Next Article in Journal
Electrochemical Behavior of Plasma-Nitrided Austenitic Stainless Steel in Chloride Solutions
Previous Article in Journal
Fabrication and Characterization of Polycaprolactone–Baghdadite Nanofibers by Electrospinning Method for Tissue Engineering Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Investigation of the Photonic Application of TeO2-K2TeO3-Nb2O5-BaF2 Glass Co-Doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 at 1.54 μm Based on Its Thermal and Luminescence Properties

1
Laboratory of Systems Integration and Emerging Energies, National Engineering School of Sfax (ENIS), University of Sfax, Sfax 3018, Tunisia
2
Department of Physics, Faculty of Science and Arts, Najran University, Najran 11001, Saudi Arabia
3
Promising Centre for Sensors and Electronic Devices (PCSED), Advanced Materials and Nano-Research, Najran University, Najran 11001, Saudi Arabia
4
LaMaCoP, Faculty of Sciences of Sfax, University of Sfax, Sfax 3018, Tunisia
5
Department of Metallurgical and Materials Engineering, Yildiz Technical University, Davutpasa Campus KMA-205, Esenler, Istanbul 34220, Turkey
6
Physics Department, Faculty of Science, Al Azhar University, Assiut 71542, Egypt
7
Department of Physics, Faculty of Science, King Khalid University, Abha 61421, Saudi Arabia
8
Department of Radiological Sciences, College of Applied Medical Sciences, King Khalid University, Abha 61421, Saudi Arabia
*
Author to whom correspondence should be addressed.
Materials 2024, 17(17), 4188; https://doi.org/10.3390/ma17174188
Submission received: 29 July 2024 / Revised: 13 August 2024 / Accepted: 20 August 2024 / Published: 23 August 2024

Abstract

:
A glass composition using TeO2-K2TeO3-Nb2O5-BaF2 co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 was successfully fabricated. Its thermal stability and physical parameters were studied, and luminescence spectroscopy of the fabricated glasses was conducted. The optical band gap, Eopt, decreased from 2.689 to 2.663 eV following the substitution of Ho2O3 with Yb2O3. The values of the refractive index, third-order nonlinear optical susceptibility (χ(3)), and nonlinear refractive index (n2) of the fabricated glasses were estimated. Furthermore, the Judd–Ofelt intensity parameters Ω t   ( t = 2 , 4 , 6 ) , radiative properties such as transition probabilities (Aed), magnetic dipole-type transition probabilities (Amd), branching ratios (β), and radiative lifetime (τ) of the fabricated glasses were evaluated. The emission cross-section and FWHM of the 4I13/24I15/2 transition around 1.54 μm of the glass were reported, and the emission intensity of the visible signal was studied under 980 nm laser excitation. The material might be a useful candidate for solid lasers and nonlinear amplifier devices, especially in the communications bands.

1. Introduction

Due to their unique optical characteristics, glasses are frequently utilized as optical materials. In recent years, there has been an obvious increase in interest in the possible applications of tellurite glasses due to their intriguing and significant optical features. Tellurite glasses have been used to fabricate a wide range of devices, such as planar waveguides, nanowires, and optic amplifiers [1,2,3,4]. Numerous studies have been conducted to examine the physical characteristics of several tellurite (TeO2) glass combinations. TeO2 glasses have attracted some technical and scientific interest because of their many applications [5]. Technological optical fiber devices, lasers, optical fibers, solar cells, sensors, memory-switching devices, gas sensors, optoelectronics, and optical waveguide applications have all demonstrated significant potential for using these glasses [5,6,7,8]. Furthermore, due to TeO2’s excellent nonlinear optical characteristics, low melting point, superior chemical stability, and elevated index of refraction, it has drawn greater interest as a glass former than other glass formers such as silicates and phosphate [9,10,11]. Given their huge transparent window, great refractive index, and outstanding stability, TeO2 glasses containing a heavy metal oxide—such as Nb2O5, which is a heavy metal tellurite glass—are appealing for the additional development of infrared lasers and amplifiers [12]. Furthermore, niobic TeO2 glasses have an elevated third-order nonlinear optical susceptibility, which makes them a viable option for nonlinear fiber devices such as all-optical switches [12]. The contents of non-bridging oxygen (NBO) and TeO3 units both rise when modifiers such as BaO2 disrupt the random arrangement of glasses [13]. Moreover, the alkaline earth metal element barium (Ba) has a high basicity, a large ionic radius, and a higher atomic number. Consequently, the presence of Ba in the TeO2 glass framework modifies the glass’s construction and enhances its chemical stability, density, gloss, and refractive index [14,15]. When fluorine ions are added to TeO2 glass, the glass’s formation range is increased, its viscosity is decreased, its degree of transparency is enhanced, and its moisture resistance is increased [16]. Also, adding fluorine to TeO2 glasses disintegrates the network of the glassy system due to fluorine’s electronegativity being greater than that of oxygen [16]. This implies that the network’s configuration of the glass in these kinds of glasses is impacted by the fluoride’s replacement of oxygen. When the framework is altered, a lot of fundamental characteristics also change. Unlike both oxygen and fluorine matrix structures, an oxide–fluoride glass matrix might offer rare earth ions a unique home. As a result, oxyfluoride glasses with a high proportion of rare earth elements are novel beneficial substances [17]. Fluoride glasses have a minimal cost limit and a wide spectrum range, which makes them ideal for optical fiber applications involving sensors. It is commonly known that increasing a glass matrix’s metal fluoride compounds improves its transparent nature and refractive index [9]. The prospective utilization of co-doping of glasses with rare earth ions in solid-state lasers, three-dimensional displays, and optical amplifiers has garnered significant interest in the last decade [18]. The rare earth ions erbium (Er3+), ytterbium (Yb3+), and holmium (Ho3+) have received the most attention. Er3+, Ho3+, and Yb3+ ion insertion into the matrix of glass allows for the production of up-conversion luminescence when it is exposed to mid-infrared (MIR) rays. A high-power laser diode energy of 980 nm might be used to actively excite Er3+ ions. According to the mutual concentration of these ions, a transfer of energy process among them might then alter the up-conversion emission intensity [19]. Moreover, erbium-doped glasses are extensively utilized in a diversity of optical implementations, mostly in the areas of eye-safe lasers and optical amplifiers for fiber networking [18,20,21]. Furthermore, the observable up-conversion emission can be amplified by co-doping the host substance with Ho2O3/Er2O3 or Yb2O3/Er2O3 couplings. This is because of a greater absorbing cross-section and greater energy transmission operations from Ho3+ to Er3+ ions or from Yb3+ to Er3+ ions, respectively [18,22,23]. The present work aimed to prepare TeO2-based glass (70TeO2-15K2TeO3-10Nb2O5-5BaF2) that was co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 and to investigate the impact of these rare earth oxides on the thermal and optical properties of the glasses. Differential scanning calorimetry and a double-beam spectrophotometer were utilized to study these characteristics. Furthermore, optical spectrum measurements were conducted to estimate the optical factors. The energy levels of the glass co-doped with Er2O3/Yb2O3 were determined, and the branching ratios (β), radiative lifetimes (τrad), electric dipole-type transition probabilities (Aed), magnetic dipole-type transition probabilities (Amd), and Judd–Ofelt intensity factors Ωt (t = 2, 4, 6) were calculated.
The primary goal of this work was to investigate how the amount of Yb3+ and Ho3+ affects the spectroscopic characteristics of Er3+-co-doped TeO2 glasses. This will help maximize the transition’s gain and emission cross-section between 4I13/24I15/2 and will also determine whether or not these glasses are suitable as optical glasses for laser and fiber amplifiers. Utilizing measurements of the absorption spectra and McCumber theory, the absorbing, emitting, and gain cross-sections of the 4I13/24I15/2 transition were derived at approximately 1.54 µm. Finally, the emission intensity of the visible signal was studied under 980 nm laser excitation. The FWHM of the 4I13/24I15/2 transition of the glass was reported. Furthermore, we estimated the nonlinear refractive index and third-order susceptibility of fabricated glass which have good transmission with multiple absorption peaks in the near-infrared wavelength spectrum. Therefore, fabricated glass is a unique property that can used in nonlinear devices.

2. Experimental Section

The TeO2 glasses with the composition 70TeO2-15K2TeO3-10Nb2O5-5BaF2 (in mol%) co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 were produced using a melt-quenching process. All the starting chemicals were procured from Aldrich and were 99.99% pure. The particulars of their compositions are displayed in Table 1.
After mixing, the mixture was heated for 25 min at 950 °C in a platinum crucible within a furnace. After that, a graphite mold was filled with the extremely viscous melt. Following two hours of annealing at 270 °C, the quenched glass was gradually cooled to room temperature (RT). Figure 1 shows pictures of the sample glasses as they were produced. The samples were cut and polished to 2.1 mm thick. To investigate the thermal properties of these glasses, a DSC Shimadzu 50 with a resolution ± 1.0 °C at a heating rate of 10 °C/min over a temperature range of 550 °C was employed. Toluene, a known-density immersing solution (0.8669 g/cm3), was employed using the Archimedes method with a resolution ± 0.001 g/cm3 to determine the glass sample’s density at RT. In the 190–2500 nm wavelength range with a resolution of 1nm, the optical absorbing and transmission spectra were attained utilizing a JASCO V-570 spectrophotometer.

