Next Article in Journal
Prediction of Bearing Capacity of the Square Concrete-Filled Steel Tube Columns: An Application of Metaheuristic-Based Neural Network Models
Previous Article in Journal
Manufacturing of a Magnetic Composite Flexible Filament and Optimization of a 3D Printed Wideband Electromagnetic Multilayer Absorber in X-Ku Frequency Bands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Yb3+ on the Structural and Visible to Near-Infrared Wavelength Photoluminescence Properties in Sm3+-Yb3+-Codoped Barium Fluorotellurite Glasses

School of Chemical and Process Engineering, University of Leeds, Clarendon Road, Leeds LS2 9JT, UK
*
Author to whom correspondence should be addressed.
Materials 2022, 15(9), 3314; https://doi.org/10.3390/ma15093314
Submission received: 31 March 2022 / Revised: 26 April 2022 / Accepted: 30 April 2022 / Published: 5 May 2022
(This article belongs to the Topic Optical and Optoelectronic Materials and Applications)

Abstract

:
We report on the Sm3+ and Sm3+:Yb3+-doped barium fluorotellurite glasses prepared using the conventional melting and quenching method. The spectroscopic characterisations were investigated with Raman and FTIR to evaluate the glasses’ structural and hydroxyl (-OH) content. The Raman analysis revealed a structural modification in the glass network upon adding and increasing the Yb3+ concentration from a TeO3 trigonal pyramid to a TeO4 trigonal bi-pyramid polyhedral. At the same time, the FTIR measurements showed the existence of -OH groups in the glass. Thus, under the current experimental conditions and nominal composition, the -OH group contents are too large to enable an effective removal. The near-infrared region of the absorption spectra is employed to determine the nephelauxetic ratio and bonding parameters. The average nephelauxetic ratio decreases, and the bonding parameter increases with the increasing Yb3+ content in the glasses. A room temperature visible and near-infrared photoluminescence ranging from 500 to 1500 nm in wavelength and decay properties were investigated for glasses doped with Sm3+ and Sm3+-Yb3+ by exciting them with 450 and 980 nm laser sources. Exciting the Sm3+- and Sm3+-Yb3+-doped glasses by 450 nm excitation reveals a new series of photoluminescence emissions at 1200, 1293, and 1418 nm, corresponding to the 6F11/2 state to the 6HJ (J = 7/2, 9/2, 11/2) transitions. Under the 976 nm laser excitation, a broad photoluminescence emission from 980 to 1200 nm was detected. A decay lifetime decreased from ~244 to ~43 μs by increasing the Yb3+ content, ascribing to concentration quenching and the OH content.

1. Introduction

During the last decades, various rare-earth (RE3+) ions doped in different host-glass matrices have been investigated for applications such as laser sources for optical telecommunications, optical amplifiers, multicolour displays, and medical diagnostics (sensing and imaging in the anatomical environment) [1,2,3,4]. Tellurium oxide (TeO2)-based glasses have numerous advantages over the borate, phosphate, and silicate host glasses [1,2,5]. This is due to their unique high linear and nonlinear refractive indices; outstanding characteristics, including high density, thermal stability, crystallisation, chemical durability, and low melting temperature; and high optical transparency in the visible to IR region up to 0.35–5 μm [5]. In addition, TeO2-based glasses have a high stimulated emission cross-section, a low-phonon-energy range from 560 to 750 cm−1 depending on the lattice modifier, and a high RE3+-ion concentration solubility [6,7,8,9]. However, a pure TeO2 has a low glass-forming ability due to its high quenching rate [10,11,12,13]. Consequently, this requires lattice modifiers such as alkali oxides, alkaline earth oxides, and transition metal oxides and halides to vitrify with a moderate quenching rate [13]. These lattice modifiers enable the TeO4 and TeO3 structural units to break down and decrease the average Te-O coordination [14], thus enhancing the glass-forming ability. For instance, adding a barium fluoride (BaF2) modifier to a TeO2-based glass network tends to decrease the phonon energy remarkably as compared to the phonon energy of the TeO2-based (Te-O~750 cm−1) [15,16]. Thus, reducing the nonradiative and multiphonon relaxation mechanism promotes optical transitions in RE3+ ions.
RE3+ ions such as trivalent samarium (Sm3+) are among the most studied active ions with low- and high-phonon-energy host materials owing to their adjacent closely lying energy-level structures, the enormous energy difference between the metastable level 4G5/2, and lower-lying levels such as 6FJ (J = 1/2, 3/2, 5/2,7/2, 9/2, and 11/2) and 6HJ (J = 5/2,7/2, 9/2, and 11/2) [16,17]. In addition, Sm3+ has more NIR absorption and photoluminescence emission transitions than erbium (Er3+), neodymium (Nd3+), and ytterbium (Yb3+). Hence, exploiting absorption and photoluminescence emission properties within the 700–1300 nm optical window [18] is attractive for monitoring soft tissue molecular signatures because of the low absorption and scattering in this spectral range. In addition, it has a vast potential for developing NIR lasers, optical amplifiers, temperature sensors, and biomedical devices.
Recently, numerous host materials doped with a Sm3+ ion have been explored because of their interesting optical properties and broad photoluminescence band from 4G5/2 to 6HJ and 6FJ transitions [19,20,21,22,23,24,25,26,27,28,29]. Here are some current studies of Sm3+-doped glasses with the focussed-on fluorescence emission properties from the orange to red wavelength region. Mawlud et al. [5] had reported optical spectroscopic properties on various concentrations of Sm3+ ion-doped TeO2-Na2O glasses. They demonstrated an increasing photoluminescence emission intensity under a 404 nm excitation source at the following transitions, 4G5/26H5/2 (561 nm), 4G5/26H7/2 (598 nm), 4G5/26H9/2 (645), and 4G5/26H11/2 (704 nm), upon increasing the Sm3+ ion concentration. Similarly, Hussain et al. [23] prepared various Sm3+ ion-doped phosphate glasses to analyse their optical absorption, radiative transition probability, and photoluminescence properties. The four main visible photoluminescence emission transitions from 4G5/2 to 6HJ were observed under an excitation of 400 nm. Even though the photoluminescence properties of Sm3+-doped glasses have been studied significantly between the 400 and 700 nm spectral region, there is limited research on the NIR transitions ranging from 800 to 1500 nm for comparative studies. Recently, Herrara et al. [24] investigated the structural and photoluminescence properties of Sm3+-doped heavy-metal oxide glass comprised of B2O3-PbO-Bi2O3-GeO2. The photoluminescence spectral of the glass revealed strong emission peaks of the Sm3+ ion at 563, 601, 648, and 711 nm, respectively. Additionally, NIR photoluminescence bands centred at 912, 954, 1036, and 1182 nm were reported, which matched with transitions from 4G5/26FJ (J = 3/2, 5/2, 7/2, 9/2). Likewise, visible and NIR photoluminescence emission peaks from tellurite zinc oxide glass doped with Sm3+ ions under an excitation wavelength of 403 nm had also been reported elsewhere [25].
The current approach to increasing the narrow absorption cross-section of the Sm3+ ions energy level in the NIR region from 850 to 1000 nm is codoping with Yb3+ to develop a long-wavelength broadband NIR (850–1150 nm) glass with enhanced optical properties through the resonant energy transfer process. Yb3+ is a well-known efficient sensitiser and activator with a broad absorption cross-section ranging from 950 to 1060 nm in wavelength, corresponding to the Yb3+: 6F7/26F5/2 transition. Therefore, codoping the Sm3+: Yb3+ with glass could broaden the absorption cross-section from 800 to 1150 nm by overlapping the Sm3+ (4G5/26F3/2, 4G5/26F5/2, 4G5/26F7/2) and Yb3+ (6F7/26F5/2) transitions. This approach would increase the pump efficiency to enable the realisation of population inversion for optical amplification via energy transfer processes between Sm3+ and Yb3+, curtailing resonant nonradiative transfer or nonresonant phonon-assisted nonradiative transfer [26,27,28].
In this article, we have fabricated Sm3+- and Yb3+:Sm3+-doped barium fluorotellurite zinc glasses by varying the Yb3+ concentrations. We performed a comprehensive analysis to investigate the influence of the Yb3+ content on the barium fluorotellurite zinc glasses’ structure and -OH content via Raman spectroscopy and Fourier transform infrared (FTIR) transmittance and absorption spectra. The Yb3+ concentration effect on the glass refractive index and nephelauxetic ratios are also determined. Furthermore, the visible and NIR photoluminescence emission and decay lifetimes are examined to understand the energy transfer process between the Yb3+ and Sm3+ ions by increasing the Yb3+ concentration under 450 and 980 nm pumping schemes.

