm diameter for physical and spectroscopic tests after cooling to the ambient temperature.

Tables Icon

Table 1. The compositions (mol %) of tellurite glass and corresponding name

The density of glass sample was determined based on the Archimedes principle by using the distilled water as an immersion solution. Refractive index was measured at 632.8 nm by prism coupler of Sialon Technology. X-ray diffractogram (XRD) pattern was obtained by Bruker D8 Advance power diffractometer from 10° to 70° (2θ) under a Cu-Ka radiation. Raman spectrum was obtained by Raman microscope of Renishaw inVia. Thermal stability was evaluated by differential scanning calorimeter (DSC) of TA Q2000 instrument. Absorption spectrum between 400 and 2200 nm was acquired by Perkin Elmer Lambda 950 UV-Vis-NIR spectrophotometer. Fluorescence spectrum in VIS-NIR range was acquired by FLSP920 fluorescence spectrometer of Edinburgh Instruments at excitation of 808 nm laser diode (LD). A digital oscilloscope was used to obtain fluorescence decay curve under 808 nm pulse excitation utilizing the same testing system. The measured conditions for all glass samples were kept the same so as to get a comparable result, and the physical parameters and the morphology of tellurite glass samples obtained are shown in Table 2.

Tables Icon

Table 2. Density ($\rho$), refractive index ($n$), doping concentrations of Tm3+ (${N_{\textrm{Tm}}}$), Ho3+ (${N_{\textrm{Ho}}}$) and Ce3+ (${N_{\textrm{Ce}}}$), and photo of the prepared glass sample.

3. Results and discussion

3.1 XRD pattern, DSC curve and Raman spectrum

Figure 1 displayed the XRD patterns of tellurite glass samples with Tm3+/Ho3+ co-doping (T-TH0 and T-TH3) and Tm3+/Ho3+/Ce3+ tri-doping (T-THC0 and T-THC3) with and without WO3 component. As seen from the figure that in the diffraction angle range (2θ) of 10–70°, the XRD profiles of glass samples looks almost the same which have the common characteristics, i.e. there is no sharp diffraction peaks but an intense hump located at 15–40°, indicating that there is no periodicity of atom arrangement in three dimensional space. This means that the physical structure of synthesized tellurite glasses is in amorphous state [14,15]. As the representative, inset of Fig. 1 displayed the DSC curves of glass samples T-TH0 and T-TH3. Generally, the thermal stability of glass hosts is characterized by difference $\Delta T$ (=${T_\textrm{x}}$${T_\textrm{g}}$) between the crystallization onset temperature (${T_\textrm{x}}$) and transition temperature (${T_\textrm{g}}$) [16]. If the difference $\Delta T$ is greater than 100 °C, it implies that the glass host has good glass-forming ability and anti-crystallization stability, and can allow a wide working temperature range during fiber drawing. Therefore, it can be concluded from the figure that the present studied tellurite glasses possess good thermal stability.

 figure: Fig. 1.

Fig. 1. XRD patterns of the synthesized tellurite glass sample, and inset is the DSC curves of T-TH0 and T-TH3 tellurite glasses.

Download Full Size | PDF

Figure 2 displayed the Raman spectra of four tellurite glass samples T-TH0, T-TH3, T-THC0 and T-THC3 in the spectral range from 200 to 1000 cm-1. The scattering peaks of T-TH0 and T-THC0 glass samples without WO3 component appear at three positions of about 390, 670 and 761 cm-1. Among which, the 390 cm-1 scattering peak is caused by the bending vibration of Zn-O bonds from ZnO4 unit [17,18] together with the symmetric tensile vibration of Te-O-Te or O-Te-O bonds from TeO4 triangular double pyramid (tbp) unit [19]. The scattering peak at 690 cm-1 corresponds to the asymmetric stretching vibration of Te-O bonds from TeO4 tbp unit [20] and the 761 cm-1 scattering peak arises from the stretching vibration of non-bridged Te and O atoms from TeO3 + 1 polyhedron and TeO3 triangular pyramid units [21]. It is seen that the scattering peak at 390 cm-1 in T-TH3 and T-THC3 tellurite glass samples with WO3 component becomes stronger and this is attributed to the addition of tensile and flexural vibration of W-O-W bonds related to the octahedral unit of WO6 [22]. Besides these three scattering peaks, in both glass samples a new scattering peak is found at about 916 cm-1, and this scattering peak originates from the bending vibration of W = O bond from WO4 tetrahedron unit together with the tensile vibration of W-O- bond from WO6 octahedron unit [16]. Obviously, the vibrational phonon energy of the tellurite glass becomes large after the introduction of WO3 component, but is still smaller compared with that of the traditional oxide glasses like silicate glass (∼1050 cm-1) [23], phosphate glass (∼1250 cm-1) [24] and borate glass (∼1300 cm-1) [25].

 figure: Fig. 2.

Fig. 2. Raman spectra of tellurite glass samples with and without WO3 component.

Download Full Size | PDF

3.2 Absorption spectrum and Judd-Ofelt analysis

Figure 3 displayed the absorption spectra of tellurite glass samples in the range of 400–2000nm and the samples selected as the representative are Ho3+ and Tm3+ single-doping (T-Ho and T-Tm), Tm3+/Ho3+ co-doping (T-TH0 and T-TH3) as well as Tm3+/Ho3+/Ce3+ tri-doping (T-THC3) respectively. The results obtained are similar to those of other glass hosts [26]. All absorption bands exhibited in the spectra originate from the transitions of f-f electrons from the ground states to respective excited levels. The bands observed at wavelengths 470, 685, 795, 1212 and 1665 in the Tm3+ single-doping sample are attributed to the absorption transitions from the ground state 3H6 to excited levels 1G4, 3F3(3F2),3H4, 3H5 and 3F4 of Tm3+ respectively [27]. The bands located at wavelengths 1955, 1150, 645, 540, 485, 458 and 420 nm in the Ho3+ single-doping sample originate from the absorption transitions from the ground state 5I8 to excited levels5I7, 5I6, 5F5, 5S2 (5F4), 5F3, 5F1 (5G6) and 3G5 (5G5) of Ho3+ respectively [28]. Because the absorption band of Ce3+ (2F5/22F7/2) is positioned at MIR region of about 4650 nm [29], thus the absorption peak of Ce3+ does not displayed in the figure. However, owing to the strong absorption induced by the inter-configuration transition of 4f→5d (2F5/25D1) of Ce3+ at about 400–500 nm [30], the UV absorption edge in the glass sample with Ce3+ is red-shifted to a much longer wavelength. Moreover, it is noted that the shapes and positions of all absorption peaks are almost unchanged in these samples. This observation implies that the three ions are homogeneously distributed in tellurite glass network.

 figure: Fig. 3.

Fig. 3. Absorption spectra of tellurite glass samples doped with Tm3+, Ho3+ and Ce3+.