3. Results and Discussion

3.1. Thermal Characteristics

The DSC thermograms for the 70TeO2-15K2TeO3-10Nb2O5-5BaF2 glass co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 at a rate of 10 °C/min are shown in Figure 2. These graphs demonstrate that at the glass transition, distinct endothermic peaks are seen, followed by an exothermic crystallization peak. Table 2 lists the glass transition temperature (Tg), crystallization start temperature (Tc), and peak crystallization temperature (Tp) for the investigated glasses.
The glassy nature of the current samples was confirmed based on the curve forms. Additionally, Tg offers details regarding the glass network’s interconnectivity and binding strength. It is understood that Tg increases as the glass’s interconnectivity and binding strength grow [24]. These values of (Tg) are close to the tellurite-based glasses [25]. The distinction between Tc and Tg is employed to determine the thermal stability (ΔT) [26,27] of the produced glasses (ΔT = TcTg). The glass’s excellent thermal stability is indicated by the significant difference between Tc and Tg. Another method for calculating the glass’s stability would be to use Sestak’s estimation of the Hruby index, H = ΔT/Tg [28,29]. The ∆T and H are shown in Table 2, and these are important for determining the glass devitrification process [27]. The next relation could be employed to evaluate the value of the factor KSP, which is associated with glass’s stability versus crystallization [28,29,30,31]:
K S P = T p T c T p T g T g
The KSP magnitudes of the produced glasses are documented in Table 2. It can be seen that all these values of the thermal stability parameters (∆T, H, KSP) of the TKNB glass co-doped with Er2O3/Yb2O3 (TKNB2) are slightly lower than those of the TKNB glass co-doped with Er2O3/Ho2O3 (TKNB1). This may suggest a decrease in rigidity in the glassy matrix due to the replacement of Ho2O3 with Yb2O3. On the other hand, this decrease in thermal stability due to the substitution of Ho2O3 with Yb2O3 might be credited to the following causes: (i) the low bond strengthening Yb-O (387.7 kJ mol−1) in contrast to Ho-O (606 kJ mol−1) [32]; or (ii) the cation radius of Yb3+ (2.28 Å) being slightly lower than the cation radius of Ho3+ (2.33 Å). In addition to reducing the length of bonds and producing a high coulombic force of attraction between opposing ions, cation’s radius is directly proportional to polarizability [33].

3.2. Density and Molar Volume

The densities of the examined glasses were determined by utilizing Archimedes’ principle. We used toluene (ρ0 = 0.8669 g/cm3) as the immersing fluid. The density (ρ) was computed employing the equation presented below after the weights of the glasses were first measured separately in the air (Wa) and then in the previously mentioned liquid (Wl).
ρ = W a W a W l ρ 0
Using the formula Vm = Mw/ρ, where Mw is the glass sample’s molecular weight, the molar volume (Vm) was computed. The values of ρ and Vm are listed in Table 1. It is worth mentioning that the substitution of Ho2O3 with Yb2O3 leads to a slight increase in glass density, while the Vm slightly decreases. This phenomenon could be ascribed to both the higher ρ and the higher Mw of Yb2O3 (9.17 g/cm3, 394.08 g/mol) compared with Ho2O3 (8.41 g/cm3, 377.86 g/mol). Given that the ρ is inversely correlated to the Vm and is proportionate to the average Mw, it is generally predicted that the two quantities will behave in opposition to one another.

3.3. Optical Properties

The spectra of optical transmission of the TKNB glasses co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 are provided in Figure 3, which indicates that the fabricated glasses are highly transmissible in the range of UV–VIS–NIR wavelengths. Figure 4 demonstrates several peaks in the spectra, which are due to the presence of Er3+, Ho3+, and Yb3+ ions in the TKNB glass system co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3. The absorption bands corresponding to the transition from 4I15/2 to the ground states 5I7(Ho3+), 4I13/2, 5I6(Ho3+), 4I11/2(Er3+) + 4F5/2(Yb3+), 4I9/2, 4F3/2, 4S3/2, 2H11/2, 4F7/2, 4F3/2, and 2G9/2 can be attributed to the following wavelengths: 1950, 1534, 1155, 970, 803, 655, 540, 520, 490, 450, and 420 nm. These correspond with the reported wavelengths in similar glasses [21,22,23].
The absorption coefficient (α) of the examined glasses is computed using the following formula [32,33,34]:
α = 1 d ln I 0 I t = 2.303 O D L
where L is the thickness of the sample before and after the light traverses the sample, its intensities are I0 and It, and OD is the optical density.
The spectrum of absorption of glasses with varying compositions is used to study changes in the optical band gap (Eopt) and refractive index (n). The Eopt, which is the difference between the highest energy in the valence band and the lowest energy in the conduction band, and temperature both affect how many electrons are excited toward the conduction band. When a material’s Eopt falls within the ranges of 0 to 4 eV, it is classified as a semiconductor. A substance is classified as an insulator if its Eopt value is substantial, falling between 4 and 12 eV. These statistics are crucial for designing semiconductor devices, since the energy gap determines the electrical and optical characteristics of the device. Clarifying the transformation and electronic band construction of substances has been shown to have been an extremely beneficial effect in research on the optical absorbing edge in the UV area. Using this connection, the optical band gap Eopt of the specimens is computed from their absorption patterns [35].
α   h ν = b h ν E o p t s
where b is a constant, hv represents the energy of the incoming photon, and α is the absorbing coefficient. An index called s, which is equivalent to ½, 2, 3/2, or 3 for directly allowed, indirectly allowed, directly forbidden, and indirectly forbidden transitions, respectively, is used to describe the optical absorption processes. In this instance, indirect transmissions in substances are covered by this formula with s = 2, which is employed to characterize experimental findings for amorphous substances [33,34,35]. The change in (αhν)1/2 vs. for indirect transitions is shown in Figure 5. By means of extrapolation of the linear fitting of the curve in the UV–VIS nm range across the axis at (αhν)1/2 = 0, the Eopt of the fabricated glasses was determined. Table 3 lists each value of Eopt.
It is discovered that the Eopt values of the two glass samples are lower than the pristine TeO2 glass’s 3.79 eV value, determined in a previous study [36]. The values of Eopt for the TKNB glass co-doped with Er2O3/Yb2O3 (TKNB2) are slightly lower than the TKNB glass doped with Er2O3/Ho2O3 (TKNB1). The BO in the glass-forming system is altered by the addition of rare earth ions, and any alteration in BO, including the creation of NBO, alters the absorption properties, which in turn reduces the optical band gap. The TeO2 matrix’s coordination number varies when rare earth is added. The sample matrix’s structural organization and chemical makeup both have an impact on the optical band gap [37]. When Ho2O3 is replaced with Yb2O3, the slight decrease in Eopt is because of the substitution of the Ho-O bond with the Yb-O bond. The decrease in Eopt with the addition of Yb2O3 can be elucidated by the change in electron density [38]. When Yb2O3 is added to a glass sample instead of Ho2O3, the amount of NBO may increase, which consequently decreases the Eopt [37]. The value of the materials’ molar mass is the cause of the increase in NBO [39]. In comparison to the molar mass of Ho2O3, the molar mass of Yb2O3 is higher. Thus, there is an increase in NBO atoms, which lowers the value of Eopt for the Yb2O3-containing glass sample. Additionally, the Eopt falls under the category of semiconductor substances, ranging from 2.689 eV to 2.663 eV [40].
The following equations characterize the extinction coefficient (k) and n of the examined glasses [41]:
k = α λ 4 π
R = n 1 2 + k 2 n + 1 2 + k 2 ;   n = 1 + R 1 R + 4 R 1 R 2 k 2 1 2
where R is the reflectance. Figure 6 and Figure 7 show the computed values of the k and n of the examined glasses, respectively. Both the variations in structure and the fluctuation in the incident wavelength (λ) can affect the n and k. According to Figure 7, the n drops as the λ increases. The following figure shows the various optical factors that are determined using the n and k spectrum distributions vs. λ for the tested glasses.
The data for the refractive indexes in Figure 7 are fitted to a three-term Sellmeier equation:
n 2 λ = A + B / 1 C λ 2 + D / 1 E λ 2
where A, B, C, D, and E are the glass substance dispersion factors (Sellmeier coefficients), and λ is the wavelength in μm. The n is influenced by both smaller and greater energy gaps from electronic adsorption, as indicated by the first and second terms. The final term indicates the refractive index-lowering effect of network absorbing [42]. The square of twice the IR transmitting edge could be utilized for calculating the E [42,43]. Table 3 displays Sellmeier’s factors, which were obtained by fitting the experimental data [44] with the use of Equation (7). The Wemple and DiDomenico (WDD) single-oscillator model was utilized to study the n dispersion [45]. The relation between n and can be represented as follows utilizing this model:
n 0 2 1 1 = E o E d 1 E d E o h ν 2
where Ed is the dispersion energy, which is an estimate of the oscillator strength or average strength of the interband optical transition, and Eo is the oscillator energy. Equation (8) may be used to derive the Eo and Ed through graphing (n2−1)−1 versus (hv)2, as Figure 8 illustrates. Applying the fit of the linear parameters, Eo and Ed may be calculated based on the graph. The intercepts and slopes of the arcs yield the Ed and Eo values.
When hv→0, the Wemple–DiDomenico dispersion relationship, Equation (8), is extrapolated to obtain the static refractive index (n0) of the as-prepared glasses, which yields the following formula:
n 0 = 1 + E d / E o
The determined values of Eo, Ed, and no are recorded in Table 4. It can be seen that the value of no for TKNB glass doped with Er2O3/Yb2O3 (TKNB2) is higher than that of TKNB glass doped with Er2O3/Ho2O3 (TKNB1) due to the higher density of the TKNB2 sample.
Important factors for the production of optical equipment, such as fiber optic and laser substances, are the n, molar polarizability (αm), and molar refraction (Rm). Consequently, the following formulas [34] are applied to calculate these properties of the examined glasses:
R m = n o 2 1 n o 2 + 2 V m
α m = 3 4 π N A R m
where NA is Avogadro’s number. The Rm and αm values are recorded in Table 3. These values for TKNB glass doped with Er2O3/Yb2O3 (TKNB2) are higher than those of glass doped with Er2O3/Ho2O3 (TKNB1). This provides a qualitative explanation for the rise in refractive indices observed when Yb2O3 replaces Ho2O3. We conclude that a rise in molar refraction may be the cause of the observed pattern, which is a rise in the oxide ion polarizability, accompanied by an elevation in n. This relationship [31] was utilized to estimate the metallization criterion (M) for the as-prepared glasses:
M = 1 R m V m
The metallic or non-metallic character of the substance is indicated by the M. M < 0 suggests a metallic nature of the substances, whilst a value of M > 0 indicates an insulating nature. The M values are listed in Table 3 and were between 0.503 and 0.507. Therefore, the produced glasses demonstrated an insulating nature [30,31]. On the other hand, the exchange of Ho2O3 with Yb2O3 causes an upsurge in the valance band’s width and a decrease in M and, thus, a decrease in Eopt. As shown in Table 3, the glass doped with Er2O3/Ho2O3 had higher values of M and Eopt, while the glass co-doped with Er2O3/Ho2O had smaller values of M and Eopt.
The nonlinear optical parameters can be determined based on Miller’s rule. The dispersion of the optical linear susceptibility χ(1) and the third-order nonlinear optical susceptibility χ(3) are, respectively, deduced based on their empirical relations [46].
χ 1 = n o 2 1 4 π
χ 3 = χ 1 4 × 1.7 × 10 10
The nonlinear refractive index n2 of the examined glasses relates to the third-order nonlinear optics χ(3) and static refractive index no, which can be determined using the following equation [47,48]:
n 2 = 12 π n o χ 3
The calculated values of χ(1), χ(3), and n2 are listed in Table 4. It can be observed that these values are higher for the TKNB glass doped with Yb2O3/Er2O3 (TKNB2) than for the TKNB glass doped with Ho2O3/Er2O3 (TKNB1). The reported results demonstrate an increase in the nonlinear susceptibility, and the examined glasses show high values of optical linear susceptibility, χ(1), replacing an atom with an atomic radius that is smaller than the removed one. This also applies when replacing Ho3+ (atomic radii = 2.33 Å) with Yb3+ (atomic radius = 2.28 Å) in the TKNB glass doped with Yb2O3/Er2O3 (TKNB2), which causes an increase in the χ(3) value. This is dependent on the value of the third-order nonlinear optical susceptibility, χ(3), and nonlinear refractive index, n2, which suggest that these glasses can be used in nonlinear optical devices, especially in the communication bands.