2. Materials and Method

Barium fluorotellurite glasses doped/codoped with Sm3+ and Sm3+: Yb3+ ions with molar composition of (97.5 − x)TeO2-10ZnO-10BaF2-0.5Sm2O3-(x = 0–1.0Yb2O3) were fabricated using conventional melting and quenching technique. The starting materials comprise high-purity TeO2 (≥99.99%), ZnO (99.99%), BaF2 (99.99%), Yb2O3 (99.99%), and Sm2O3 (99.99%), which were purchased from Alfa Aesar. A 30 g batch of Sm3+- and Sm3+: Yb3+-doped barium fluorotellurite glasses were synthesised by weighing the appropriate amount of each powder raw material. A homogenous mixture of the powder samples was obtained using a mortar and a pestle to mix thoroughly for about 20 min. The mixture was transferred into a gold crucible and melted at 750–800 °C for about 3 h by employing an electrical furnace (Elite Thermal Systems Limited, Market Harborough, UK). During the glass melting, the furnace atmosphere was maintained under a dry oxygen atmosphere by purging a high-purity oxygen gas to control moisture and OH-ion ingress in the glass melt. The homogeneous molten glass was poured into a preheated brass mould at 300 °C for 3 h and then annealed at 300 °C for 4 h to remove thermal stress.
Finally, the annealing furnace was allowed to cool down to room temperature overnight and polished for optical measurements. A prism coupler (Metricon model 2010/M) and Thermo Pycnomatic ATC Helium Pycnometer were also utilised to measure the glass’s linear refractive index and density. The molar composition, density, and refractive index of the achieved glasses are depicted in Table 1. Raman spectroscopy measurements were performed utilising a Renishaw via Raman spectrometer with a green Ar+ laser (λ = 514.5 nm) excitation source (Renishaw, Wotton-under-Edge, UK) to determine the molecular structure of as-prepared glasses within the spectral range of 100–1000 cm−1. Similar, infrared absorbance spectroscopy was conducted to investigate the effect of the single-doped local structures compared to codoped glasses. The infrared absorbance spectra were collected in a Vertex 70 FITR spectrometer (Brucker, Coventry, UK) over 4 cm−1 resolution between 12,000 and 400 cm−1. The visible and NIR photoluminescence properties were investigated at room temperature with an FS920 spectrometer (Edinburgh Instruments, Livingston, UK) by exciting with 450 and 980 nm laser sources in the wavelength range of 500–750 and 850–1500 nm. The visible and NIR photoluminescence emission measurements were performed with a 495 nm short-pass filter cut-off wavelength.