Download Full Size | PDF

Besides the absorption transitions observed, the radiative transition characteristics of doped rare earth ions can be evaluated from the absorption spectrum. Based on the J-O (Judd-Ofelt) theory [31,32], some significant spectral parameters, for instance, the fluorescence branching ratio, radiative transition probability and radiative lifetime for the transition from upper level to lower level can be determined. The detailed description of J-O theory can be found in Refs. [33,34]. First, applying the least mean square fitting algorithm to the theoretical and the experimental oscillator strengths, the three J-O intensity parameters (${\Omega _2}$, ${\Omega _4}$ and ${\Omega _6}$) are determined and the obtained results of Tm3+ ions in Tm3+ single-doping glass sample (T-Tm) are 5.26, 1.45 and 1.84 × 10−20 respectively, while those of Ho3+ ions in Ho3+ single-doping glass sample (T-Ho) are 5.21, 2.81 and 1.72 × 10−20 respectively. The parameter ${\Omega _2}$ mainly reflect the short-range coordination effect [35], because it is associated with the covalent nature of chemical bonds between the ligand anions and rare earth cations, together with the local environment symmetry around rare earth sites. When the value of ${\Omega _2}$ increases, the covalent bond rises in number and the local environment of rare earth ions becomes more asymmetrical. The parameters ${\Omega _4}$ and ${\Omega _6}$ reveal the long-range effect of doped glass [36], wherein the hardness and viscosity of glass matrix are mainly associated with ${\Omega _6}$. In this work, the relatively large ${\Omega _2}$ values of Tm3+ and Ho3+ ions imply that the local symmetry around rare earth ions is lower while the covalent character of the chemical bonds between the rare earth cations and ligand anions is relatively higher.

Further, using the relevant formulas introduced in Refs. [33,34], the radiative transition probability, fluorescence branching ratio and radiative lifetime can be determined based on the values of three intensity parameters. The spectral parameters calculated for the partial lower-excited levels of Tm3+ and Ho3+ ions are summarized in Tables 3 and 4 respectively. It is well known that high radiative transition probability is necessary to obtain strong fluorescence emission. For the two NIR band emissions at 1.83 and 2.0 µm studied in this paper, that is, the Tm3+:3F43H6 and Ho3+:5I75I8 radiative transitions, the radiative probabilities are about 505.11 and 232.34 s-1 respectively. Among them, the radiative probability of Tm3+:3F43H6 transition is higher than that of the germanate glass (∼398.29 s-1) [37], gallate glass (∼457.25 s-1) [38], while that of Ho3+:5I75I8 transition is higher in comparison to that of germanate glass (∼177.70 s-1) [39] as well as fluorate glass (∼103.02 s-1) [40]. This indicates that tellurite glass doped with Tm3+ and Ho3+ ions in this work has great potential application prospect in the ∼2.0 µm band.

Tables Icon

Table 3. The electric- and magnetic-dipole radiative transition probabilities (${A_{\textrm{ed}}}$, ${A_{\textrm{md}}}$), branching ratio ($\beta$) and lifetime (${\tau _{\textrm{rad}}}$) of partial energy levels in Tm3+ doped glass sample.

Tables Icon

Table 4. The electric- and magnetic-dipole radiative transition probabilities (${A_{\textrm{ed}}}$, ${A_{\textrm{md}}}$), branching ratio ($\beta$) and lifetime (${\tau _{\textrm{rad}}}$) of partial energy levels in Ho3+ doped glass sample.

3.3 Fluorescence emission and luminescence mechanism

Figures 4 and 5 displayed the fluorescence emission spectra of Tm3+/Ho3+ co-doping and Tm3+/Ho3+/Ce3+ tri-doping tellurite glass samples with various amounts of WO3 from 1200 to 2200 nm and from 620 to 740 nm spectral ranges respectively under the excitation of 808 nm LD. It can be seen that the co-doping and tri-doping tellurite glass samples have the similar luminescence properties. On the one hand, 705 nm, 1.47 µm, 1.83 µm and 2.0 µm four bands are observed in the measured wavelength range, and they originate from Tm3+:3F2,33H6, 3H43F4,3F43H6 and Ho3+:5I75I8 radiative transitions [4144] respectively. It should be noted that the appearance of fluorescence of Ho3+ in the 2.0 µm band is caused by the energy transfer from Tm3+ to Ho3+ ions which will be discussed later, because Ho3+ does not have an absorption level that is compatible with the 808 nm LD. Interestingly, this 2.0 µm band fluorescence of Ho3+ ions overlaps the 1.83 µm band fluorescence of Tm3+ ions, as a result, a broadband and relatively flat fluorescence emission spectrum between 1600 and 2200 nm is formed. On the other hand, after introducing WO3 component, the fluorescence intensities of 1.47, 1.83 and 2.0 µm bands are enhanced to the different degrees until the amount of WO3 reaches about 7.5 mol%, while the intensity at 705 nm band fluorescence exhibits an opposite trend. As discussed in section 3.1, the introduction of WO3 component brings the increase of vibrational phonon energy of glass network, and such increase of phonon energy is responsible for the luminescence phenomena observed. Thirdly, for both NIR bands of 2.0 µm and 1.83 µm that we are interested in, their fluorescence intensity in the tellurite sample with Tm3+/Ho3+/Ce3+ tri-doping is noticeably stronger compared to that in the sample with Tm3+/Ho3+ co-doping, and this is attributed to the cross relaxation among Ce3+ and Tm3+ ions.

 figure: Fig. 4.

Fig. 4. Fluorescence spectra in the range of (a) 1200-2200 nm and (b) 620-740 nm in Tm3+/Ho3+ co-doping tellurite glass samples with different WO3 amounts under 808 nm LD excitation.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Fluorescence spectra in the range of (a) 1200-2200 nm and (b) 620-740 nm in Tm3+/Ho3+/Ce3+ tri-doping tellurite glass samples with different WO3 amounts under 808 nm LD excitation.

Download Full Size | PDF

The luminescence observed under 808 nm LD excitation can be understood by means of a schematic diagram which includes the interactions and transitions of Ce3+, Ho3+ and Tm3+ ions as displayed in Fig. 6. First, through the ground state absorption (GSA), Tm3+ ions at 3H6 ground state are excited to the pumping level of 3H4, where a few of them radiate directly to the level 3F4 accompanying with a faint 1.47 µm band fluorescence due to the inherent small fluorescence branching ratio. The majority of ions at 3H4 level quickly return to the lower3F4 and 3H5 levels via the cross relaxation (CR1, Tm3+:3H4 + Tm3+:3H6 → Tm3+:3F4 + Tm3+:3F4) and multi-phonon relaxation (MPR) processes. A fraction of Tm3+ ions at 3H5 level keep decaying to the 3F4 level through the MPR process and meanwhile transfer the energy to the Ho3+:5I6 level (ET1, Tm3+:3H5 + Ho3+:5I8 → Tm3+:3H6 + Ho3+:5I6), while the remainder are excited to a higher level 1G4 via the excited state absorption (ESA, Tm3+:3H5 + a photon → Tm3+:1G4). After they return to the level 3F2,3 by a process of MPR, the radiative transition to ground state 3H6 produces a fluorescence in 705 nm band. Lastly, at the level 3F4, the Tm3+ ions radiate to the ground state 3H6 and generate a strong fluorescence in 1.83 µm band. In the Tm3+/Ho3+ co-doping, energy transfer can easily take place since the energy differences between the certain energy levels of Tm3+ and Ho3+ ions are small, for example, from Tm3+:3H5 level to Ho3+:5I6 level (ET1) and from Tm3+:3F4 level to Ho3+:5I7 level (ET2, Tm3+:3F4 + Ho3+:5I8 → Tm3+:3H6 + Ho3+:5I7). Therefore, at 5I7 level, the Ho3+ ions from the indirect excitation of energy transfer ET2 together with those from the MPR process of Ho3+ ions at the 5I6 level, radiate to the ground state 5I8 and produce a fluorescence in the 2.0 µm band. In the case of Tm3+/Ho3+/Ce3+ tri-doping, the introduced Ce3+ ions can interact with Tm3+ ions through the cross relaxations of Tm3+:3F2,3 + Ce3+:2F5/2 → Tm3+:3H4 + Ce3+:2F7/2 (CR2) and Tm3+:3H5 + Ce3+:2F5/2 → Tm3+:3F4 + Ce3+:2F7/2 (CR3). These two CR2 and CR3 processes promote the Tm3+ ions return from 3F2,3 level to 3H4 level and from 3H5 level to 3F4 level respectively, which is conducive to the intensity increases of the 1.47, 1.83 and 2.0 µm three bands, but is negative for the 705 nm band. Further, with the addition of WO3 component, the increase in phonon energy of glass matrix facilitates the MPR processes from the high level to the low levels of Ho3+ and Tm3+ ions, especially the energy transfer process from Tm3+ to Ho3+ ions. For energy transfer ET2, the difference in energy between the Ho3+:5I7 level and Tm3+:3F4 level is approximately 900 cm-1, which nearly corresponds to the maximal phonon energy of glass matrix (∼916 cm-1). Thus, one phonon is enough to bridge the difference in this phonon-assisted energy transfer process. Just due to the increased MPR rate and energy transfer efficiency, the 1.83 and 2.0 µm band fluorescence intensities increase further with the introduction of WO3 component.