3.4. Absorption Spectra and Judd and Ofelt Analysis

Figure 9 displays the Er3+/Yb3+-co-doped glasses and their absorption spectra in the UV–VIS–NIR wavelength between 400 and 1800 nm. The 4f11-4f11 transitions of Er3+ and Yb3+ are responsible for all of the bands. Three bands at 800, 975, and 1530 nm are detectable in the infrared section of the spectrum (Figure 9a). The first and third bands result from Er3+ transitions between 4I9/2 and 4I13/2, which are ground states, and 4I15/2. With an increased optical density, the second band represents an overlapping absorption among the two transmissions, 2F7/22F5/2 (originating from Yb3+ ion) and 4I15/24I11/2 (originating from Er3+ ion). As seen in Figure 9b, the UV–visible region of the absorbed spectrum demonstrates well-resolved lines, which are ascribed to the transition of Er3+ ions in the ground state, 4I15/2 to the 4F3/2, 4F3/2, 4F5/2, 4F7/2, 4H11/2, 4S3/2, and 4F9/2. These absorption bands were assigned and located in accordance with Carnall et al. [49] and all relevant research.
The absorption spectra of this glass medium (TKNB2) were subjected to a Judd and Ofelt (JO) investigation in order to ascertain their spectroscopic characteristics. Below is a quick synopsis of the JO investigation.
The well-known theory developed by JO in 1962 makes it possible to compute the probability of an electric dipole transition between rare earth ion energy levels in diverse contexts [50]. Theoretically, the electric S e d c a l and magnetic ( S m d ) dipole line strengths can be utilized for describing the radiative transmissions of the first J level to the succeeding J′ level in the 4fn conformation of rare earth ions [51,52].
The calculated line strength S e d c a l c S L J , S L J between the initial state J , described by S , L , J , and the final state J , given by ( S , L , J ) , can be expressed using the following relation [53,54]:
S e d c a l c S L J , S L J = t = 2,4 , 6 Ω t S L J U ( t ) S L J 2
The influence of the host glasses on the luminescence intensity is represented by the JO intensity factors, Ω t   ( t = 2 ,   4 ,   6 ) , and the doubly reduced matrix elements of grade t among the statuses with the quantum numbers of ( S , L , J ) and S , L , J are S L J U ( t ) S L J 2 [51]. The reduced matrix elements, which may be obtained in the literature [52,53,54,55,56,57,58], are only dependent on the angular momentum of the Er3+ states. The reduced matrices are defined as the total of the relevant matrix elements for two or more manifolds. Table 5 provides the matrix element values for each Er3+ absorption band.
Meanwhile, the measured electric dipole line strength S e d m e a s S L J , S L J can also be calculated from the absorption spectra (Figure 1) [56,57,58] by using the following equation:
S e d m e a s = 9 n n 2 + 2 2 4 π ϵ 0 3 c h 2 J + 1 8 π 3 e 2 × 2.303 L N λ ¯ J J j O D λ d λ n · S m d S L J , S L J
where J is the total angular momentum quantum number of the ground state J = 15 2 , e is the charge of the electron, and N is the ion content (ions/cm3). The average wavenumber of the absorbing band is denoted by λ ¯   ( nm ) , the refractive index of the host with respect to λ ¯ is represented by n, L is the thickness of the sample under study (L = 1 mm), and J J j O D λ d λ symbolizes the experimentally determined integrated optical density in the respective wavelength ranges.
The magnetic dipole transitions S m d S L J , S L J contribute as follows [59]:
S m d S L J , S L J = h 4 π m c 2 S L J L + 2 S S L J 2
The symbols m and c stand for the electron mass and light velocity, respectively. The magnetic dipole matrix elements between the LS-coupled states are represented by S L + 2 S S L J [60].
Table 6 demonstrates that the only transition with a contribution of the magnetic dipole, S m d = 0.7148 × 10 20   c m 2 , in the case of Er3+ ions is the 4I15/24I13/2 ( s e l e c t i o n   r u l e s :   J = 1 ) .
The measured line strengths ( S e d m e a s ) of the electric dipoles were utilized to calculate the values of the JO intensity factors Ω 2 , Ω 3 ,   a n d   Ω 6 . If the JO factors produce a column vector Ω , and the double-decreased matrix elements produce an n × 3 matrix A , where n is the number of transmissions to fit and 3 matches the three JO factors, then the measured electric dipole line strength can be transcribed as a 1 × n column vector. The equality among Equations (13) and (14) can be articulated as S e d m e a s = A . Ω . The group of JO factors was obtained from the matrix Ω = ( A T . A ) 1 . A T . S e d m e a s , where A T is the transposition of matrix A . This matrix-based process is perfect for computations that are carried out on a computer.
The optimal modification was performed, accounting for the initial six transitions and yielding the following values: Ω 2 = 2.387 × 10 20   c m 2 , Ω 4 = 1.881 × 10 20   c m 2 , and Ω 6 = 0.657 × 10 20   c m 2 . These computed Ω t values were utilized in Equation (12) to obtain the values of S e d c a l c . Table 6 provides an overview of the outcomes of some of the calculated parameters, as well as the S e d c a l c and S e d m e a s absorption line strengths for the Er3+/Yb3+-co-doped TKNB2 sample.
Table 6 presents the values of some of the factors that were employed in computation, as well as the S e d m e a s and S e d c a l c absorbing line strengths for the Er3+/Yb3+-co-doped TKNB2 sample.
The formula below indicates the root mean square δ r m s difference between the predicted and observed line strengths of the transitions, which serves as a gauge for the fit’s accuracy:
δ r m s = i = 1 N t r a n s S e d m e a s ( i ) S e d c a l ( i ) N t r a n s 3 1 / 2
where N t r a n s is the number of strong absorption transitions.
The value found in this study, δ r m s = 0.0361 × 10 20   c m 2 , is tiny when compared with values found in other kinds of glasses [61,62,63,64,65,66].
The calculated Ω t parameters for different glasses accord well with those reported in existing works. The Ω t measurements of Er3+ ions in several other common glasses [67,68,69,70,71,72,73] are shown in Table 7. The intensity parameters comprise two terms, based on the JO theory [74]. Firstly, the symmetry and distortions associated with the constructional alteration in the presence of rare earth ions are described by the crystal field parameter. The other represents the covalency between the ligand anions and doped rare earth ions, which is connected to the excited opposite parity electronic states and 4f radial integral states of wave functions. Further, the covalent bonding of binder anions and rare earth ions (less ionic in nature) in the host and the symmetry of the immediate environment surrounding them are correlated with the intensity parameter Ω 2 . The degree of asymmetry surrounding rare earth sites increases with an increasing value of Ω 2 , indicating a greater covalency between the metal and ligand link. The bulk characteristics of the glass framework, such as stiffness, viscosity, and basicity, are primarily described by the intensity factors Ω 4 and Ω 6 , which are also influenced by the acidity and alkalinity of the host substance [75]. The host material’s hardness and basicity increase and decrease, respectively, with higher values of Ω 4 and Ω 6 . Table 7 demonstrates that all three intensity parameters exhibit an upward trend of Ω 2 > Ω 4 > Ω 6 , which aligns with the majority of results from previous studies. A greater asymmetry and superior covalence are indicated by the higher Ω 2 value compared with Ω 4 and Ω 6 [76]. In the meantime, all three Er3+ intensity characteristics that were measured in the current study were reduced to within the values of other glass hosts, suggesting a local asymmetry associated with the Er3+/Yb3+ ions and moderate covalency of the Er-O/Yb-O bonds in the current TeO2 glass. The manufactured glass’s spectroscopic quality factor, Q = Ω 4 / Ω 6 , has a value of 2.8630. When compared with the glass hosts ZBLAN [77], Boro-tellurite [78], and TLNT [79], the generated glass had a higher value of Q = Ω 4 / Ω 6 . The glass that was developed had higher spectroscopic quality factor values, Q = Ω 4 / Ω 6 , than those found for TeO2 glasses in earlier investigations, indicating that it is a better fit for optical devices. For optical amplifier and laser usage, the glass used in the present research (TKNB2) is a great option [80].
The relationship between the emission line strengths related to the transmissions from the highest energy levels, 4I15/2, 4I13/2, 4I11/2, 4I9/2, 4F9/2, 4S3/2, 2H11/2, and 4F7/2, and their related lower-lying energy levels can be computed utilizing the JO intensity parameters. The spontaneous emission probabilities A r a d ( J J ) can be calculated utilizing these line strengths as follows:
A r a d J J = A e d + A m d = e 2 4 π ϵ 0 64 π 4 3 h 2 J + 1 λ ¯ 4 n n 2 + 2 2 9 S e d + n 3 S m d
where A e d and A m d denote the respective radiative transmission probabilities for electric and magnetic dipoles. Equations (13) and (15) are employed to calculate S e d and S m d , respectively.
The relative transmission probability, or branching ratio β r a d J J , can be obtained using the following equation if there are many transitions from the level:
β r a d J J = A r a d J J A r a d J J
where all terminal states J′ are covered by the total.
An energy level’s radiative lifetime τ r a d , which is based on the rates of spontaneous emission across all transitions from this level, can be calculated utilizing the following equation:
τ r a d = 1 A r a d J J
Table 8 presents the electric dipole transmission probability A e d , magnetic dipole transmission probability A m d , spontaneous emission transmission probability A r a d , fluorescence branch ratio β r a d , and radiation lifetime τ r a d of the Er3+ ions in TKNB2 glass, which relate to the transmissions that occur from the higher energy levels of 4I13/2, 4I11/2, 4I9/2, 4F9/2, 4S3/2, 4H11/2, and F 7 / 2 to their lower-lying energy levels, respectively.
In general, the green and red UV emissions were below NIR excitement for the Er3+ and Yb3+-co-doped glass (see Figure 10). According to estimates, the branched ratios of the transitions 2H11/24I15/2 (green), 4S3/24I15/2 (green), and 4F9/24I15/2 (red) were 94.8585%, 69.0467%, and 92.5887%, respectively. These findings indicate that the green and red emission transition channels predominate over all related higher-level radiative transitions.
The ions’ capacity for absorbing or emitting light is measured using the absorption and emission cross-sections. A significant emission cross-section indicates a large gain coefficient and lower pump laser threshold energy. Utilizing the Beer–Lambert equation, the absorbing cross-section ( σ a ) of the 4I13/24I15/2 transmission (1540 nm) of Er3+ was found based on the absorbing patterns of Er3+/Yb3+-co-doped TeO2 glass [41]:
σ a λ = 2.303 × O D ( λ ) N × L
where N is the content of Er3+ ions (ions/cm3), L is the thickness of the sample, and OD(λ) is the optical density of the realistic absorbing patterns of the manufactured glass. The McCumber theory allows for the extraction of the stimulated emission cross-section ( σ e ) utilizing the following relationship [42,43,44]:
σ e λ = σ a λ Z l Z u e x p h c k B T 1 λ Z L 1 λ
The absorption cross-section and partition functions of the lesser and higher states engaged in the optical transmission under consideration are represented by σ a λ , Z l , and Z u . The Planck constant 6.63 × 10 34   J   S , Boltzmann constant ( 1.38 × 10 23 , and temperature (RT in this case) are denoted by the parameters h , k B , and T . Additionally, the wavelength at which the Stark sublevels of the emission are lower and multiple transmissions are achieved is denoted by λ Z L . Figure 11 displays the computed values of the absorbing and emission cross-sections of Er3+/Yb3+-co-doped TeO2 glass (between 1430 and 1630).
The stimulation-emitted cross-section peak, σ e p e a k , is located at 6.86 × 10 21   c m 2 . This value is in good agreement with those of other TeO2 glasses reported in the literature [77,78,79,80]. Moreover, the emitted cross-section σ e p e a k and the full width at half-maximum ( F W H M ) are crucial factors in obtaining high gain and wideband amplification in optical amplifiers. The resulting F W H M × σ e p e a k could be employed to determine an optical amplifier’s bandwidth characteristics. A broader gain bandwidth and lower pump threshold power are implied by the higher values of the F W H M × σ e p e a k produced and the higher radiative lifetime τ r a d . Table 9 lists the F W H M × σ e p e a k values of Er3+/Yb3+-co-doped glasses. It can be seen that the F W H M × σ e p e a k in TKNB2 has a comparable bandwidth characteristic to other glass hosts such as phosphate tellurite glass ( 337.5 × 10 21   n m · c m 2 ) and PBGG ( 338.5 × 10 21   n m · c m 2 ) [35]. Finally, we found that the effect of Er3+/Ho3+ was slightly changed compared with Er3+/Yb3+ in the same host glass; therefore, we only reported the emission cross-section and increased spectroscopic values for a transition energy level of 4I3/24I5/2 of TKNB2.
When assessing the effectiveness of laser mediums, the optical gain coefficient is an essential factor to consider. Once the absorbed and emitted cross-sections for the changes between the two working laser levels are known, the following equation can be applied to determine the optical gain coefficient G ( λ ) [41]:
G λ = N P σ e λ ( 1 P ) σ a λ
where P is the rate of population inversion.
The electron population ratio across the two energy levels is represented by the p value, which progressively increases from 0 to 1, as seen in Figure 12. This figure demonstrates that a positive gain can be achieved when p ≥ 0.6, suggesting that TKNB2 glass may find a use as a matrix substance for 1.54 μm fiber lasers.

4. Conclusions

The thermal stability parameters for the Er2O3/Yb2O3-co-doped glass were slightly lower than those for Er2O3/Ho2O3 due to the substitution of Ho2O3 with Yb2O3 in the glassy matrix. The value of Eopt reduced from 2.689 to 2.663 eV following the substitution of Ho2O3 with Yb2O3, because this substitution increased the amount of NBO in the glass network. The physical parameters (αm, Rm, Ed, Eo, and no) were correlated to the existence of rare earth oxides (Er2O3/Ho2O3 and Er2O3/Yb2O3) in the glassy matrix. The spectroscopic characteristics of the Er3+/Yb3+-co-doped glasses were evaluated based on their intensity factors, radiative rates, branching ratios, and radiative lifetimes. The maximum emission cross-section reported was 6.8 × 10 21   c m 2 , and the gain coefficient of Er3+/Yb3+ for the transition of 4I13/2 4I15/2 was 6.0 cm−1 with a high quality factor ( σ e × F W H M = 377.3 × 10 21   n m · c m 2 ) . When the Ho3+ ions are replaced with Yb3+ ions in the same host glass, it leads to a slight change in the spectroscopic properties, viz, the emission cross-section and gain and intensity parameters of the transition energy level, 4I13/2 4I15/2. These findings suggest that the co-doping of tellurite glass is a viable choice for optical amplifier and laser manufacturing in the communication bands.