3. Results and Discussion

3.1. Physical and Structural Properties—Density, Refractive Index, Raman and FTIR Spectroscopies

Table 1 shows the measured densities and the refractive index as a function of the Y3+ ion concentration. The average density values were determined using the Thermo Pycnomatic ATC Helium Pycnometer instrument range from 5.632 to 5.618 g/cm3. The refractive index increases slightly as the content of the Yb3+ increases, which could be attributed to its dependence on the polarisability of the ions or electron density [29] in the glass
Raman spectroscopy is a well-known technique for determining the host materials’ vibrational, structural, and functional groups and the phonon energy for the optical applications. Hence, the Raman spectra of the SBT glass series with different Yb3+ concentrations were monitored within a wavenumber spectral range of 0–1000 cm−1, as depicted in Figure 1. The Boson peak of the tellurite glass due to the disordered structure of the amorphous state occurring between 60 and 100 cm−1 is observed in these glass samples.
It can be noticed that the Raman spectra of all the glasses prepared exhibit a broad band between 0 and 200 cm−1 and a strong peak centred at ~65 cm−1, which confirms the presence of the Boson peak in the SBT series’ glassy structure. This apparent characteristic peak can be attributed to acoustic phonon and rotational molecular modes [29]. The Raman spectra of all the glass samples prepared were characterised by deconvoluting three prominent broad bands ranging from 220 to 370, 370 to 546, and 546 to 900 cm−1, as shown in Figure 1a–d, using origin software and fitted with Gaussian sub-band curves. The weak Raman intensity observed at 295 cm−1 can be attributed to the Ag and Fg Raman active modes of Sm-O [30,31]. The vibration band that peaked at 450 cm−1 is assigned to the bending and stretching vibrations of the Te-O-Te or O-Te-O linkage bonds, increasing the amplitude by increasing the Yb3+ content. On the other hand, enhancing this prominent peak could be attributed to the formation of Te-O-Ba/Yb sharing a vertex with TeO4, TeO3+δ, and TeO3 [32,33,34]. In addition, the vibration bands centred at 664 and 734 cm−1 show a vigorous Raman intensity and are ascribed to the structural fragment of the TeO4 trigonal bi-pyramid polyhedral (tbps), TeO3 trigonal pyramid (tp), and intermediate TeO3+δ polyhedron units, respectively [2,31,33,34]. The TeO4 (tbps) vibration peak is accredited to the asymmetric stretching vibrations of the Te-O-Te bridges which connect the TeO3 tbps at the vertices, whereas the TeO3 (tp) vibration band represents symmetric and asymmetric stretching modes due to the formation of nonbridge oxygen. It can be noticed that the relative intensity of a Sm3+-doped barium fluorotellurite (sample SBT1) Raman peak centred at 734 cm−1 (TeO3) is more intense as compared to the vibration band that occurred at 664 cm−1 (TeO4). The vibrational band centred at 781 cm−1 is ascribed to the stretching mode of the nonbridging oxygen TeO3+1 unit. Nevertheless, upon doping and increasing the Yb3+ concentration, the amplitude of the vibration band at 664 cm−1 progressively increases and predominates over the 734 cm−1 peak, as shown in Figure 2a. This demonstrates that the TeO3 (tps) unit has undergone a structural transformation by reducing the network connectivity to the TeO4 unit in the glass network. Figure 2b illustrates the intensity ratio of the bridge oxygen to nonbridge oxygen atoms ( I 664 I 735 ) as a function of the Yb2O3 concentration (mol%). It was observed that an increase in the Yb2O3 concentration shows a slight enhancement in the TeO4 (tbps) vibration band amplitude without any shift in the wavenumber. According to Maaoui et al. [35], such behaviour could be due to F (149 pm) substituting O2− (152 pm) because of the similarities in the ionic radius. For instance, the network form can be broken down by replacing two F- with one oxygen atom or interchanging F- with Ba in the TeO3 (tps) to become Te(O, F)3 someplace within the glass network.
Figure 3a illustrates a mid-infrared transmittance spectra of samples SBT1, SBT2, SBT3, and SBT4 with corresponding thicknesses of 4.45, 4.58, 4.56, and 4.48 mm recorded using a Brucker Vertex 70 FTIR spectrometer. Sample SBT1 at 4.45 mm thick exhibits a high average transmittance of about 52% within the spectral range of 2.5 to 3.0 μm and longer mid-infrared cut-off edges near 6.25 μm. The transmission of sample SBT1 revealed two broad bands dipping around 3.3 μm (3000 cm−1) and 4.38 μm (2304 cm−1), which are assigned to the -OH groups connected in the glass network [36,37,38]. The absorption band that occurred at 3000 cm−1 is attributed to the free -OH groups coupled to cations in the glass network, with a typical example being tellurium ion (Te-OH) and weakly bonded with hydrogen (H-OH) [36]. The absorption peak centred at 2304 cm−1 is due to the -OH absorption strongly connected with hydrogen (H-OH) and multiphonon absorption. However, no noticeable changes are observed upon doping with various concentrations of Yb3+ with sample STBT1; nevertheless, transmission decreases with decreasing Yb3+ content in the glass, as shown in Figure 3a.
The absorption coefficient of -OH, α OH ( cm 1 ) in the fluorotellurite glass was estimated by employing the Beer–Lambert law equation, which is expressed as [36]:
α OH = l n ( T ( % ) 100 ) L = A L
Meanwhile, the corresponding concentration N OH ( ions / cm 3 ) is obtained using Equation (2) [37]
N OH = N A V O ε α OH
where T(%) is the maximum transmittance in percentage at 3000 and 2304 cm−1, L is the glass thickness (cm), A is the absorbance, ε is the -OH group molar absorptivity in the tellurite glass (4.91 × 104 cm2mol−1) [38], and NAVO represents Avogadro’s constant (6.02 × 1023 mol−1).
The absorption coefficient spectra were determined using the transmittance spectra and Equation (1), as depicted in Figure 3b. The maximum absorption peaks centred at 3.30 and 4.38 μm in Figure 3b were employed to estimate the -OH concentration in all the glasses prepared, which is summarised in Table 2. It is observed that the α OH (cm−1) and N OH ( ions / cm 3 ) increase with increasing Yb3+ content in the glass series and these values correlate with oxyfluoride tellurite glasses reported elsewhere, respectively [35]. Therefore, high hydroxide absorption in the glasses can be reduced by either increasing the BaF2 content or substituting a portion of the ZnO with ZnF2.

3.2. Absorption Spectra, Bonding between Sm3+ and Surrounding Oxygen Atoms

Figure 4a represents the absorption coefficient spectra of the SBT glass series measured at room temperature using a FITIR spectrometer in the wavenumber region of 4000 to 12,000 cm−1. The absorption bands of the Yb3+ and various Sm3+ transitions are labelled in Figure 4a,b. About seven of the 4f5-4f5 electronic transitions of the Sm3+ from the 5H5/2 ground state to different excited states are observed in the hypersensitive frequency, as shown in Figure 4b. The Sm3+-hypersensitive absorption bands occur as a result of the transition from the 5H5/2 ground state to excited states such as 6F11/2, 6F9/2, 6F7/2, 6F5/2, 6F3/2, 6H15/2, and 6F1/2, which correspond to wavenumbers of 10,333 cm−1 (968 nm), 9238 cm−1 (1082 nm), 811 cm−1 (1233 nm), 7244 cm−1 (1381 nm), 6731 cm−1 (1486 nm), 6513 cm−1 (1536 nm), and 6292 cm−1 (1589 nm), respectively. In addition, the intense and broad absorption band in the wavenumber range of 11,200–9601 cm−1 are attributed to the overlap transitions between Sm3+ (5H5/26F11/2) and Yb3+ (2F7/22F5/2). The NIR absorption transitions are sensitive to an environmental change around the ions, thus leading to broad absorption ranging from 7542 to 9560 cm−1 as the Yb3+ content increases to 1 mol%. The broad absorption band in the SBT4 glass could be accredited to the presence of a -OH second overtone in the glass. In addition, most of the NIR transitions from 5H5/26F9/2, 6F7/2, 6F5/2, 6F3/2, and 6F1/2 in Figure 4 obey the selection rule with J 6 inducing electric-dipole interactions, thus exhibiting sharp and more intense absorption bands.
Furthermore, the absorption spectra of the as-prepared glasses were employed to investigate the significant nature of the bonding between Sm3+ and their surrounding oxygen atoms in the host glasses. The slight change in the absorption peak position at the NIR transitions of the glasses is attributed to the variation in the surrounding environment of Sm3+ and the structure complexity. These were evaluated using the nephelauxetic ratio (β) and bonding parameters (δ). The nephelauxetic effect arises from the expansion of partially filled 4f-4f-shells owing to different glass compositions with a charge transfer from the ligand to the core of the Sm3+ and Yb3+ [39,40]. The nephelauxetic ratio, ( β ) , is expressed as the ratio of a particular RE3+ ion-absorption transition in wavenumber (cm−1), ( v g l ) , in the host glass, to the corresponding transition of the aquo-ion in wavenumber (cm−1), ( v a ) [5,41,42]:
β = v g l v a
The average value of the nephelauxetic ratio, β ¯ = β N , (N is the total amount of observed absorption in the glass matrix) was utilised to determine the bonding parameter, ( σ ), from the following equation [17,41,42]:
σ = ( 1 β ¯ β ¯ ) × 100
The σ can be either positive or negative, which signifies the covalent or ionic nature of the RE3+-ligand bond depending on the surrounding environment. All absorption coefficient spectra of the glass samples were fitted with multiple Gaussian function to determine the peak positions of each NIR transition of the Sm3+ ions. This provides an accurately centred wavenumber for each transition and enables accuracy in assessing the nephelauxetic ratio (β) and bonding parameters. The nephelauxetic effect is caused by the shortening of the metal–ligand distance bond formation due to the expansion of the electron cloud, which leads to a decrease in the coordination number. Table 3 shows β and σ values obtained from Equations (3) and (4) for the SBT glass series. The negative values of σ confirm the ionic character of the Sm3+- and Sm3+:Yb3+-doped and -codoped barium fluorotellurite glasses. The ionicity, σ , values increased with the increasing Yb3+ content leading to a structural modification. Such structural alterations have been observed in various glasses, as shown in Figure 2, leading to an increase in the TeO4 bridge oxygen peak, which accounts for increasing bonding parameters with increasing Yb3+.