 figure: Fig. 6.

Fig. 6. Energy level diagram of Tm3+, Ho3+, Ce3+ transitions and interactions among them under LD excitation at 808 nm.

Download Full Size | PDF

To see the influence of Ce3+ ions and WO3 component on the luminescence properties of Ho3+ and Tm3+ ions more clearly, Fig. 7 displayed the fluorescence emission spectra of T-TH0 glass sample without Ce3+ ions and WO3 component, T-TH3 glass sample with 7.5 mol% WO3 component and without Ce3+ ions, T-THC0 glass sample with 0.1 mol% CeO2 and without WO3 component, and T-THC3 glass sample with 0.1 mol% CeO2 and 7.5 mol% WO3 component. In comparison to glass sample T-TH0, it displays that the introduction of Ce3+ ions or WO3 component both enhances the fluorescence intensities in the 1.47, 1.83 and 2.0 µm three bands but weakens the 705 nm band emission. Importantly, the simultaneous introduction of Ce3+ ions and WO3 component has a more obvious effect. For the broadband emission band of 1600–2200 nm, an increase of about 103% in the glass sample T-THC3 was observed. In addition, this broadband emission band has a full width at half maximum (FWHM) of ∼360 nm, which is better than that of most reports such as Tm3+/Ho3+ co-doped antimony-silicate glass (356 nm) [45], lanthanum aluminosilicate glass (343 nm) [46], silicate glass (350 nm) [47], silicate-germanate glass (231.5 nm) [48], Tm3+/Ho3+/Yb3+ tri-doped gallo-germanate glass (343 nm) [49] and Tm3+/Ho3+/Ce3+ tri-doped bismuth tellurite glass (216 nm) [50].

 figure: Fig. 7.

Fig. 7. Fluorescence spectra in the range of (a) 1200-2200 nm and (b) 620-740 nm in T-TH0, T-TH3, T-THC0 and T-THC3 glass samples under 808 nm LD excitation.

Download Full Size | PDF

The fluorescence transient behavior of 1.47, 1.83 and 2.0 µm three bands in samples with Tm3+/Ho3+ co-doping and Tm3+/Ho3+/Ce3+ tri-doping was examined. Figures 8 and 9 displayed the measured fluorescence decay curves of corresponding samples. All of the curves deviated from mono-exponential decay law, implying interactions like cross relaxation together with energy transfer exist among the doping rare earth ions. Decay curve is fitted by a double exponential function [51]:

$$I(t) = {A_1}\exp ( - \frac{t}{{{\tau _1}}}) + {A_2}\exp ( - \frac{t}{{{\tau _2}}}) + I({t_0})$$
in which $I({{t_0}} )$ represents the constant background such as the initial intensity and $I(t )$ denotes fluorescence intensity at t time. ${A_1}$ and ${A_2}$ represent the intensity coefficients corresponding to two channel decays with lifetime ${\tau _1}$ and ${\tau _2}$ respectively. Hence, the estimation of fluorescence lifetime is conducted with the formula below:
$${\tau _{m}} = \frac{{{A_1}\tau _2^2 + {A_2}\tau _2^2}}{{{A_1}{\tau _1} + {A_2}{\tau _2}}}$$

 figure: Fig. 8.

Fig. 8. Fluorescence decay curves of (a) Tm3+:3H43F4, (b) Tm3+:3F43H6 and (c) Ho3+:5I75I8 transitions in Tm3+/Ho3+ co-doped tellurite glass samples under 808 nm LD excitation, and the solid line is fitting result.

Download Full Size | PDF

 figure: Fig. 9.

Fig. 9. Fluorescence decay curves of (a) Tm3+:3H43F4, (b) Tm3+:3F43H6 and (c) Ho3+:5I75I8 transitions in Tm3+/Ho3+/Ce3+ tri-doped tellurite glass samples under 808 nm LD excitation, and the solid line is fitting result.

Download Full Size | PDF

It is seen that when 7.5 mol% WO3 component is introduced into samples with Tm3+/Ho3+ co-doping and Tm3+/Ho3+/Ce3+ tri-doping, the lifetimes of Tm3+:3H43F4,3F43H6 and Ho3+:5I75I8 radiative transitions are all increased, which are in good accordance with the fluorescence emission intensities observed. Among them, the lifetime of Tm3+:3H43F4 transition is slightly increased from 0.051 to 0.063 ms in the Tm3+/Ho3+ co-doping case, that of Tm3+:3F43H6 transition is increased from 1.04 to 1.21 ms, and that of Ho3+:5I75I8 transition is increased from 1.32 to 1.40 ms. While in the case of Tm3+/Ho3+/Ce3+ tri-doping, the lifetimes of above transitions are increased from 0.064 to 0.076 ms, 1.27 to 1.44 ms and 1.44 to 1.87 ms, respectively. As analyzed above, the addition of WO3 component brings a raise of vibrational phonon energy in glass network, which enhances the process of energy transfer ET2 as well as the MPR process from the high levels to the low-lying levels, and this is responsible for the increase of radiative transition lifetimes. However, when the amount of WO3 component in glass continues to increase, the vibrational phonon density at 916 cm-1 increases accordingly, and the MPR processes of these three NIR emitting levels will become stronger synchronously, which leads to the fluorescence and lifetime of these transitions decrease. Therefore, the amount of WO3 oxide with high phonon energy introduced into the tellurite glass must be appropriate to ensure that the fluorescence emission tends to the best state. In addition, it is pointed out that the lifetime of the three transitions in samples with Tm3+/Ho3+/Ce3+ tri-doping is larger in comparison to that in the case of Tm3+/Ho3+ co-doping, and this is attributed to the CR2 and CR3 processes between the Ce3+ and Tm3+ ions.