Author Contributions

A.B.: conceptualization, methodology, investigation, writing—original draft, formal analysis, and writing—review and editing; A.M.A.-S.: conceptualization, methodology, formal analysis, investigation, and writing—original draft; H.B.A.: conceptualization, methodology, investigation, writing—original draft, and writing—review and editing; K.D.: conceptualization, methodology, investigation, writing—original draft, and writing—review and editing; A.E.E.: conceptualization, methodology, investigation, writing—original draft, formal analysis, and writing—review and editing; M.Ç.E.: investigation, writing—original draft, conceptualization, methodology, investigation, and visualization; E.R.: investigation, writing—original draft, conceptualization, methodology, investigation, and visualization; A.M.A.: methodology and writing—review and editing; K.I.H.: conceptualization, methodology, writing—original draft, visualization, writing—review and editing, and funding acquisition; R.M.: conceptualization, methodology, investigation, and writing—original draft; E.S.Y.: conceptualization, methodology, investigation, visualization, writing—original draft, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education in KSA through project number KKU-IFP2-DA-6.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yuan, J.; Zheng, G.; Ye, Y.; Chen, Y.; Deng, T.; Xiao, P.; Ye, Y.; Wang, W. Enhanced 1.5 μm emission from Yb3+/Er3+ codoped tungsten tellurite glasses for broadband near-infrared optical fiber amplifiers and tunable fiber lasers. RSC Adv. 2021, 1, 27992–27999. [Google Scholar] [CrossRef]
  2. Chillcce, E.F.; Mazali, I.O.; Alves, O.L.; Barbosa, L.C. Optical and physical properties of Er3+-doped oxy-fluoride tellurite glasses. Opt. Mater. 2011, 33, 389–396. [Google Scholar] [CrossRef]
  3. Nandi, P.; Jose, G.; Jayakrishnan, C.; Debbarma, S.; Chalapathi, K.; Alti, K.; Dharmadhikari, A.K.; Dharmadhikari, J.A.; Mathur, D. Femtosecond laser written channel waveguides in tellurite glass. Opt. Express 2006, 14, 12145–12150. [Google Scholar] [CrossRef]
  4. Tong, L.; Hu, L.; Zhang, J.; Qiu, J.; Yang, Q.; Lou, J.; Shen, Y.; He, J.; Ye, Z. Photonic nanowires directly drawn from bulk glasses. Opt. Express 2006, 14, 82–87. [Google Scholar] [CrossRef]
  5. Alazoumi, S.H.; Aziz, A.A.; El-Mallawany, R.; Aliyu, U.S.; Kamari, H.M.; Zaid, M.H.M.; Matori, K.A.; Ushah, A. Optical properties of zinc lead tellurite glasses. Results Phys. 2018, 9, 1371–1376. [Google Scholar] [CrossRef]
  6. Meena, S.L.; Bhatia, B. Polarizability and optical basicity of Er3+ ions doped zinc lithium bismuth borate glasses. J. Pure Appl. Ind. Phys. 2016, 6, 175–183. [Google Scholar]
  7. Thirumaran, S.; Sathish, K. Spectroscopic investigations on structural characterization of borate glass specimen doped with transition metal ions. Res. J. Chem. Environ. 2015, 18, 77–82. [Google Scholar]
  8. Amin Matori, K.; Mohd Zaid, M.H.; Quah, H.J.; Abdul Aziz, S.H.; Abdul Wahab, Z.; Mohd Ghazali, M.S. Studying the effect of ZnO on physical and elastic properties of (ZnO)x(P2O5)1−x glasses using nondestructive ultrasonic method. Adv. Mater. Sci. Eng. 2015, 2015, 6. [Google Scholar] [CrossRef]
  9. Sekhar, K.C.; Ahmed, M.R.; Narsimlu, N.; Deshpande, U.; Sathe, V.G.; Shareefuddin, M. The effect of the addition of CaF2 and PbF2 on boro-tellurite glasses doped with chromium ions. Mater. Res. Express 2020, 6, 125206. [Google Scholar] [CrossRef]
  10. Yousef, E.S.; Mansour, S.F.; Hassaan, M.Y.; Emara, A.M. Synthesis optical properties of novel TeO2 based glasses. Optik 2016, 127, 8933–8939. [Google Scholar] [CrossRef]
  11. Yousef, E.; Hotzel, M.; Russel, C. Linear and non-linear refractive indices of tellurite glasses in the system TeO2–WO3–ZnF2. J. Non-Cryst. Solids 2004, 342, 82–88. [Google Scholar] [CrossRef]
  12. Dai, S.; Wu, J.; Zhang, J.; Wang, G.; Jiang, Z. The spectroscopic properties of Er3+-doped TeO2–Nb2O5 glasses with high mechanical strength performance. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2005, 62, 431–437. [Google Scholar] [CrossRef] [PubMed]
  13. Kaur, A.; Khanna, A.; Aleksandrov, L.I. Structural, thermal, optical and photo-luminescent properties of barium tellurite glasses doped with rare-earth ions. J. Non-Cryst. Solids 2017, 476, 67–74. [Google Scholar] [CrossRef]
  14. Mhareb, M.H.A.; Alajerami, Y.S.M.; Dwaikat, N.; Al-Buriahi, M.S.; Alqahtani, M.; Alshahri, F.; Saleh, N.; Alonizan, N.; Saleh, M.A.; Sayyed, M.I. Investigation of photon, neutron and proton shielding features of H3BO3–ZnO–Na2O–BaO glass system. Nucl. Eng. Technol. 2021, 53, 949–959. [Google Scholar] [CrossRef]
  15. Yin, S.; Wang, H.; Li, A.; Huang, H.; Zhang, J.; Liu, L.; Zhu, Y. Study on the Optical Properties of High Refractive Index TeO2-PbO-ZnO-BaF2 Glass System. Adv. Mater. Sci. Eng. 2021, 2021, 6466344. [Google Scholar] [CrossRef]
  16. Elkhoshkhany, N.; Marzouk, S.Y.; Shahin, S. Synthesis and optical properties of new fluoro-tellurite glass within (TeO2-ZnO-LiF-Nb2O5-NaF) system. J. Non-Cryst. Solids 2017, 472, 39–45. [Google Scholar] [CrossRef]
  17. Kilic, G.; Issever, U.G.; Ilik, E. Synthesis, characterization and crystalline phase studies of TeO2–Ta2O5–ZnO/ZnF2 oxyfluoride semiconducting glasses. J. Non-Cryst. Solids 2020, 527, 119747. [Google Scholar] [CrossRef]
  18. Hraiech, S.; Ferid, M.; Guyot, Y.; Boulon, G. Structural and optical studies of Yb3+, Er3+ and Er3+/Yb3+ co-doped phosphate glasses. J. Rare Earths 2013, 31, 685–693. [Google Scholar] [CrossRef]
  19. Hou, Z.; Xue, Z.; Li, F.; Wang, M.; Hu, X.; Wang, S. Luminescence and up-conversion mechanism of Er3+/Ho3+ co-doped oxyfluoride tellurite glasses and glass–ceramics. J. All. Comp. 2013, 577, 523–527. [Google Scholar] [CrossRef]
  20. Miniscalco, W.J. Erbium-doped glasses for fiber amplifiers at 1550 nm. J. Light. Technol. 1991, 9, 234. [Google Scholar] [CrossRef]
  21. Obaton, A.F.; Parent, C.; Le Flem, G.; Thony, P.; Brenier, A.; Boulon, G. Yb3+- Er3+-codoped LaLiP4O12 glass: A new eye-safe laser at 1535 nm. J. Alloys Compd. 2000, 300–301, 123–130. [Google Scholar] [CrossRef]
  22. Weng, F.Y.; Chen, D.Q.; Wang, Y.S.; Yu, Y.L.; Huang, P.; Lin, H. Energy transfer and up-conversion luminescence in Er3+/Yb3+co- doped transparent glass ceramic containing YF3 nano- crystals. Ceram. Int. 2009, 35, 2619. [Google Scholar] [CrossRef]
  23. Zhang, J.J.; Dai, S.X.; Wang, G.N.; Zhang, L.; Sun, H.T.; Hu, L.L. Investigation on upconversion luminescence in Er3+/Yb3+ codoped tellurite glasses and fibers. Phys. Lett. A. 2005, 345, 409. [Google Scholar] [CrossRef]
  24. Al-Syadi, A.M.; Al-Assiri, M.S.; Hassan, H.M.A.; El Enany, G.; El-Desoky, M.M. Effect of sulfur addition on the electrochemical performance of lithium-vanadium-phosphate glasses as electrodes for energy storage devices. J. Electroanal. Chem. 2017, 804, 36–41. [Google Scholar] [CrossRef]
  25. Krishna, G.; Veeraiah, N.; Venkatramaiah, N.; Venkatesan, R. Induced crystallization and physical properties of Li2O–CaF2–P2O5, TiO2 glass system: Part I. Characterization, spectroscopic and elastic properties. J. Alloys Compd. 2008, 450, 477–485. [Google Scholar] [CrossRef]
  26. Sakka, S.; Mackenzie, J.D. Relation between apparent glass transition temperature and liquids temperature for inorganic glasses. J. Non-Cryst. Solids 1971, 6, 145–162. [Google Scholar] [CrossRef]
  27. Al-Syadi, A.M. Electrochemical performance of Na2O–Li2O–P2S5–V2S5 glass–ceramic nanocomposites as electrodes for supercapacitors. Appl. Phys. A 2021, 127, 755. [Google Scholar] [CrossRef]
  28. Sestak, J. Applicability of DTA to the Study of Crystallization Kinetics of Glasses. Phys. Chem. Glas. 1974, 15, 137. [Google Scholar]
  29. Hruby, A.J. Evaluation of Glass-Forming Tendency by Means of DTA. Czech. Phys. 1972, 22, 1187. [Google Scholar] [CrossRef]
  30. Hussein, K.I.; Al-Syadi, A.M.; Alqahtani, M.S.; Elkhoshkhany, N.; Algarni, H.; Reben, M.; Yousef, E.S.; Stability, T. Thermal stability, optical properties, and Gamma Shielding Properties of Tellurite Glass Modified with Potassium Chloride. Materials 2022, 15, 2403. [Google Scholar] [CrossRef]
  31. Elkhoshkhany, N.; Mohamed, H.M.; Yousef, E.S. UV–Vis-NIR spectroscopy, structural and thermal properties of novel oxyhalide tellurite glasses with composition TeO2-B2O3-SrCl2-LiF-Bi2O3 for optical application. Results Phys. 2019, 13, 102222. [Google Scholar] [CrossRef]