3.3. Photoluminescence Spectra and Energy Transfer Process

The photoluminescence emission spectra of various Yb3+ concentrations in the Sm3+-doped barium fluorotellurite glass were initially investigated in the wavelength range 500–1600 nm. Under the 450 nm excitation, the exciting radiation undergoes a ground state absorption from a 6H5/2 state to a higher-lying 4G7/2 state with stark levels of Sm3+ (acting as a sensitizer). The populated 4G9/2 level dissipates with no emission light via multiphonon relaxation and a nonradiative process to lower levels such as 4G9/24G7/2, 4F3/2, and 4G5/2 shown in Figure 5. The emitted light undergoes radiative decay from the 4G5/2 energy level to various lower-lying states. On the other hand, the energy absorbed by the Sm3+ can also be transferred to an acceptor, Yb3+ (4G9/22F5/2 or 4F7/22F5/2), through energy transfer (ET) and cross-relaxation (CR) processes as illustrated in Figure 5. Figure 6 shows photoluminescence emission intensity spectra of the four different Yb3+ concentrations in the Sm3+-doped fluorotellurite glasses recorded at room temperature under a 450 nm wavelength excitation. An average diode laser power of 60 mW was employed to excite the glass samples.
Figure 6a,b represent the photoluminescence spectra of the Sm3+-doped and Sm3+:Yb3+-codoped barium fluorotellurite glass series in the visible and near-infrared spectral ranges. The visible photoluminescence emission spectra reveal four broad and intense peaks centred at 562 nm (yellow-green band), 598 nm (orange band), 646 nm (red band), and 707 nm (red band), attributing to the 4f-4f interconfigurational transitions. These emission peaks correspond to the 4G5/26H5/2, 4G5/26H7/2, 4G5/26H9/2, and 4G5/26H11/2 transition bands shown in Figure 5. Similarly, the NIR photoluminescence emission spectrum peaks of the Sm3+-doped glass centred at 906, 949, 1030, 1130, 1200, 1293, and 1418 nm correlating with the 4G5/2 state to 6FJ (J = 3/2, 5/2, 7/2, 9/2) and the 6F11/2 state to 6HJ (J = 7/2, 9/2, 11/2) transitions. It is essential to mention that this is the first report on Sm3+ photoluminescence emission peaks occurring at 1200, 1293, and 1418 nm. In addition, fluorotellurite glasses codoping with Sm3+:Yb3+ exhibit a broad photoluminescence emission band range between 850 and 1100 nm owing to the overlap of the Sm3+ (4G5/26F3/2, 6F5/2, 6F7/2) and Yb3+ (2F5/22F7/2) transitions. It can be seen that the photoluminescence emission bands measured at the visible and NIR decrease with an increasing Yb3+ concentration, which could be ascribed to concentration quenching. The photoluminescence emission of sample SBT1 was collected without and with a 495 nm short-pass filter in the NIR wavelength using a 976 nm excitation for comparison as shown in Figure 7a. No significant difference is observed between a 495 nm filter-based and without photoluminescence emission measured under the same laser pump power. Eight transitions are observed in the Sm3+-doped barium fluorotellurite glass around 888, 1126, 1173, 1243, 1350, 1386, 1461, and 1486 nm, which are assigned to the 4G5/26F3/2, 4G5/26F9/2, 6F9/26H7/2, 4G5/26F9/2, 4G5/26F11/2, 6F11/26H11/2, 6F9/26H9/2 transitions, respectively. The intense broadband emission peak between 900 and 1050 nm is attributed to the 976 nm laser first-order. Nevertheless, these transitions were only detected from the Sm3+ single-doped glass, which contradicts the results of the 1.0 mol% Yb3+, 0.5 mol% Sm3+-codoped glass reported by Bolton et al. [25].
Figure 7b shows the photoluminescence emission spectra in the NIR (950–1250 nm) region recorded at room temperature under the same experimental conditions using a 976 nm excitation and a 495 nm filter cut-off wavelength. The spectral revealed two emission bands from the Yb3+ and Sm3+ centred at 1040 and 1125 nm. These emission bands are ascribed to the Yb3+ (2F5/22F7/2) and Sm3+ (4G5/26F9/2) transitions. The full-width–half-maximum (FWHM) values of the as-prepared Sm3+:Yb3+-codoped glasses in the NIR emission at 1040 nm were obtained by fitting with the Gaussian curve. The FWHM values measured are ~43, ~40, and ~35 nm, respectively, corresponding to the SBT2, SBT3, and SBT4 glasses.
The photoluminescence–decay curves of samples SBT2, SBT3, and SBT4 for the Yb3+:2F5/2 transitions were measured under a 976 nm excitation at room temperature, as depicted in Figure 7b. The decay profiles were analysed by fitting with a single exponential function to estimate the decay lifetime. Figure 7c illustrates the photoluminescence– decay curves fitted with a single exponential function of the Yb3+:2F5/2 level observed at 1040 nm. Table 4 presents the fitted single exponential decay equation of the decay curves and R-square parameters. According to Figure 7c, the measured lifetimes of 244, 153, and 45 μs were obtained for glass samples SBT2, SBT3, and SBT4, respectively. The monotonically decreased lifetime is attributed to nonradiative decay processes such as concentration quenching, multiphonon relaxation, more efficient energy transfer from the Yb3+ to the Sm3+, and the coupling of the Yb3+:Sm3+ to the -OH group. It is important to mention that the lifetimes obtained from these samples from the Yb3+:2F5/2 at 1040 nm in the Sm3+:Yb3+-codoped glasses are far greater than those reported elsewhere [25].
Figure 8 illustrates the schematical energy transfer processes between the Yb3+ and Sm3+ ions excited by a 976 nm laser diode. Initially, the 976 nm pump is absorbed from the ground states of the Yb3+: 2F7/2 to the excited states of the Yb3+:2F5/2, respectively. This is followed by spontaneous decay or is de-excited radiatively from the excited states of Yb3+:2F5/2 to 2F7/2 to emit a photoluminescence emission at 1040 nm. The populated photon in the Yb3+:2F5/2 states can act as an indirect pumping scheme to transfer its energy to the neighbouring Sm3+ ions through four main routes such as ET1 (Yb3+:2F5/2 → Sm3+:4G9/2), ET2(Yb3+:2F5/2 → Sm3+:4G5/2), ET3 (Yb3+:2F5/2 → Sm3+:4G11/2), and ET4 (Yb3+:2F5/2 → Sm3+:4G9/2) via an energy transfer process or excited state absorption, respectively. Considering that if the Sm3+:4G9/2 energy level dominates in the energy transfer mechanism, it undergoes resonant nonradiative and nonresonant phonon-assisted energy transfer to the Sm3+:4G5/2 level and then is returned radiatively from Sm3+:4G5/2 to Sm3+:6F9/2 to generate a photoluminescence emission at 1125 nm. On the other hand, similar energy transfer processes can occur via ET3 and ET4 by transferring the majority of the ions in the Yb3+:2F5/2 level to the Sm3+:4G11/2 or 4G9/2 levels, which may decay nonradiatively and radiatively to 6H7/2 and emit a photoluminescence emission observed at 1125 nm. Furthermore, the schematic energy-level diagram in Figure 8 of the Sm3+:Yb3+-codoped glass series shows four possible resonant cross-relaxation processes, CR1, CR2, CR3, and CR4, with a change in energy of practically zero ( E 0 ) as reported elsewhere [25]. The following sequence equations represent these multipolar interactions due to the nonradiative decay: CR1, 6F9/26F7/2 and 6H5/26H7/2 (resonant E 0 ); CR2, 6F9/26F3/2 and 6H5/26H9/2 (resonant E 0 ); CR3, 6F9/26H9/2 and 6H5/26F3/2 (resonant E 0 ); and CR4, 6F9/26H7/2 and 6H5/26F7/2 (resonant E 0 ).