From the measured fluorescence lifetime (${\tau _{m}}$) and the calculated spontaneous radiative lifetime (${\tau _{\textrm{rad}}}$) according to the J-O theory, the quantum efficiency (${\eta _{\textrm{rad}}}$) of Tm3+:3F43H6 and Ho3+:5I75I8 transitions can be estimated by the following [52]:

$${\eta _{\textrm{rad}}} = \frac{{{\tau _{m}}}}{{{\tau _{\textrm{rad}}}}}$$

Table 5 lists the quantum efficiency of Tm3+:3F43H6 (1.83 µm) and Ho3+:5I75I8 (2.0 µm) transitions including the fluorescence lifetime and spontaneous radiative lifetime of T-TH0 and T-TH3 tellurite glass samples. With the introduction of WO3 component, the quantum efficiency of Tm3+:1.83 µm and Ho3+:2.0 µm two bands has been improved to different degrees, which again shows that the introduction of an appropriate amount of WO3 component has a beneficial effect on these fluorescence emissions. In addition, the quantum efficiency of Tm3+:3F43H6 transition in this work is higher than that of germanate glass (23.9%) [53], but lower than that of fluorotellurite glass (75.93%) [54]. While the quantum efficiency of Ho3+:5I75I8 transition is higher than that of silica glass (9.38%) [55], but lower than that of low silica aluminosilicate glass (51%) [56]. High quantum efficiency is mainly attributed to the low phonon energy of glass and is conducive to the fluorescence emission.

Tables Icon

Table 5. The spontaneous radiative lifetime (${\tau _{\textrm{rad}}}$), fluorescence lifetime (${\tau _{m}}$) and quantum efficiency (${\eta _{\textrm{rad}}}$)

3.4 Multiphonon relaxation and energy transfer

The introduction of an appropriate amount of WO3 can enhance the NIR fluorescence emissions, which is attributed to the increased MPR rates from the high levels of Tm3+ and Ho3+ ions to their fluorescence emitting levels together with the enhanced energy transfer from the Tm3+:3F4 level to the Ho3+:5I7 level. Hence, it is essential to examine the effect of the addition of WO3 on the MPR between the relevant levels and the energy transfer from Tm3+ to Ho3+ ions. The MPR rate from a high level to a low level can be estimated from an empirical formula [8,57]:

$${W_{\textrm{mpr}}} = B\exp [{ - (\Delta E - 2\hbar \omega )\alpha } ]$$
in which B and $\alpha$ both represent the characteristic parameters of glass host, which is unrelated to the electron level of rare earth ions from which the decay occurs. $\Delta E$ is the energy gap and $\hbar \omega$ is the host phonon energy. For tellurite glass, B and $\alpha$ are given as the values of 107.97 s-1 and 4.7 × 10−3 cm respectively [47]. According to the above formula, for example, the MPR rate of Tm3+:3H53F4 transition with energy gap ∼2240 cm-1 in tellurite glass sample T-TH0 without WO3 is about 3.19 × 106 s-1 and increases significantly to 1.37 × 107 s-1 in tellurite glass sample T-TH3 with WO3, while that of Tm3+:3F43H6 transition with energy gap ∼6000 cm-1 is basically unchanged in these two glass samples because this rate is nearly zero compared with the former. Therefore, with the increase of host phonon energy from 761 to 916 cm-1 after the introduction of WO3 component, the number of Tm3+ ions at 3F4 level increases greatly, as a result, the radiative transition of Tm3+:3F43H6 (1.85 µm) and the energy transfer from Tm3+:3F4 level to Ho3+:5I7 level (ET2) are enhanced, and the latter leads to the increase of Ho3+:5I75I8 radiative transition (2.0 µm).

The energy transfer among rare earth ions is primarily based on the dipole-dipole interaction, and the microscopic probability from the donor (D) to the acceptor (A) was defined by D.L. Dexter [58] as below:

$${W_{\textrm{DA}}} = \frac{{{C_{\textrm{DA}}}}}{{{R^6}}}$$
where R denotes the separation distance between the acceptor and donor ions, and ${C_{\textrm{DA}}}$ is a microscopic coefficient characterizing the energy transfer, as determined by [58]:
$${C_{\textrm{DA}}} = \frac{{R_\textrm{C}^6}}{{{\tau _\textrm{D}}}}$$
where ${\tau _\textrm{D}}$ is the donor lifetime at excitation level and ${R_\textrm{C}}$ denotes the critical radius for the interaction of donor with acceptor:
$$R_\textrm{C}^6 = \frac{{6\textrm{c}{\tau _\textrm{D}}}}{{{{(2\mathrm{\pi })}^4}{n^2}}}\frac{{g_{\textrm{low}}^\textrm{D}}}{{g_{\textrm{up}}^\textrm{D}}}\int {\sigma _{\textrm{em}}^\textrm{D}(\lambda )} \sigma _{\textrm{abs}}^\textrm{A}(\lambda )d\lambda$$
where $g_{\textrm{up}}^\textrm{D}$ and $g_{\textrm{low}}^\textrm{D}$ represent the donor ion degeneracies in upper and lower levels, $\sigma _{\textrm{em}}^\textrm{D}$ represents the emission cross section of donor ions while $\sigma _{\textrm{abs}}^\textrm{A}$ represents the absorption cross section of acceptor ions, and they can be determined by Beer-Lambert law [59] from the absorption spectrum and McCumber theory [60] respectively:
$${\sigma _{\textrm{abs}}}(\lambda )= \frac{{2.303OD(\lambda )}}{{N \cdot L}}$$
$${\sigma _{\textrm{em}}}(\lambda ) = {\sigma _{\textrm{abs}}}\frac{{{Z_\textrm{l}}}}{{{Z_\textrm{u}}}}\exp \left[ {\frac{{\textrm{hc}}}{{\textrm{k}T}}\left( {\frac{1}{{{\lambda_\textrm{p}}}} - \frac{1}{\lambda }} \right)} \right]$$
where N is the doping concentration, L represents the glass sample thickness, $OD(\lambda )$ denotes the optical density at wavelength ${\lambda _{}}$. ${Z_\textrm{u}}$ and ${Z_\textrm{l}}$ represent the partition functions of high and low energy levels respectively, ${\lambda _\textrm{p}}$ is the peak wavelength of absorption band, $\textrm{k}$ is the Boltzmann constant and $T$ is the room temperature.

In the case of existing energy difference between the donor and the acceptor levels, the energy transfer between them requires the assistance of host phonons, and the overlap integral in Eq. (7) was revised as the following [61]:

$$R_\textrm{C}^6 = \frac{{6\textrm{c}{\tau _\textrm{D}}}}{{{{(2\mathrm{\pi })}^4}{n^2}}}\frac{{g_{\textrm{low}}^\textrm{D}}}{{g_{\textrm{up}}^\textrm{D}}}\sum\limits_{N = 0}^\infty {\sum\limits_{m = 0}^N {\int {\sigma _{\textrm{em(}m\textrm{ - phonon)}}^\textrm{D}(\lambda )} \sigma _{\textrm{abs(}k\textrm{ - phonon)}}^\textrm{A}(\lambda )} } d\lambda$$
where $\sigma _{\textrm{em(}m\textrm{ - phonon)}}^\textrm{D}$ and $\sigma _{\textrm{abs(}k\textrm{ - phonon)}}^\textrm{A}$ represent the m-phonon emission and k-phonon absorption sidebands for donor and acceptor ions respectively, N denotes the number of phonon needed in the energy transfer (N = m + k), and they are determined by:
$$\sigma _{\textrm{em(}m\textrm{ - phonon)}}^\textrm{D}(\lambda ) = {e^{[ - (2\bar{n} + 1){s_0}]}}\frac{{s_0^m}}{{m!}}{(\bar{n} + 1)^m}\sigma _{\textrm{em}}^\textrm{D}(\lambda _m^\textrm{ + })\;$$
$$\;\sigma _{\textrm{abs(}k\textrm{ - phonon)}}^\textrm{A}(\lambda ) = {e^{[ - 2\bar{n}{s_0}]}}\frac{{s_0^k}}{{k!}}{(\bar{n})^k}\sigma _{\textrm{abs}}^\textrm{A}(\lambda _k^\textrm{ - })$$
where ${s_0}$ denotes the Huang-Rhys factor of rare earth ions (${s_0}$=0.31), $\overline n = {1 / {({{e^{\hbar {\omega_0}/\textrm{k}T}} - 1} )}}$ is the average occupancy of phonon, and $\hbar {\omega _0}$ is the maximum phonon energy of glass host. $\lambda _m^ +{=} {1 / {({{1 / {\lambda - m\hbar {\omega_0}}}} )}}$ and $\lambda _k^ -{=} {1 / {({{1 / {\lambda + k\hbar {\omega_0}}}} )}}$ denote the wavelength shifts of donor emission cross section caused by m-phonon emission and acceptor absorption cross section caused by k-phonon absorption, respectively. If one takes into account only the m-phonon emission and ignores k-phonon absorption for simplicity (N = m), then the micro probability and the energy transfer coefficient from the donor to the acceptor ions are expressed below [61]:
$$\;{W_{\textrm{DA}}} = \frac{{6\textrm{c}g_{\textrm{low}}^\textrm{D}}}{{{{(2\mathrm{\pi })}^4}{n^2}{R^6}g_{\textrm{up}}^\textrm{D}}}\sum\limits_{N = 0}^\infty {\int {\sigma _{\textrm{em(}N\textrm{ - phonon)}}^\textrm{D}({\lambda_N^ + } )\sigma _{\textrm{abs}}^\textrm{A}(\lambda )} \textrm{d}\lambda }$$
$${C_{\textrm{DA}}} = \frac{{6\textrm{c}g_{\textrm{low}}^\textrm{D}}}{{{{(2\mathrm{\pi })}^4}{n^2}g_{\textrm{up}}^\textrm{D}}}\sum\limits_{N = 0}^\infty {\int {\sigma _{\textrm{em(}N\textrm{ - phonon)}}^\textrm{D}({\lambda_N^ + } )\sigma _{\textrm{abs}}^\textrm{A}(\lambda )} \textrm{d}\lambda }$$

The energy transfer micro coefficients, the number of phonons involved and their contributions to the total probability calculated for the energy transfer from Tm3+:3F4 level to Ho3+:5I7 level (ET2) in the Tm3+/Ho3+ co-doping tellurite glass samples (T-TH0 and T-TH3) with and without WO3 component are summarized in Table 6. The absorption cross section of acceptor ions, and the emission cross section and emission sideband of donor ions required in the calculation are shown in Fig. 10. It is concluded from the table that the energy transfer ET2 is a non-resonant interaction process, which is mainly assisted by one phonon, and its contribution ratio in T-TH0 and T-TH3 glass sample has reached about 87% and 92% respectively. The increase of one phonon contribution ratio in glass sample after the introduction of WO3 component indicates that the increased phonon energy is more suitable for bridging the energy difference between the Tm3+:3F4 level and Ho3+:5I7 level, thus enhancing the energy transfer, which is conducive to the 2.0 µm band fluorescence of Ho3+ ions.

 figure: Fig. 10.

Fig. 10. Emission cross-section of Tm3+:3F43H6 transition and absorption cross-section of Ho3+:5I85I7 transition in (a) T-TH0 and (b) T-TH3 tellurite glass samples, and the dash line is one and two-phonon emission sidebands of Tm3+:3F43H6 transition.

Download Full Size | PDF

Tables Icon

Table 6. Microscopic parameters of energy transfer Tm3+:3F4→Ho3+:5I7 in T-TH0 and T-TH3 tellurite glass samples.

4. Conclusions

An enhanced broad 1600–2200 nm band luminescence with a FWHM of ∼360 nm was obtained in tellurite glasses tri-doped with Tm3+, Ho3+ and Ce3+ containing WO3 excited at 808 nm LD. The spectral overlapping of 1.83 µm band fluorescence of Tm3+ and 2.0 µm band fluorescence of Ho3+ resulted in this broadband luminescence. The tri-doping of Ce3+ with Tm3+ and Ho3+ accelerated the return of Tm3+ ions to the fluorescence level 3F4 through the cross relaxation among Ce3+ and Tm3+ ions which is beneficial to the 1.83 µm band, while the increase of glass phonon energy induced by the addition of WO3 component enhanced the MPR rates as well as the energy transfer from Tm3+:3F4 level to Ho3+:5I7 level which is beneficial to the 2.0 µm band, all these led to an improvement of broadband luminescence by about 103% in comparison to the tellurite glass only with Tm3+/Ho3+ co-doping. The present study proposed an experimental scheme to improve the broadband luminescence, which has significant implications for the further development of optoelectronic devices operated in ∼2.0 µm band.

Funding

National Natural Science Foundation of China (61875095, U22A2085).

Disclosures

The authors declare no conflicts of interest.

Data availability

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

References

1. P. P. Xiao and B. Wang, “Design of an erbium-doped Al2O3 optical waveguide amplifier with on-chip integrated laser pumping source,” Opt. Commun. 508, 127709 (2022). [CrossRef]  

2. F. Wang, L. L. Hu, W. B. Xu, M. Wang, S. Y. Feng, J. J. Ren, L. Zhang, D. P. Chen, N. Ollier, G. J. Gao, C. L. Yu, and S. K. Chen, “Manipulating refractive index, homogeneity and spectroscopy of Yb3+-doped silica-core glass towards high-power large mode area photonic crystal fiber lasers,” Opt. Express 25(21), 25960–25969 (2017). [CrossRef]  

3. K. Suresh, K. V. Krishnaiah, C. Basavapoornima, S. R. Depuru, and C. K. Jayasankar, “Enhancement of 1.8 mm emission in Er3+/Tm3+ co-doped tellurite glasses: Role of energy transfer and dual wavelength pumping schemes,” J. Alloys Compd. 827, 154038 (2020). [CrossRef]  

4. G. W. Tang, X. Wen, Q. Qian, T. T. Zhu, W. W. Liu, M. Sun, X. D. Chen, and Z. M. Yang, “Efficient 2.0 µm emission in Er3+/Ho3+ co-doped barium gallogermanate glasses under different excitations for mid-infrared laser,” J. Alloys Compd. 664, 19–24 (2016). [CrossRef]  

5. M. Kochanowicz, J. Zmojda, P. Miluski, T. Ragin, W. A. Pisarski, J. Pisarska, R. Jadach, M. Sitarz, and D. Dorosz, “Structural and luminescent properties of germanate glasses and double-clad optical fiber co-doped with Yb3+/Ho3+,” J. Alloys Compd. 727, 1221–1226 (2017). [CrossRef]  

6. J. J. Zhang, Y. Lu, M. Z. Cai, Y. Tian, F. F. Huang, and S. Q. Xu, “Highly efficient 2.84-µm emission in Ho3+/Yb3+ co-doped tellurite-germanate glass for mid-infrared laser materials,” IEEE Photonics Technol. Lett. 29(17), 1498–1501 (2017). [CrossRef]  

7. M. Kochanowicz, J. Zmojda, P. Miluski, A. Baranowska, M. Leich, A. Schwuchow, M. Jäger, M. Kuwik, J. Pisarska, W. A. Pisarski, and D. Dorosz, “Tm3+/Ho3+ co-doped germanate glass and double-clad optical fiber for broadband emission and lasing above 2 µm,” Opt. Mater. Express 9(3), 1450–1458 (2019). [CrossRef]  