  32. Luo, Y.R. Comprehensive Handbook of Chemical Bond Energies; CRC Press: Boca Raton, FL, USA, 2007. [Google Scholar]
  33. Brik, M.G.; Srivastava, A.M.; Popov, A.I. A few common misconceptions in the interpretation of experimental spectroscopic data. Opt. Mater. 2022, 127, 112276. [Google Scholar] [CrossRef]
  34. El Sayed Yousef, B.; Al-Qaisi, U.V. spectroscopy, refractive indices and elastic properties of the (76 − x) TeO2· 9P2O5· 15ZnO· xLiNbO3 glass. Solid State Sci. 2013, 19, 6–11. [Google Scholar] [CrossRef]
  35. Eevon, C.; Halimah, M.K.; Zakaria, A.; Azurahanim, C.A.C.; Azlan, M.N.; Faznny, M.F. Linear and nonlinear optical properties of Gd3+ doped zinc borotellurite glasses for all-optical switching applications. Results Phys. 2016, 6, 761–766. [Google Scholar] [CrossRef]
  36. Dimitrov, V.; Sakka, S. Electronic oxide polarizability and optical basicity of simple oxides. I. J. Appl. Phys. 1996, 79, 1736–1740. [Google Scholar] [CrossRef]
  37. Jlassi, I.; Elhouichet, H.; Ferid, M. Thermal and optical properties of tellurite glasses doped erbium. J. Mater. Sci. 2011, 46, 806–812. [Google Scholar] [CrossRef]
  38. Saddeek, Y.B.; Yahia, I.S.; Aly, K.A.; Dobrowolski, W. Spectroscopic, mechanical and magnetic characterization of some bismuth borate glasses containing gadolinium ions. Solid State Sci. 2010, 12, 1426–1434. [Google Scholar] [CrossRef]
  39. Sushama, D.; Predeep, P. Thermal and optical studies of rare earth doped tungston-tellurite glasses, International. J. Appl. Phys. Math. 2014, 4, 139. [Google Scholar]
  40. AbuShanab, W.S.; Moustafa, E.B.; Hammad, A.H.; Ramadan, R.M.; Wassel, A.R. Enhancement the structural, optical and non-linear optical properties of cadmium phosphate glasses by nickel ions. J. Mater. Sci. Mater. Electron. 2019, 30, 18058–18064. [Google Scholar] [CrossRef]
  41. Al-Assiri, M.S.; El-Desoky, M.M.; Alyamani, A.; Al-Hajry, A.; Al-Mogeeth, A.; Bahga, A.A. Spectroscopic study of nanocrystalline V2O5· nH2O films doped with Li ions. Opt. Laser Technol. 2010, 42, 994–1003. [Google Scholar] [CrossRef]
  42. El-Mallawany, R.M.; Ribeiro, M.A.; Lara, L.S.; Lenzi, E.K.; Alsadig, I.A.A.; Novatski, A. Refractive index behavior of tellurite glasses. Opt. Mater. 2021, 112, 110810. [Google Scholar] [CrossRef]
  43. Ghosh, G. Sellmeier coefficients and dispersion of thermo-optic coefficients for some optical glasses. Appl. Optic. 1997, 36, 1540–1546. [Google Scholar] [CrossRef] [PubMed]
  44. El-Mallawany, R.M.; Abdalla, D.; Ahmed, I.A. New tellurite glass: Optical properties. Mater. Chem. Phys. 2008, 109, 291–296. [Google Scholar] [CrossRef]
  45. Wemple, S.H.; DiDomenico, M., Jr. Behavior of the electronic dielectric constant in covalent and ionic materials. Phys. Rev. 1971, 3, 1338. [Google Scholar] [CrossRef]
  46. Rao, A.S. Saturation effects in nonlinear absorption, refraction, and frequency conversion: A review. Optik 2022, 267, 169638. [Google Scholar] [CrossRef]
  47. Christodoulides, D.N.; Khoo, I.C.; Salamo, G.J.; Stegeman, G.I.; Van Stryland, E.W. Nonlinear refraction and absorption mechanisms and magnitudes. Adv. Opt. Photonics 2010, 2, 60–200. [Google Scholar] [CrossRef]
  48. Suresh, S.; Ramanand, A.; Jayaraman, D.; Mani, P. Review on theoretical Aspect of nonlinear optics. Rev. Adv. Mater. Sci. 2012, 30, 175–183. [Google Scholar]
  49. 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+, and Tm3+. J. Chem. Phys. 1968, 49, 4424–4442. [Google Scholar] [CrossRef]
  50. Reisfeld, R.; Eckstein, Y. Dependence of spontaneous emission and nonradiative relaxations of Tm3+ and Er3+ on glass host and temperature. J. Chem. Phys. 1975, 63, 4001–4012. [Google Scholar] [CrossRef]
  51. Damak, K.; Al-Shihri, A.S.; Seo, H.J.; Rüssel, C.; Maâlej, R. Quantifying Raman and emission gain coefficients of Ho3+ doped TeO2· ZnO· PbO· PbF2· Na2O (TZPPN) tellurite glass. Solid State Sci. 2014, 28, 74–80. [Google Scholar] [CrossRef]
  52. Yusof, N.N.; Ghoshal, S.K.; Jupri, S.A. Luminescence of neodymium ion-activated magnesium zinc sulfophosphate glass: Role of titanium nanoparticles sensitization. Opt. Mater. 2020, 109, 110390. [Google Scholar] [CrossRef]
  53. Reddy, K.S.R.K.; Swapna, K.; Mahamuda, S.; Venkateswarulu, M.; Rao, A.S. Structural, optical and photoluminescence properties of alkaline-earth boro tellurite glasses doped with trivalent Neodymium for 1.06 μm optoelectronic devices. Opt. Mater. 2021, 111, 110615. [Google Scholar] [CrossRef]
  54. Judd, B.R. The Matrix Elements of Tensor Operators for the Electronic Configurations fn. Proc. Phys. Soc. 1959, 74, 330. [Google Scholar] [CrossRef]
  55. Ofelt, G.S. Intensities of crystal spectra of rare-earth ions. J. Chem. Phys. 1962, 37, 511–520. [Google Scholar] [CrossRef]
  56. Carnall, W.T.; Hessler, J.P.; Wagner, F.W., Jr. Transition probabilities in the absorption and fluorescence spectra of lanthanides in molten lithium nitrate-potassium nitrate eutectic. J. Phys. Chem. 1978, 82, 2152–2158. [Google Scholar] [CrossRef]
  57. Singh, G.; Tiwari, V.S.; Gupta, P.K. Spectroscopic analysis on the basis Judd–Ofelt theory of Nd3+ in (Y0.985Nd0.015)2O3: A transparent laser-host ceramic. Mater. Res. Bull. 2014, 60, 838–842. [Google Scholar] [CrossRef]
  58. Kaminskii, A.; Alexander, A. Laser Crystals: Their Physics and Properties; Springer: Berlin/Heidelberg, Germany, 2013; Volume 14. [Google Scholar]
  59. Damak, K.; Yousef, E.; AlFaify, S.; Rüssel, C.; Maâlej, R. green and infrared emission cross-sectionsof Er3+ doped TZPPN tellurite glass. Opt. Mater. Express 2014, 4, 597–612. [Google Scholar] [CrossRef]
  60. Wang, X.; Fan, S.; Li, K.; Zhang, L.; Wang, S.; Hu, L. Compositional dependence of the 1.8 μm emission properties of Tm3+ ions in silicate glass. J. Appl. Phys. 2012, 112, 103521. [Google Scholar] [CrossRef]
  61. Tanabe, S.; Hanada, T.; Ohyagi, T.; Soga, N. Correlation between Eu 151 Mössbauer isomer shift and Judd-Ofelt Ω 6 parameters of Nd3+ ions in phosphate and silicate laser glasses. Phys. Rev. B 1993, 48, 10591. [Google Scholar] [CrossRef]
  62. Judd, B.R. Optical absorption intensities of rare-earthions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  63. Damak, K.; Maâlej, R.; Qusti, A.H.; Rüssel, C. Thermal and spectroscopic properties of Tm3+ doped TZPPN transparent glass laser material. J. Non-Cryst. Solids 2012, 358, 2974–2980. [Google Scholar] [CrossRef]
  64. León-Luis, S.F.; Rodríguez-Mendoza, U.R.; Martín, I.R.; Lalla, E.; Lavín, V. Effects of Er3+ concentration on thermal sensitivity in optical temperature fluorotellurite glass sensors. Sens. Actuators B Chem. 2013, 176, 1167–1175. [Google Scholar] [CrossRef]
  65. Sui, G.Z.; Li, X.P.; Cheng, L.H.; Zhang, J.S.; Sun, J.S.; Zhong, H.Y.; Tian, Y.; Fu, S.B.; Chen, B.J. Laser cooling with optical temperature sensing in Er3+-doped tellurite-germanate glasses. Appl. Phys. B 2013, 110, 471–476. [Google Scholar] [CrossRef]
  66. Vijaya, N.; Babu, P.; Venkatramu, V.; Jayasankar, C.K.; León-Luis, S.F.; Rodríguez-Mendoza, U.R.; Martín, I.R.; Lavín, V. Optical characterization of Er3+-doped zinc fluorophosphate glasses for optical temperature sensors. Sens. Actuators B Chem. 2013, 186, 156–164. [Google Scholar] [CrossRef]
  67. Feng, L.; Lai, B.; Wang, J.; Du, G.; Su, Q. Spectroscopic properties of Er3+ in a oxyfluoride glass and upconversion and temperature sensor behaviour of Er3+/Yb3+-codoped oxyfluoride glass. J. Lumin. 2010, 130, 2418–2423. [Google Scholar] [CrossRef]
  68. León-Luis, S.F.; Rodríguez-Mendoza, U.R.; Haro-González, P.; Martín, I.R.; Lavín, V. Role of the host matrix on the thermal sensitivity of Er3+ luminescence in optical temperature sensors. Sens. Actuators B Chem. 2012, 174, 176–186. [Google Scholar] [CrossRef]
  69. Rodríguez-Mendoza, U.R.; Lalla, E.A.; Cáceres, J.M.; Rivera-López, F.; León-Luís, S.F.; Lavín, V. Optical characterization, 1.5 μm emission and IR-to-visible energy upconversion in Er3+-doped fluorotellurite glasses. J. Lumin. 2011, 131, 1239–1248. [Google Scholar] [CrossRef]
  70. Buse, G.; Preda, E.; Stef, M.; Nicoara, I. Influence of Yb3+ ions on the optical properties of double-doped Er, Yb: CaF2 crystals. Phys. Scr. 2011, 83, 025604. [Google Scholar] [CrossRef]
  71. Lin, L.; Ren, G.; Chen, M.; Liu, Y. The behavior of Er3+ dopants during crystallization in oxyfluoride silicate glass ceramics. J. Alloys Compd. 2009, 486, 261–264. [Google Scholar] [CrossRef]