4. Conclusions

The Sm3+- and Sm3+:Yb3+-doped barium fluoride-substituted zinc tellurite glasses were fabricated and characterised for their numerous properties such as density, refractive index, structural, -OH content, and photoluminescence. Adding the appropriate amount of Yb3+ ions into the glass network composition leads to a structural reformation, and the -OH content increases significantly. The photoluminescence emission and decay dynamics of the RE3+ ion single-doped and codoped barium fluorotellurite glasses were investigated under 450 and 976 nm band excitations. The photoluminescence emission spectra were recorded using a 450 nm laser diode excitation to monitor the glasses’ visible and NIR emission spectra. This measurement revealed four intense emission peaks at the yellow-green (4G5/26H5/2,), orange (4G5/26H7/2), and red (4G5/26H9/2 and 4G5/26H11/2) bands in the visible spectrum. In contrast, seven emission peaks were observed at the NIR region corresponding to the radiative emission from the 4G5/2 state to 6FJ (J = 3/2, 5/2, 7/2, 9/2) transitions and the 6F11/2 state to 6HJ (J = 7/2, 9/2, 11/2) transitions. It is observed that the photoluminescence emission intensities measured from visible to NIR decrease with increasing Yb3+ concentrations. This quenching process is attributed to resonant energy transfer processes from Sm3+ to Yb3+ due to a large absorption cross-section of Yb3+. Two emission peaks were observed from the Sm3+:Yb3+-codoped glasses at 1040 and 1125 nm, with a 976 nm excitation correlating to the Yb3+ (2F5/22F7/2) and Sm3+ (4G5/26F9/2) transitions, whereas the photoluminescence lifetime decreases with increasing Yb3+ concentrations. These initial photoluminescence property results of the Sm3+:Yb3+-codoped glasses suggest new NIR windows for developing lasers suitable for medical diagnostics and imaging in anatomical environments.