8. J. J. Kang, Z. F. Mo, Z. Y. Huang, J. H. Yang, W. C. Ma, J. T. Liu, C. M. Xia, Z. Y. Hou, and G. Y. Zhou, “Fabrication and optical properties of Tm3+/Ho3+ co-doped lanthanum aluminosilicate photonic crystal fiber for 2 µm fiber lasers,” J. Non-Cryst. Solids 596, 121869 (2022). [CrossRef]  

9. S. K. Taherunnisa, K. S. Rudramamba, D. Vijaya Sri, N. Veeraiah, T. Srikanth, and M. Rami Reddy, “Optimized NIR and MIR emission properties of Tm3+/Ho3+ ions in lead sulfo phosphate glasses,” Infrared Phys. Technol. 113, 103617 (2021). [CrossRef]  

10. N. M. Ty, D. C. Zhou, J. B. Qiu, and H. K. Dan, “Broadband flat near/mid-infrared emissions of Tm3+–Ho3+ co-doped, and Tm3+–Ho3+–Yb3+ tri-doped zinc silicate glasses under 808 and 980 nm laser diode excitations,” Infrared Phys. Technol. 111, 103483 (2020). [CrossRef]  

11. A. N. Trukhin, “Luminescence of α-quartz crystal and silica glass under excitation of excimer lasers ArF (193 nm), KrF (248 nm),” J. Lumin. 188, 524–528 (2017). [CrossRef]  

12. Z. Wang, F. F. Huang, B. Li, Y. Li, and S. Xu, “Enhanced luminescence properties of Ho/ Yb ions regulated by the nanocrystalline environment and phonon energy in silicate glasses,” J. Lumin. 219, 116949 (2020). [CrossRef]  

13. J. T. James, J. K. Jose, M. Manjunatha, K. Suresh, and A. Madhu, “Structural, luminescence and NMR studies on Nd3+-doped sodium-calcium-borate glasses for lasing applications,” Ceram. Int. 46(17), 27099–27109 (2020). [CrossRef]  

14. M. Walas, P. Piotrowski, T. Lewandowski, A. Synak, M. Łapinski, W. Sadowski, and B. Koscielska, “Tailored white light emission in Eu3+/Dy3+ doped tellurite glass phosphors containing Al3+ ions,” Opt. Mater. 79, 289–295 (2018). [CrossRef]  

15. F. Ahmadi, R. Hussin, and S. K. Ghoshal, “Physical and structural properties of dysprosium ion doped phosphate glasses,” Optik 227, 166000 (2021). [CrossRef]  

16. W. H. Tang, Y. Tian, B. P. Li, Y. Y. Xu, Q. H. Liu, J. J. Zhang, and S. Q. Xu, “Effect of introduction of TiO2 and GeO2 oxides on thermal stability and 2 µm luminescence properties of tellurite glasses,” Ceram. Int. 45(13), 16411–16416 (2019). [CrossRef]  

17. N. Manikandan, A. Ryasnyanskiy, and J. Toulouse, “Thermal and optical properties of TeO2–ZnO–BaO glasses,” J. Non-Cryst. Solids 358(5), 947–951 (2012). [CrossRef]  

18. A. Bachvarova-Nedelcheva, R. Iordanova, S. Ganev, and Y. Dimitriev, “Glass formation and structural studies of glasses in the TeO2–ZnO–Bi2O3–Nb2O5 system,” J. Non-Cryst. Solids 503-504, 224–231 (2019). [CrossRef]  

19. Y. D. Jiao, Z. X. Jia, X. H. Guo, Z. P. Zhao, Y. Ohishi, W. P. Qin, and G. S. Qin, “Third-order cascaded Raman shift in all-solid fluorotellurite fiber pumped at 1550 nm,” Opt. Lett. 47(3), 690–693 (2022). [CrossRef]  

20. N. A. Wójcik, N. Tagiara, S. Ali, K. Górnicka, H. Segawa, T. Klimczuk, B. Jonson, D. Möncke, and E. Kamitsos, “Structure and magnetic properties of BeO-Fe2O3-Al2O3-TeO2 glass-ceramic composites,” J. Eur. Ceram. Soc. 41(10), 5214–5222 (2021). [CrossRef]  

21. S. L. Zhao, B. Y. Chen, L. Wen, and L. L. Hu, “Spectral properties and stability of Er3+-doped TeO2–WO3 glass,” Mater. Chem. Phys. 99(2-3), 210–213 (2006). [CrossRef]  

22. R. Siripuram, P. S. G. Rao, and S. Sripada, “Optical, structural and thermal studies of xWO3+(30-x) As2O3+70TeO2 (where x=10, 20, 30 mol%) glasses by UV–vis, Raman, IR and DSC Studies,” Optik 224, 165450 (2020). [CrossRef]  

23. R. Gao, A. Shi, Z. Cao, X. Chu, A. Wang, X. Lu, C. Yao, and X. Li, “Construction of 2D up-conversion calcium copper silicate nanosheet for efficient photocatalytic nitrogen fixation under full spectrum,” J. Alloys Compd. 910, 164869 (2022). [CrossRef]  

24. M. Kumar, Y. Ratnakaram, R. Vijayalakshmi, and T. R. Raman, “Probing into structural and spectroscopic properties of Dy3+ doped and Eu3+/Dy3+ co-doped bismuth phosphate (BiPO4) glass ceramics with different modifier fluorides,” J. Lumin. 245, 118786 (2022). [CrossRef]  

25. B. Shanmugavelu, V. V. Ravi Kanth Kumar, R. Kuladeep, and D. Narayana Rao, “Third order nonlinear optical properties of bismuth zinc borate glasses,” J. Appl. Phys. 114(24), 243103 (2013). [CrossRef]  

26. W. J. Zhang, Q. Y. Zhang, Q. J. Chen, Q. Qian, Z. M. Yang, J. R. Qiu, P. Huang, and Y. S. Wang, “Enhanced 2.0 µm emission and gain coefficient of transparent glass ceramic containing BaF2:Ho3+, Tm3+ nanocrystals,” Opt. Express 17(23), 20952–20958 (2009). [CrossRef]  

27. R. F. Falci, T. Guerineau, J. L. Delarosbil, and Y. Messaddeq, “Spectroscopic properties of gallium-rich germano-gallate glasses doped with Tm3+,” J. Lumin. 249, 119014 (2022). [CrossRef]  

28. G. W. Tang, Z. H. Liang, W. H. Huang, D. L. Yang, L. Tu, W. Lin, X. G. Song, D. D. Chen, Q. Qian, X. M. Wei, and Z. M. Yang, “Broadband high-gain Tm3+ /Ho3+ co-doped germanate glass multimaterial fiber for fiber lasers above 2 µm,” Opt. Express 30(18), 32693–32703 (2022). [CrossRef]  

29. X. E. Su, Y. X. Zhou, Y. R. Zhu, M. H. Zhou, J. Li, and H. R. Shao, “Energy transfer induced 2.0: µm luminescent enhancement in Ho3+/Yb3+/Ce3+ tri-doped tellurite glass,” J. Lumin. 203, 26–34 (2018). [CrossRef]  

30. T. Sambasiva Rao, K. S. Rudramamba, D. Vijayasri, and M. Rami Reddy, “Optical properties and energy transfer probing for white light emission in Ce3+/Tb3+/ Sm3+ tri-doped barium gallium borosilicate glasses,” Opt. Mater. 120, 111412 (2021). [CrossRef]  