  72. Zou, X.; Izumitani, T. Spectroscopic properties and mechanisms of excited state absorption and energy transfer upconversion for Er3+-doped glasses. J. Non-Cryst. Solids 1993, 162, 68–80. [Google Scholar] [CrossRef]
  73. Wang, H.; Qian, G.; Wang, Z.; Wang, M. Spectroscopic properties and Judd–Ofelt theory analysis of erbium chelates. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2005, 62, 146–152. [Google Scholar] [CrossRef] [PubMed]
  74. Labbe, C.; Doualan, J.L.; Girard, S.; Moncorge, R.; Thuau, M. Crystal Growth, Spectroscopic and Laser Investigation of the Er3+ Doped Fluoride Crystals for 2.8 µm Laser Applications. In Proceedings of the LASERS’98, Tucson, Arizona, 7 November 1998; Corcoran, V.J., Goldman, T.A., Eds.; STS Press: McLean, VA, USA, 1999; p. 217. [Google Scholar]
  75. Djellab, S.; Diaf, M.; Labbaci, K.; Guerbous, L. Crystal growth and spectroscopic properties of Er3+ ions doped CdF2 single crystals. Phys. Scr. 2014, 89, 045101. [Google Scholar] [CrossRef]
  76. Khiari, S.; Bendjedaa, F.; Diaf, M. Optical Characterization of Erbium Doped KY3F10 Fluoride Single Crystals. Opt. Phot. J. 2013, 3, 13. [Google Scholar] [CrossRef]
  77. Ma, Y.; Guo, Y.; Huang, F.; Hu, L.; Zhang, J. Spectroscopic properties in Er3+ doped zinc-and tungsten-modified tellurite glasses for 2.7 μm laser materials. J. Lumin. 2014, 147, 372–377. [Google Scholar] [CrossRef]
  78. Liu, R.; Chen, M.; Zhu, X.; Zhou, Y.; Zeng, F.; Su, Z. Luminescent properties and structure of Dy3+ doped germanosilicate glass. J. Lumin. 2020, 226, 117378. [Google Scholar] [CrossRef]
  79. Arunkumar, S.; Marimuthu, K. Spectroscopic properties of Er3+ doped bismuth leadtelluroborate glasses for 1.53 μm optical amplifiers. J. Alloys Compd. 2015, 627, 54–68. [Google Scholar] [CrossRef]
  80. Soga, K.; Tsuda, M.; Sakuragi, S.; Inoue, H.; Inoue, S.; Makishima, A. Effects of chloride introduction on the optical properties and the upconversion emission with 980-nm excitation of Er3+ in ZBLAN fluoride glasses. J. Non-Cryst. Solids 1997, 222, 272–281. [Google Scholar] [CrossRef]
  81. Miniscalco, W.J.; Quimby, R.S. General procedure for the analysis of Er3+ cross sections. Opt. Lett. 1991, 16, 258–260. [Google Scholar] [CrossRef]
  82. Payne, S.A.; Chase, L.L.; Smith, L.K.; Kway, W.L.; Krupke, W.F. Infrared cross-section measurements for crystals doped with Er/sup3+/, Tm/sup3+/, and Ho/sup3+. IEEE J. Quantum Electron. 1992, 28, 2619–2630. [Google Scholar] [CrossRef]
  83. Yang, G.F.; Shi, D.M.; Zhang, Q.Y.; Jiang, Z.H. Spectroscopic Properties of Er3+/Yb3+-codoped PbO–Bi2O3–Ga2O3–GeO2 glasses. J. Fluoresc. 2008, 18, 131–137. [Google Scholar] [CrossRef]
Figure 1. Photo of polished glass samples of 70TeO2-15K2TeO3-10Nb2O5-5BaF2-30,000 ppmHo2O3-30,000 ppm Er2O3 (TKNB1) and 70TeO2-15K2TeO3-10Nb2O5-5BaF2-30,000 ppmYb2O3-30,000 ppm Er2O3 (TKNB2).
Figure 1. Photo of polished glass samples of 70TeO2-15K2TeO3-10Nb2O5-5BaF2-30,000 ppmHo2O3-30,000 ppm Er2O3 (TKNB1) and 70TeO2-15K2TeO3-10Nb2O5-5BaF2-30,000 ppmYb2O3-30,000 ppm Er2O3 (TKNB2).
Materials 17 04188 g001
Figure 2. DSC thermograms for TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Figure 2. DSC thermograms for TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Materials 17 04188 g002
Figure 3. Optical transmission spectra of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Figure 3. Optical transmission spectra of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Materials 17 04188 g003
Figure 4. Optical absorption spectra of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Figure 4. Optical absorption spectra of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Materials 17 04188 g004
Figure 5. Variation in (αhν)1/2 versus for the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Figure 5. Variation in (αhν)1/2 versus for the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Materials 17 04188 g005
Figure 6. Variation in the extinction coefficient (k) with wavelength (λ) of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Figure 6. Variation in the extinction coefficient (k) with wavelength (λ) of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Materials 17 04188 g006
Figure 7. Variation in the refractive index (n) with wavelength (λ) of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Figure 7. Variation in the refractive index (n) with wavelength (λ) of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Materials 17 04188 g007
Figure 8. Variation in refractive index (n) with wavelength (λ) of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Figure 8. Variation in refractive index (n) with wavelength (λ) of the TKNB glass doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Materials 17 04188 g008
Figure 9. RT absorption spectrum of Er3+/Yb3+-co-doped tellurite glass (TKNB2): (a) near-infrared range and (b) UV–visible range.
Figure 9. RT absorption spectrum of Er3+/Yb3+-co-doped tellurite glass (TKNB2): (a) near-infrared range and (b) UV–visible range.
Materials 17 04188 g009
Figure 10. Emission spectra of Er3+/Yb3+-co-doped glass at excitation wavelength of 980 nm.
Figure 10. Emission spectra of Er3+/Yb3+-co-doped glass at excitation wavelength of 980 nm.
Materials 17 04188 g010
Figure 11. The computed absorption and emission cross-sections of Er3+/Yb3+-co-doped TeO2 glass (TKNB2) for the transition of 4I3/24I5/2.
Figure 11. The computed absorption and emission cross-sections of Er3+/Yb3+-co-doped TeO2 glass (TKNB2) for the transition of 4I3/24I5/2.
Materials 17 04188 g011
Figure 12. The gain coefficient for the 4I3/24I5/2 transition of Er3+/Yb3+-co-doped tellurite glass (TKNB2).
Figure 12. The gain coefficient for the 4I3/24I5/2 transition of Er3+/Yb3+-co-doped tellurite glass (TKNB2).
Materials 17 04188 g012
Table 1. The composition, density ρ, and molar volume Vm of the TKNB glassy system co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3.
Table 1. The composition, density ρ, and molar volume Vm of the TKNB glassy system co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3.
Sample CodeComposition
(mol%)
Ρ
(g/cm3) ± 0.001
Vm
(cm3/mol) ± 0.0056
TKNB170TeO2-15K2TeO3-10Nb2O5-5BaF2-30,000 ppm Ho2O3-30,000 ppm Er2O34.958535.373
TKNB270TeO2-15K2TeO3-10Nb2O5-5BaF2-30,000 ppm Yb2O3-30,000 ppm Er2O34.962235.353
Table 2. Thermal parameters of the TKNB glass system co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3.
Table 2. Thermal parameters of the TKNB glass system co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3.
Sample CodeTg
(°C) ± 1.0
Tc
(°C) ± 0.1
Tp
(°C) ± 1.0
T
(°C) ± 1.0
H ± 1.0KSP ± 1.0
TKNB1333422447890.2678.56
TKNB2332419442870.2627.62
Table 3. The Sellmeier coefficients, molar refractivity Rm, molar polarizability αm, metallization criterion M, and optical band gap Eopt of the TKNB glass system co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3.
Table 3. The Sellmeier coefficients, molar refractivity Rm, molar polarizability αm, metallization criterion M, and optical band gap Eopt of the TKNB glass system co-doped with Er2O3/Ho2O3 and Er2O3/Yb2O3.
Sellmeier CoefficientsRm
(Mol−1) ± 0.0465
αm
−3) ± 0.0064
M ± 0.0008Eopt (eV) ± 0.001
Sample CodeABCD
TKNB13.1950.73930.25520.0139717.4496.9210.5072.689
TKNB2 3.205 0.7512 0.2593 0.01695 17.5586.9640.5032.663
Table 4. Dispersion and nonlinear optical parameters of the TKNB glass system co-doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Table 4. Dispersion and nonlinear optical parameters of the TKNB glass system co-doped with Ho2O3/Er2O3 and Yb2O3/Er2O3.
Sample CodeDispersion Parameters Nonlinearity Parameters
Eo (eV)Ed (eV)no ± 0.001χ(1)χ(3) × 10−13
(esu)
n2 × 10−12
(esu)
TKNB18.5925.281.980.234075.19.68
TKNB28.7525.951.990.235965.269.97
Table 5. Reduced matrix element values for Er3+ absorption transitions.
Table 5. Reduced matrix element values for Er3+ absorption transitions.
Transition from 4I15/2 to U ( 2 ) 2 U ( 4 ) 2 U ( 6 ) 2
4I13/20.01940.11731.4332
4I9/20.00.16940.0101
4F9/20.00.5480.4769
4S3/20.00.00.2369
2H11/20.81200.46690.0951
4F7/20.00.14470.6348
Table 6. Values of average wavelengths, refractive index, and integrated absorption coefficients’ electric and magnetic dipole line strengths for Er3+/Yb3+-co-doped tellurite glass (TNBK2).
Table 6. Values of average wavelengths, refractive index, and integrated absorption coefficients’ electric and magnetic dipole line strengths for Er3+/Yb3+-co-doped tellurite glass (TNBK2).
Transition from 4I15/2 to λ ¯
( n m )