Author Contributions

Conceptualization, E.K.-B. and A.J.; methodology, E.K.-B. and Y.C.; formal analysis, E.K.-B., Y.C. and M.A.-M.; investigation, E.K.-B., Y.C. and R.T.; resources, A.J.; data curation, E.K.-B., Y.C. and M.A.-M.; writing—original draft preparation, E.K.-B. and A.J.; writing—review and editing M.A.-M., G.S., Y.C., R.T. and A.J.; visualization, E.K.-B., Y.C. and A.J.; supervision, E.K.-B., and A.J.; project administration, E.K.-B.; funding acquisition, A.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Engineering and Physical Sciences Research Council through the following Grants (Grant Nos. EP/M015165/1 and EP/M022854/1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. O’Donnell, M.D.; Miller, C.A.; Furniss, D.; Tikhomirov, V.K.; Seddon, A.B. Fluorotellurite glasses with the improved mid-infrared transmission. J. Non-Cryst. Solids 2003, 331, 48–57. [Google Scholar] [CrossRef]
  2. Li, J.; Xiao, X.; Gu, S.; Xu, Y.; Zhou, Z.; Guo, H. Preparation and optical properties of TeO2-BaO-ZnO-ZnF2 fluorotellurite glass for mid-infrared fibre Raman laser applications. Opt. Mater. 2017, 66, 567–572. [Google Scholar] [CrossRef]
  3. Mandal, P.; Chowdhury, S.; Ghosh, S. Spectroscopic characterisation of Er3+ doped lead zinc phosphate glass via Judd–Ofelt analysis. Bull. Mater. Sci. 2019, 42, 1–6. [Google Scholar] [CrossRef] [Green Version]
  4. Barbi, S.; Mugoni, C.; Montorsi, M.; Affatigato, M.; Gatto, C.; Siligardi, C. Structural and optical properties of rare-earths doped barium bismuth borate glasses. J. Non-Cryst. Solids 2018, 481, 239–247. [Google Scholar] [CrossRef]
  5. Mawlud, S.Q.; Ameen, M.M.; Sahar, M.R.; Mahraz, Z.A.S.; Ahmed, K.F. Spectroscopic properties of Sm3+ doped sodium-tellurite glasses: Judd-Ofelt analysis. Opt. Mater. 2017, 66, 318–327. [Google Scholar] [CrossRef]
  6. El-Mallawany, R.A.H. Tellurite Glass Handbook: Physical Properties and Data; CRC Press: Boca Raton, FL, USA, 2002. [Google Scholar]
  7. Devi, C.B.A.; Mahamuda, S.K.; Swapna, K.; Venkateswarlu, M.; Rao, A.S.; Prakash, G.V. Pr3+ ions doped single alkali and mixed alkali fluoro tungsten tellurite glasses for visible red luminescent devices. J. Non-Cryst Solids 2018, 498, 345–351. [Google Scholar] [CrossRef]
  8. Rivera, V.A.G.; Ledemi, Y.; El-Amraoui, M.; Messaddeq, Y.; Marega, E. Green-to-red light tuning by up-conversion emission via energy transfer in Er3+–Tm3+-codoped germanium–tellurite glasses. J. Non-Cryst. Solids 2014, 392–393, 45–50. [Google Scholar] [CrossRef]
  9. Lin, H.; Liu, K.; Pun EY, B.; Ma, T.C.; Peng, X.; An, Q.D. Infrared and visible fluorescence in Er3+-doped gallium tellurite glasses. Chem. Phys. Lett. 2004, 398, 146–150. [Google Scholar] [CrossRef]
  10. Tang, Y.; Liu, Y.; Cai, P.; Maalej, R.; Seo, H.J. Thermal stability and spectroscopic properties of Ho3+ doped tellurite-borate glasses. J. Rare Earths 2015, 33, 939–945. [Google Scholar]
  11. Tong, T.H.; Yan, X.; Duan, Z.; Liao, M.; Suzuki, T.; Ohishi, Y. Novel tellurite-phosphate composite microstructured optical fibres for nonlinear applications. Proc. SPIE 2012, 9, 2598–2601. [Google Scholar]
  12. Jha, A.; Richards, B.D.O.; Jose, G.; Fernandez, T.T.; Hill, C.J.; Lousteau, J. Review on structural, thermal, optical and spectroscopic properties of tellurium oxide-based glasses for fibre optic and waveguide applications. Int. Mater. Rev. 2012, 57, 357–382. [Google Scholar] [CrossRef]
  13. Pach, K.; Golis, E.; Yousef, E.S.; Sitarz, M.; Filipecki, J. Structural studies of tellurite glasses doped with erbium ions. J. Mol. Struct. 2018, 1164, 32–37. [Google Scholar] [CrossRef]
  14. Santos, F.A.; Figueiredo, M.S.; Barbano, E.C.; Misoguti, L.; Lima, S.M.; Andrade, L.H.C.; Yukimitu, K.; Moraes, J.C.S. Influence of lattice modifier on the nonlinear refractive index of tellurite glass. Ceram. Int. 2017, 43, 15201–15204. [Google Scholar] [CrossRef] [Green Version]
  15. Pollack, S.A.; Chang, D.B. Ion-pair upconversion pumped laser emission in Er3+ ions in YAG, YLF, SrF2, and CaF2 crystals. J. Appl. Phys. 1988, 64, 2885. [Google Scholar] [CrossRef]
  16. Wang, G.; Dai, S.; Zhang, J.; Xu, S.; Hu, L.; Jiang, Z. Effect of F ions on physical and spectroscopic properties of Yb3+-doped TeO2-based glasses. J. Lumin. 2005, 113, 27. [Google Scholar] [CrossRef]
  17. Carnall, W.T.; Fields, P.R.; Rajnak, K. Electronic Energy Levels in the Trivalent Lanthanide Aquo Ions. I. Pr3+, Nd3+, Pm3+, Sm3+, Dy3+, Ho3+, Er3+, Tm3+. J. Chem. Phys. 1968, 49, 4424. [Google Scholar] [CrossRef]
  18. Kaczkan, M.; Frukacz, Z.; Malinowski, M. Infra-red-to-visible wavelength upconversion in Sm3+-activated YAG crystals. Alloys Compd. 2001, 736, 323–324. [Google Scholar] [CrossRef]
  19. Zhou, Y.; Lin, J.; Wang, S. Energy transfer and upconversion luminescence properties of Y2O3:Sm and Gd2O3:Sm Phosphors. J. Solid State Chem. 2003, 171, 391. [Google Scholar] [CrossRef]
  20. Biju, P.R.; Jose, G.; Thomas, V.; Nampoori, V.P.N.; Unnikrishnan, N.V. Energy transfer in Sm3+:Eu3+ system in zinc sodium phosphate glasses. Opt. Mater. 2004, 24, 671. [Google Scholar] [CrossRef]
  21. Farries, M.C.; Morkel, R.P.; Townsend, J.E. Samarium-doped glass laser operating at 651 nm. Electron. Lett. 1988, 24, 539. [Google Scholar] [CrossRef] [Green Version]
  22. France, P.W.; Drexhage, M.G.; Parker, J.M.; Moore, M.W.; Carter, S.F.; Wright, J.V. Fluoride Glass Optical Fibers; CRC Press: Boca Raton, FL, USA, 1990; p. 200. [Google Scholar]