31. B. R. Judd, “Optical absorption intensities of rare-earth ions,” Phys. Rev. 127(3), 750–761 (1962). [CrossRef]  

32. G. S. Ofelt, “Intensity of crystal spectra rare-earth ions,” J. Chem. Phys. 37(3), 511–520 (1962). [CrossRef]  

33. Y. A. Yamusa, Y. Musa, T. Muhammad, A. Sa’id, A. Isma’ila, A. U. Abubakar, and I. Bulus, “Spectroscopic characteristics of barium sulfoborophosphate glasses doped samarium ions: Evaluation of Judd-Ofelt intensity parameters and radiative properties,” Optik 267, 169710 (2022). [CrossRef]  

34. M. S. A. Mohd Saidi, S. K. Ghoshal, R. Arifin, M. K. Roslan, R. Muhammad, W. N. W. Shamsuri, M. Abdullah, and M. S. Shaharin, “Spectroscopic properties of Dy3+ doped tellurite glass with Ag/TiO2 nanoparticles inclusion: Judd-Ofelt analysis,” J. Alloys Compd. 754, 171–183 (2018). [CrossRef]  

35. Y. Xu, Y. Li, Y. Tian, Y. Hua, F. Huang, J. Zhang, and S. Xu, “Excellent luminescence performance of Ho3+ ions based on a new system of oxyhalide glass,” J. Non-Cryst. Solids 594, 121823 (2022). [CrossRef]  

36. C. D. Viswanath, P. Babu, I. Martín, V. Venkatramu, V. Lavín, and C. Jayasankar, “Near-infrared and upconversion luminescence of Tm3+ and Tm3+/Yb3+-doped oxyfluorosilicate glasses,” J. Non-Cryst. Solids 507, 1–10 (2019). [CrossRef]  

37. X. Y. Song, K. X. Han, D. C. Zhou, P. F. Xu, and P. Zhang, “Broadband ∼1.8 µm emission characteristics of Tm3+-doped bismuth germanate glass based on Ga2O3 modification,” J. Non-Cryst. Solids 557, 120575 (2021). [CrossRef]  

38. Y. Y. Liu, X. H. Wang, Y. Wang, Z. Y. You, J. F. Li, Z. J. Zhu, and C. Y. Tu, “Crystal growth and spectroscopic analysis of Tm, Ho singly-doped and Tm/Ho co-doped CaLaGa3O7 NIR laser crystals,” Opt. Mater. 75, 744–750 (2018). [CrossRef]  

39. K. Biswas, A. D. Sontakke, R. Sen, and K. Annapurna, “Enhanced 2 µm broad-band emission and NIR to visible frequency up-conversion from Ho3+/Yb3+ co-doped Bi2O3-GeO2-ZnO glasses,” Spectrochim. Acta, Part A 112, 301–308 (2013). [CrossRef]  

40. B. Zhou, T. Wei, M. Cai, Y. Tian, J. Zhou, D. Deng, S. Xu, and J. Zhang, “Analysis on energy transfer process of Ho3+ doped fluoroaluminate glass sensitized by Yb3+ for mid-infrared 2.85 µm emission,” J. Quant. Spectrosc. Radiat. Transfer 149, 41–50 (2014). [CrossRef]  

41. I. E. Kolesnikov, M. A. Kurochkin, A. A. Kalinichev, E. Y. Kolesnikov, and E. Lähderanta, “Optical temperature sensing in Tm3+/Yb3+-doped GeO2-PbO-PbF2 glass ceramics based on ratiometric and spectral line position approaches,” Sens. Actuators, A 284, 251–259 (2018). [CrossRef]  

42. J. J. Wang, Z. X. Jia, C. Z. Zhang, Y. Sun, Y. Ohishi, W. P. Qin, and G. S. Qin, “Thulium-doped fluorotellurite glass fibers for broadband S-band amplifiers,” Opt. Lett. 47(8), 1964–1967 (2022). [CrossRef]  

43. L. Tu, G. Tang, Q. Qian, and Z. Yang, “Controllable structural tailoring for enhanced∼ 2 µm emission in heavily Tm3+-doped germanate glasses,” Opt. Lett. 46(2), 310–313 (2021). [CrossRef]  

44. C. Z. Wang, Y. Tian, X. Gao, Q. Liu, F. Huang, B. Li, J. Zhang, and S. Xu, “Mid-infrared fluorescence properties, structure and energy transfer around 2 µm in Tm3+/Ho3+ co-doped tellurite glass,” J. Lumin. 194, 791–796 (2018). [CrossRef]  

45. J. Żmojda, D. Dorosz, and J. Dorosz, “2.1 µm emission of Tm3+/Ho3+-doped antimony-silicate glasses for active optical fiber,” Bull. Pol. Acad. Sci. Chem. 59(4), 381–387 (2011). [CrossRef]  

46. N. Wang, R. J. Cao, M. Z. Cai, L. L. Shen, Y. Tian, F. F. Huang, S. Q. Xu, and J. J. Zhang, “Ho3+/Tm3+ codoped lead silicate glass for 2 µm laser materials,” Opt. Laser Technol. 97, 364–369 (2017). [CrossRef]  

47. S. Meng, G. Zhao, J. Hou, Y. Liu, Y. Guo, Y. Fang, Y. Zhou, and J. Zou, “High performance of near-infrared emission for S-band amplifier from Tm3+-doped bismuth glass incorporated with Ag nanoparticles,” J. Lumin. 224, 117313 (2020). [CrossRef]  

48. X. Gao, Y. Tian, Q. Liu, B. Li, W. Tang, J. Zhang, and S. Xu, “Broadband 2 µm emission characteristics and energy transfer mechanism of Ho3+ doped silicate-germanate glass sensitized by Tm3+ ions,” Opt. Laser Technol. 111, 115–120 (2019). [CrossRef]  

49. M. Kochanowicz, J. Żmojda, P. Miluski, A. Baranowska, T. Ragin, J. Dorosz, M. Kuwik, W. A. Pisarski, J. Pisarska, M. Leśniak, and D. Dorosz, “2 µm emission in gallo-germanate glasses and glass fibers co-doped with Yb3+/Ho3+ and Yb3+/Tm3+/ Ho3+,” J. Lumin. 211, 341–346 (2019). [CrossRef]  

50. C. Yun, C. M. Zhang, X. L. Miao, Z. W. Li, S. Y. Lai, and T. Sang, “Ultra-broadband 4.1 µm mid-infrared emission of Ho3+ realized by the introduction of Tm3+ and Ce3+,” J. Lumin. 239, 118368 (2021). [CrossRef]  

51. M. F. H. Schuurmans and J. M. F. van Dijk, “On radiative and non-radiative decay times in the weak coupling limit,” Phys. B+C 123(2), 131–155 (1984). [CrossRef]  

52. T. F. Xue, L. Y. Zhang, J. J. Hu, M. S. Liao, and L. L. Hu, “Thermal and spectroscopic properties of Nd3+-doped novel fluorogallate glass,” Opt. Mater. 47, 24–29 (2015). [CrossRef]  

53. C. Z. Liu, Y. X. Zhuang, J. J. Han, J. Ruan, X. J. Zhao, and T. Kavetskyy, “Enhanced ∼1.8 µm photoluminescence under blue light excitation in Tm-Bi co-doped germanate glass and its temperature dependence,” J. Non-Cryst. Solids 525, 119645 (2019). [CrossRef]  