n
O D λ d λ
( n m )
S e d c a l
( p m 2 )
S e d m e s
( p m 2 )
S m d
( p m 2 )
4I13/215301.997517.884141.20871.18550.7148
4I9/28002.01551.067150.32540.29380
4F9/26502.02864.023071.34461.34830
4S3/25452.04800.468980.15580.18450
2H11/25202.05547.030412.87962.88010
4F7/24902.06691.704830.69490.73410
δ r m s = 0.0361 × 10 20   c m 2
Table 7. Judd–Ofelt parameters and spectroscopic quality factor ( Q = Ω 4 / Ω 6 ) in different host materials, along with reported Er3+-doped systems.
Table 7. Judd–Ofelt parameters and spectroscopic quality factor ( Q = Ω 4 / Ω 6 ) in different host materials, along with reported Er3+-doped systems.
Glass Host Ω 2 × 10 20   c m 2 Ω 4 × 10 20   c m 2 Ω 6 × 10 20   c m 2 Q = Ω 4 / Ω 6 Reference
Er3+/Yb3+2.3871.8810.6572.8630[Present work]
ZBLAN2.731.401.101.2727[77,80]
Boro-tellurite 4.2320.7790.6121.2729[78,81]
Tellurite5.62.110.732.8904[79,82]
TLNT6.481.821.271.4331[80,83]
Table 8. Calculated radiative parameters of the different states of Er3+ ion-doped tellurite glass (TKNB2).
Table 8. Calculated radiative parameters of the different states of Er3+ ion-doped tellurite glass (TKNB2).
Transitions ν ¯ ( c m 1 ) n S e d   ( p m 2 ) S m d   ( p m 2 ) A e d   ( s 1 ) A m d   ( s 1 ) β ( % ) = τ r ( m s )
4I13/24I15/265361.9981.20790.700027780.4906100.002.7956
4I11/24I15/210,2562.0070.32050339089.6172.6416
4I11/24I13/237201.9950.40400.79102019.485510.383
4I9/24I15/212,5002.0160.32530762085.5481.1221
4I9/24I13/259641.9970.50630123013.838
4I9/24I11/222441.995 0.24900.348032.25760.6143
4F9/24I15/215,3852.0291.34410605.5092.5890.1529
4F9/24I13/288492.0030.3787030604.6771
4F9/24I11/251291.9961.00180.172015513.34202.5720
4F9/24I9/228851.9950.33030.112091.54410.1623
4S3/24I15/218,3492.0480.15560310.9069.0470.2221
4S3/24I13/211,8132.0130.22980113.0025.086
4S3/24I11/280932.0010.064609902.2064
4S3/24I9/258491.9960.2848016403.6314
4S3/24F9/229641.9950.01810100.0299
2H3/24I15/219,2312.0552.879002245.2094.8590.0423
2H3/24I13/212,6952.0160.24730.1160507117.27632.6367
2H3/24I11/289752.0030.36950.086026030.13531.2244
2H3/24I9/267311.9980.81820.03002394.39801.0303
2H11/24F9/238461.9951.08190.0090590.24500.2489
2H11/24S3/28821.9950.42190000.0012
4F7/24I15/220,4082.0670.69450996.8071.6270.0719
4F7/24I13/213,8722.0210.62940255.4018.355
4F7/24I11/210,1522.00700.6296096906.9622
4F7/24I9/279082.00050.51050.05136618.25762.7589
4F7/24F9/250231.99550.11310.2122119.30500.2862
4F7/24S3/220591.99470.00870000.0008
4F7/24H11/211771.99470.62070100.0104
Table 9. Obtained values of F W H M   ( n m ) , σ e   ( c m 2 ) 1.54 μm, and σ e × F W H M   ( n m · c m 2 ) at 1.53 μm.
Table 9. Obtained values of F W H M   ( n m ) , σ e   ( c m 2 ) 1.54 μm, and σ e × F W H M   ( n m · c m 2 ) at 1.53 μm.
F W H M   ( n m ) σ e   ( c m 2 ) σ e × F W H M   ( n m · c m 2 )
Present work
Er3+/Yb3+
556.86 ×   10 21   c m 2 377.3 × 10 21   n m · c m 2
PBGG [35]388.9 338.5 × 10 21   n m · c m 2
Phosphate tellurite glass [36]44.8427.7 337.5 × 10 20   n m · c m 2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Boussetta, A.; Al-Syadi, A.M.; Albargi, H.B.; Damak, K.; Ersundu, A.E.; Ersundu, M.Ç.; Ramadan, E.; Alshehri, A.M.; Hussein, K.I.; Maalej, R.; et al. An Investigation of the Photonic Application of TeO2-K2TeO3-Nb2O5-BaF2 Glass Co-Doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 at 1.54 μm Based on Its Thermal and Luminescence Properties. Materials 2024, 17, 4188. https://doi.org/10.3390/ma17174188