  23. Hussaina, S.; Amjadb, R.J.; Walsh, B.M.; Mehmood, H.; Akbar, N.; Alvi, F.; Doustid, M.R.; Sattar, A.; Iqbal, A.; Hussain, A.; et al. Calculation of Judd Ofelt parameters: Sm3+ ions doped in zinc magnesium phosphate glasses. Solid State Commun. 2019, 298, 113632. [Google Scholar] [CrossRef]
  24. Herrera, A.; Fernandes, R.G.; de Camargo, A.S.S.; Hernandes, A.C.; Buchner, S.; Jacinto, C.; Balzaretti, N.M. Visible–NIR emission and structural properties of Sm3+ doped heavy-metal oxide glass with composition B2O3–PbO–Bi2O3–GeO2. J. Lumin. 2016, 171, 106–111. [Google Scholar] [CrossRef]
  25. Bolton, J.; Jha, A. New NIR emission from Sm3+ in Yb3+-Sm3+ co-doped tellurite glass. J. Lumin. 2021, 231, 117717. [Google Scholar] [CrossRef]
  26. Herrera, A.; Becerra, A.; Balzaretti, N.M. Sm3+/Yb3+ co-doped GeO2-PbO glass for efficiency enhancement of silicon solar cells. Opt. Mater. 2021, 111, 110730. [Google Scholar] [CrossRef]
  27. Nawaz, F.; Sahar, M.R.; Ghoshal, S.K.; Awang, A.; Ahmed, I. Concentration-dependent structural and spectroscopic properties of Sm3+/Yb3+ co-doped sodium tellurite glass. Physica B 2014, 433, 89–95. [Google Scholar] [CrossRef]
  28. Nawaz, F.; Sahar, M.F.; Ghoshal, S.K.; Amjad, J.; Dosti, M.R. Asmahani Awang, Spectral investigation of Sm3+/Yb3+co-doped sodium tellurite glass. Chin. Opt. Lett. 2013, 11, 061605. [Google Scholar] [CrossRef]
  29. Jackson, J.; Smith, C.; Massera, J.; Rivero-Baleine, C.; Bungay, C.; Petit, L.; Richardson, K. Estimation of peak Raman gain coefficients for Barium-Bismuth-Tellurite glasses from spontaneous Raman cross-section experiments. Opt. Express 2009, 17, 19071–19079. [Google Scholar] [CrossRef]
  30. Dilawar, N.; Mehrotra, S.; Varandani, D.; Kumaraswamy, B.V.; Haldar, S.K.; Bandyopadhyay, A.K. A Raman spectroscopic study of C-type rare earth sesquioxides. Mater. Characterisation 2008, 59, 462–467. [Google Scholar] [CrossRef]
  31. Sone, B.T.; Manikandan, E.; Gurib-Fakim, A.; Maaza, M. Sm2O3 nanoparticles green synthesis via Callistemon viminalis extract. J. Alloys Compd. 2015, 650, 357–362. [Google Scholar] [CrossRef]
  32. Aishwarya, K.; Vinitha, G.; Sreevidy, G.; Asokan, S.; Manikandan, N. Synthesis and characterisation of barium fluoride substituted zinc tellurite glasses. Physica B 2017, 526, 84–88. [Google Scholar] [CrossRef]
  33. Lin, S.-B.; Shi, D.-Y.; Feng, A.; Xu, Q.; Zhao, L.; Dong, J.; Liu, C.-X. Spectroscopic properties of Yb3+ doped TeO2–TiO2–Bi2O3 laser glasses. Results Phys. 2020, 16, 102852. [Google Scholar] [CrossRef]
  34. Linganna, K.; In, J.-H.; Kim, S.H.; Han, K.; Choi, J.H. Engineering of TeO2-ZnO-BaO-Based Glasses for Mid-Infrared Transmitting Optics. Materials 2020, 13, 5829. [Google Scholar] [CrossRef] [PubMed]
  35. Maaoui, A.; Haouari, M.; Bulou, A.; Boulard, B.; Ouida, H.B. Effect of BaF2 on the structural and spectroscopic properties of Er3+/Yb3+ ions codoped fluoro-tellurite glasses. J. Lumin. 2018, 196, 1–10. [Google Scholar] [CrossRef]
  36. Zhang, J.; Lu, Y.; Cai, M.; Tian, Y.; Huang, F.; Guo, Y.; Xu, S. 2.8 μm emission and OH quenching analysis in Ho3+ doped fluorotellurite-germanate glasses sensitized by Yb3+ and Er3+. Sci. Rep. 2017, 7, 16794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Fan, X.; Li, K.; Li, X.; Kuan, P.; Wang, X.; Hu, L. Spectroscopic properties of 2.7 μm emission in Er3+ doped telluride glasses and fibres. J. Alloys Compd. 2014, 615, 475–481. [Google Scholar] [CrossRef]
  38. Yan, Y.; Faber, A.J.; de Waal, H. Luminescence quenching by OH groups in highly Er-doped phosphate glasses. J. Non-Cryst. Solids 1995, 181, 283–290. [Google Scholar] [CrossRef] [Green Version]
  39. Balaji, S.; Sontakke, A.D.; Annapurna, K. Yb3+ ion concentration effects on ∼one μm emission in tellurite glass. J. Opt. Soc. Am. B 2012, 29, 1569–1579. [Google Scholar] [CrossRef]
  40. Tanko, Y.A.; Ghoshal, S.K.; Saha, M.R. Ligand field and Judd-Ofelt intensity parameters of samarium doped tellurite glass. J. Mol. Struct. 2016, 1117, 64–68. [Google Scholar] [CrossRef]
  41. Singh, S.P. Chapter 2—Complexes of the rare earth. In Complexes of the Rare Earths; Sinha, S.P., Ed.; Pergamon: Oxford, NY, USA, 1966; pp. 17–23. [Google Scholar]
  42. Azam, M.; Rai, V.K. Effect of the addition of Pb3O4 and TiO2 on the optical properties of Er3+/Yb3+:TeO2-WO3 glasses. ACS Omega 2019, 4, 16280–16291. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Deconvolution of Raman spectra in four vibration frequency regions of prepared glasses: (a) SBT1, (b) SBT2, (c) SBT3, and (d) SBT4.
Figure 1. Deconvolution of Raman spectra in four vibration frequency regions of prepared glasses: (a) SBT1, (b) SBT2, (c) SBT3, and (d) SBT4.
Materials 15 03314 g001
Figure 2. (a) Raman spectra of Sm3+/Sm3+:Yb3+-doped barium fluorotellurite glasses, (b) corresponding vibration intensity ratio of bridge oxygen to nonbridge oxygen atoms vs. Yb2O3 concentration.
Figure 2. (a) Raman spectra of Sm3+/Sm3+:Yb3+-doped barium fluorotellurite glasses, (b) corresponding vibration intensity ratio of bridge oxygen to nonbridge oxygen atoms vs. Yb2O3 concentration.
Materials 15 03314 g002
Figure 3. Infrared (a) transmission and (b) absorption coefficient spectra of Sm3+-doped BaF2-TeO2-ZnO glasses by varying Yb3+ additive concentration.
Figure 3. Infrared (a) transmission and (b) absorption coefficient spectra of Sm3+-doped BaF2-TeO2-ZnO glasses by varying Yb3+ additive concentration.
Materials 15 03314 g003