54. D. C. Tang, Q. S. Liu, X. J. Liu, X. Wang, X. Y. Yang, Y. Y. Liu, T. Ying, Y. Renguang, X. H. Zhang, and S. Q. Xu, “High quantum efficiency of 1.8 µm luminescence in Tm3+ fluoride tellurite glass,” Infrared Phys. Technol. 123, 104055 (2022). [CrossRef]  

55. X. Wang, L. L. Hu, W. B. Xu, S. K. Wang, L. Zhang, C. L. Yu, and D. P. Chen, “Spectroscopic properties of Ho3+ and Al3+ co-doped silica glass for 2-µm laser materials,” J. Lumin. 166, 276–281 (2015). [CrossRef]  

56. G. S. N. Eliel, N. G. C. Astrath, T. O. Sales, J. H. Rohling, L. C. Malacarne, M. L. Baesso, L. A. O. Nunes, A. Novatski, T. Catunda, and C. Jacinto, “Spectroscopic investigation and heat generation of Tm3+/Ho3+-codoped aluminosilicate glasses emitting at 2.0 µm,” J. Opt. Soc. Am. B 38(11), 3222–3229 (2021). [CrossRef]  

57. S. L. Zhao, F. Zheng, S. Q. Xu, H. P. Wang, and B. L. Wang, “Er3+-doped phosphor-tellurite glass for broadband short-length Er3+-doped fiber amplifier,” Chin. Opt. Lett. 6(4), 276–278 (2008). [CrossRef]  

58. D. L. Dexter, “A theory of sensitized luminescence in solids,” J. Chem. Phys. 21(5), 836–850 (1953). [CrossRef]  

59. X. Wang, H. Lin, D. Yang, L. Lin, and E. Y. Pun, “Optical transitions and up-conversion fluorescence in Ho3+/Yb3+ doped bismuth tellurite glasses,” J. Appl. Phys. 101(11), 113535 (2007). [CrossRef]  

60. X. Wang, S. J. Fan, K. F. Li, L. Zhang, S. K. Wang, and L. L. Hu, “Compositional dependence of the 1.8 µm emission properties of Tm3+ ions in silicate glass,” J. Appl. Phys. 112(10), 103521 (2012). [CrossRef]  

61. L. V. G. Tarelho, L. Gomes, and I. M. Ranieri, “Determination of microscopic parameters for nonresonant energy-transfer processes in rare-earth-doped crystals,” Phys. Rev. B 56(22), 14344–14351 (1997). [CrossRef]  

Data availability

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

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1.
Fig. 1. XRD patterns of the synthesized tellurite glass sample, and inset is the DSC curves of T-TH0 and T-TH3 tellurite glasses.
Fig. 2.
Fig. 2. Raman spectra of tellurite glass samples with and without WO3 component.
Fig. 3.
Fig. 3. Absorption spectra of tellurite glass samples doped with Tm3+, Ho3+ and Ce3+.
Fig. 4.
Fig. 4. Fluorescence spectra in the range of (a) 1200-2200 nm and (b) 620-740 nm in Tm3+/Ho3+ co-doping tellurite glass samples with different WO3 amounts under 808 nm LD excitation.
Fig. 5.
Fig. 5. Fluorescence spectra in the range of (a) 1200-2200 nm and (b) 620-740 nm in Tm3+/Ho3+/Ce3+ tri-doping tellurite glass samples with different WO3 amounts under 808 nm LD excitation.
Fig. 6.
Fig. 6. Energy level diagram of Tm3+, Ho3+, Ce3+ transitions and interactions among them under LD excitation at 808 nm.
Fig. 7.
Fig. 7. Fluorescence spectra in the range of (a) 1200-2200 nm and (b) 620-740 nm in T-TH0, T-TH3, T-THC0 and T-THC3 glass samples under 808 nm LD excitation.
Fig. 8.
Fig. 8. Fluorescence decay curves of (a) Tm3+:3H43F4, (b) Tm3+:3F43H6 and (c) Ho3+:5I75I8 transitions in Tm3+/Ho3+ co-doped tellurite glass samples under 808 nm LD excitation, and the solid line is fitting result.
Fig. 9.
Fig. 9. Fluorescence decay curves of (a) Tm3+:3H43F4, (b) Tm3+:3F43H6 and (c) Ho3+:5I75I8 transitions in Tm3+/Ho3+/Ce3+ tri-doped tellurite glass samples under 808 nm LD excitation, and the solid line is fitting result.
Fig. 10.
Fig. 10. Emission cross-section of Tm3+:3F43H6 transition and absorption cross-section of Ho3+:5I85I7 transition in (a) T-TH0 and (b) T-TH3 tellurite glass samples, and the dash line is one and two-phonon emission sidebands of Tm3+:3F43H6 transition.

Tables (6)

Tables Icon

Table 1. The compositions (mol %) of tellurite glass and corresponding name

Tables Icon

Table 2. Density ( ρ ), refractive index ( n ), doping concentrations of Tm3+ ( N Tm ), Ho3+ ( N Ho ) and Ce3+ ( N Ce ), and photo of the prepared glass sample.

Tables Icon

Table 3. The electric- and magnetic-dipole radiative transition probabilities ( A ed , A md ), branching ratio ( β ) and lifetime ( τ rad ) of partial energy levels in Tm3+ doped glass sample.

Tables Icon

Table 4. The electric- and magnetic-dipole radiative transition probabilities ( A ed , A md ), branching ratio ( β ) and lifetime ( τ rad ) of partial energy levels in Ho3+ doped glass sample.

Tables Icon

Table 5. The spontaneous radiative lifetime ( τ rad ), fluorescence lifetime ( τ m ) and quantum efficiency ( η rad )

Tables Icon

Table 6. Microscopic parameters of energy transfer Tm3+:3F4→Ho3+:5I7 in T-TH0 and T-TH3 tellurite glass samples.

Equations (14)

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

I ( t ) = A 1 exp ( t τ 1 ) + A 2 exp ( t τ 2 ) + I ( t 0 )
τ m = A 1 τ 2 2 + A 2 τ 2 2 A 1 τ 1 + A 2 τ 2
η rad = τ m τ rad
W mpr = B exp [ ( Δ E 2 ω ) α ]
W DA = C DA R 6
C DA = R C 6 τ D
R C 6 = 6 c τ D ( 2 π ) 4 n 2 g low D g up D σ em D ( λ ) σ abs A ( λ ) d λ
σ abs ( λ ) = 2.303 O D ( λ ) N L
σ em ( λ ) = σ abs Z l Z u exp [ hc k T ( 1 λ p 1 λ ) ]
R C 6 = 6 c τ D ( 2 π ) 4 n 2 g low D g up D N = 0 m = 0 N σ em( m  - phonon) D ( λ ) σ abs( k  - phonon) A ( λ ) d λ
σ em( m  - phonon) D ( λ ) = e [ ( 2 n ¯ + 1 ) s 0 ] s 0 m m ! ( n ¯ + 1 ) m σ em D ( λ m  +  )
σ abs( k  - phonon) A ( λ ) = e [ 2 n ¯ s 0 ] s 0 k k ! ( n ¯ ) k σ abs A ( λ k  -  )
W DA = 6 c g low D ( 2 π ) 4 n 2 R 6 g up D N = 0 σ em( N  - phonon) D ( λ N + ) σ abs A ( λ ) d λ
C DA = 6 c g low D ( 2 π ) 4 n 2 g up D N = 0 σ em( N  - phonon) D ( λ N + ) σ abs A ( λ ) d λ
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.