AMA Style

Boussetta A, Al-Syadi AM, Albargi HB, Damak K, Ersundu AE, Ersundu MÇ, Ramadan E, Alshehri AM, Hussein KI, Maalej R, et al. An Investigation of the Photonic Application of TeO2-K2TeO3-Nb2O5-BaF2 Glass Co-Doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 at 1.54 μm Based on Its Thermal and Luminescence Properties. Materials. 2024; 17(17):4188. https://doi.org/10.3390/ma17174188

Chicago/Turabian Style

Boussetta, Ahlem, Aref M. Al-Syadi, Hasan B. Albargi, Kamel Damak, Ali Erçin Ersundu, Miray Çelikbilek Ersundu, Essam Ramadan, Ali M. Alshehri, Khalid I. Hussein, Ramzi Maalej, and et al. 2024. "An Investigation of the Photonic Application of TeO2-K2TeO3-Nb2O5-BaF2 Glass Co-Doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 at 1.54 μm Based on Its Thermal and Luminescence Properties" Materials 17, no. 17: 4188. https://doi.org/10.3390/ma17174188

APA Style

Boussetta, A., Al-Syadi, A. M., Albargi, H. B., Damak, K., Ersundu, A. E., Ersundu, M. Ç., Ramadan, E., Alshehri, A. M., Hussein, K. I., Maalej, R., & Yousef, E. S. (2024). An Investigation of the Photonic Application of TeO2-K2TeO3-Nb2O5-BaF2 Glass Co-Doped with Er2O3/Ho2O3 and Er2O3/Yb2O3 at 1.54 μm Based on Its Thermal and Luminescence Properties. Materials, 17(17), 4188. https://doi.org/10.3390/ma17174188

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