Figure 4. (a) Optical absorption coefficient spectra of Sm3+ and Sm3+:Yb3+-codoped barium fluorotellurite glasses and (b) corresponding absorption energy-level diagram.
Figure 4. (a) Optical absorption coefficient spectra of Sm3+ and Sm3+:Yb3+-codoped barium fluorotellurite glasses and (b) corresponding absorption energy-level diagram.
Materials 15 03314 g004
Figure 5. Partial energy-level diagram of Yb3+ and Sm3+ under 450 nm excitation source showing ground excitation state absorption, nonradiative processes (NR), cross-relaxation (CR), and resonant energy transfer (ET).
Figure 5. Partial energy-level diagram of Yb3+ and Sm3+ under 450 nm excitation source showing ground excitation state absorption, nonradiative processes (NR), cross-relaxation (CR), and resonant energy transfer (ET).
Materials 15 03314 g005
Figure 6. (a) Visible and (b) near-infrared (NIR) wavelengths photoluminescence spectra of different concentrations of Yb3+ in Sm3+-doped barium fluorotellurite glass series.
Figure 6. (a) Visible and (b) near-infrared (NIR) wavelengths photoluminescence spectra of different concentrations of Yb3+ in Sm3+-doped barium fluorotellurite glass series.
Materials 15 03314 g006
Figure 7. Near-infrared photoluminescence emission spectra of Sm-doped and Sm3+: Yb3+-codoped glasses under 978 nm excitation source: (a) sample SBT1 with 495 short-pass filter and without (an inset shows emission spectra from 1100 to 1500 nm), (b) with 450 nm filter, (c) photoluminescence–decay curves of different Yb3+ content in glasses at 2F5/2 transition.
Figure 7. Near-infrared photoluminescence emission spectra of Sm-doped and Sm3+: Yb3+-codoped glasses under 978 nm excitation source: (a) sample SBT1 with 495 short-pass filter and without (an inset shows emission spectra from 1100 to 1500 nm), (b) with 450 nm filter, (c) photoluminescence–decay curves of different Yb3+ content in glasses at 2F5/2 transition.
Materials 15 03314 g007
Figure 8. Energy-level diagrams of Sm3+ and Yb3+ observed in the NIR emission transitions under 976 nm excitation and possible energy transfer processes.
Figure 8. Energy-level diagrams of Sm3+ and Yb3+ observed in the NIR emission transitions under 976 nm excitation and possible energy transfer processes.
Materials 15 03314 g008
Table 1. Molar composition, density, and refractive index of the Sm3+/Sm3+:Yb3+-doped barium fluorotellurite glass samples.
Table 1. Molar composition, density, and refractive index of the Sm3+/Sm3+:Yb3+-doped barium fluorotellurite glass samples.
Sample IDComposition (mol%)Density
(g/cm3)
Refractive Index
TeO2ZnOBaF2Sm2O3Yb2O3
SBT179.5010100.50-5.6221 ± 0.00122.0021 ± 0.0005
SBT279.2510100.500.255.6237 ± 0.00142.0060 ± 0.0005
SBT379.0010100.500.505.6217 ± 0.00372.0083 ± 0.0017
SBT478.5010100.501.005.6183 ± 0.00302.0141 ± 0.0013
Table 2. -OH absorption coefficient, α OH (cm−1), and concentration, N OH ( ions / cm 3 ) , at mid-infrared wavelengths of 3.3 and 4.38 μm.
Table 2. -OH absorption coefficient, α OH (cm−1), and concentration, N OH ( ions / cm 3 ) , at mid-infrared wavelengths of 3.3 and 4.38 μm.
Sample ID α OH   ( cm 1 ) -OH Concentration (×1019 ions/cm−3)
3.3 μm4.38 μm3.3 μm4.38 μm
SBT11.5161.2411.859 ± 0.0221.522 ± 0.018
SBT22.9732.2563.645 ± 0.0412.766 ± 0.025
SBT33.1172.2783.822 ± 0.0122.793 ± 0.017
SBT44.6553.4925.707 ± 0.0374.281 ± 0.006
Table 3. Transition energy levels and corresponding wavenumber, aquo-ion wavenumber, nephelauxetic effect ( β ) , and bonding parameters.
Table 3. Transition energy levels and corresponding wavenumber, aquo-ion wavenumber, nephelauxetic effect ( β ) , and bonding parameters.
TransitionEnergy Levels (cm−1)Aquo-ion (cm−1) [17]
STB1SBT2SBT3SBT4
6H5/26F11/210,52110,44410,35410,31210,517
6H5/26F9/292409229924492409136
6H5/26F7/281118122811681087977
6H5/26F5/272247235724172337131
6H5/26F3/267196719671967146641
6H5/26H13/264456435645664566508
6H5/26F1/262926270629262836397
β 7.0238 ± 0.00687.0202 ± 0.00467.0165 ± 0.00817.0078 ± 0.00317.0000
β ¯ 1.0034 ± 0.01371.0029 ± 0.01461.0024 ± 0.01501.0011 ± 0.01121
σ −0.339 ± 0.002−0.287 ± 0.005−0.235 ± 0.008−0.111 ± 0.004-
Table 4. Lifetime decay equation, lifetime, and R-square of the equation.
Table 4. Lifetime decay equation, lifetime, and R-square of the equation.
Sample IDDecay Equation ProfileLifetime (μs)R-Square
SBT2 I ( t ) = 0.023 + 0.969 exp ( t 0.244 ) 244 ± 0.30.996
SBT3 I ( t ) = 0.022 + 0.997 exp ( t 0.154 ) 154 ± 0.40.996
SBT4 I ( t ) = 0.004 + 1.235 exp ( t 0.043 ) 43 ± 0.10.998
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kumi-Barimah, E.; Chen, Y.; Tenwick, R.; Al-Murish, M.; Sherma, G.; Jha, A. Effect of Yb3+ on the Structural and Visible to Near-Infrared Wavelength Photoluminescence Properties in Sm3+-Yb3+-Codoped Barium Fluorotellurite Glasses. Materials 2022, 15, 3314. https://doi.org/10.3390/ma15093314

AMA Style

Kumi-Barimah E, Chen Y, Tenwick R, Al-Murish M, Sherma G, Jha A. Effect of Yb3+ on the Structural and Visible to Near-Infrared Wavelength Photoluminescence Properties in Sm3+-Yb3+-Codoped Barium Fluorotellurite Glasses. Materials. 2022; 15(9):3314. https://doi.org/10.3390/ma15093314

Chicago/Turabian Style

Kumi-Barimah, Eric, Yan Chen, Rebekah Tenwick, Mohanad Al-Murish, Geeta Sherma, and Animesh Jha. 2022. "Effect of Yb3+ on the Structural and Visible to Near-Infrared Wavelength Photoluminescence Properties in Sm3+-Yb3+-Codoped Barium Fluorotellurite Glasses" Materials 15, no. 9: 3314. https://doi.org/10.3390/ma15093314

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