1 Introduction

Alkali/alkaline earth borate glasses doped with transition metal elements, rare earth elements or both attracted the interest of researchers due to the importance of these glasses in developing many technological applications such as solid-state lasers, solar energy converters, optical fibers, and electronic devices [1,2,3]. Barium borate glasses have acquired great importance due to their unique optical properties, low melting point, high thermal and mechanical stability, and good solubility of rare earth ions as well as transition metals [4,5,6]. The fundamental structural units of borate-based glasses are trigonal BO3 and tetrahedral BO4 which can combine to form different structural groups such as diborates, tetraborates, orthoborates, pentaborates, metaborates, and boroxol rings. The coordination number of the boron atom can be easily changed from three to four or vice versa depending on the type and the concentration of the added modifier [7,8,9,10].

The introduction of transition metals into glasses improves the structural, optical, electrical, and magnetic properties of them due to the multiplicity of valance states of these elements, which depend mainly on the glass composition and melting conditions [8, 11, 12]. Among transition metal oxides, TiO2 is a useful glass constituent for preparation some special optical glasses and glass-ceramics. Titanium ions usually exist in the glassy host in trivalent and tetravalent valence states and participate in the glassy matrix with TiO4, TiO6, and sometimes TiO5 structural groups [13].

Barium titanate-based glasses are promising materials due to their unique and distinctive structural, optical, and electrical properties which make them good candidates in many modern applications such as sensors and transducers, electronics, bioactive materials, non-linear optical devices, and reversible electrodes [14, 15]. The incorporation of barium titanate with any glass former helps to form a homogeneous glass matrix and enhances the radiation shielding properties of this matrix [16]. Glasses containing rare earth elements as luminescent centers are considered as an important class of photonic materials due to their unique and distinct spectroscopic properties resulting from f–f transitions in the visible and near to mid IR spectral ranges [17, 18]. The luminescent properties of these glasses are mainly dependent on the glass type, composition, and the ratio of the rare earth elements [19].

The objective of this work is to investigate the impact of Yb3+ ions on the structural and optical properties of barium titanium borate glasses as optical materials.

2 Experimental details

2.1 Glass fabrication

Pure chemicals of boric acid (H3BO3), titanium dioxide (TiO2), barium carbonate (BaCO3), and ytterbium oxide (Yb2O3) were used as starting materials to synthesize Yb3+-doped barium titanium borate glasses by the standard melting-quenching method. Appropriate molar ratios of the starting materials were weighed, thoroughly mixed, and melted in a porcelain crucible using an electric furnace (VECSTAR, UK) at 1100 ± 10 °C for 2 h. The molten materials were rotated and checked every half hour to obtain a high degree of homogeneity, then casted into slightly warm stainless-steel molds and rapidly moved to a muffle furnace for annealing at 330 ± 5 °C for half hour. The molar composition of the synthesized glasses and their codes are presented in Table 1.

Table 1 Compositions of barium titanium borate glasses doped with Yb2O3 in mol% and their codes

2.2 Density and molar volume measurements

Archimedes method was utilized for measuring the density (ρ) of the bulk glasses by weighing them in both of air and xylene (0.86 g/cm3) based on the next relation:

$$\rho = \frac{{W_{{\text{a}}} \times \rho_{{\text{x}}} }}{{W_{{\text{a}}} - W_{{\text{x}}} }} \frac{{\text{g}}}{{{\text{cm}}^{3} }},$$
(1)

where Wa is the glass weight in air, Wx is the glass weight in xylene, and ρx is the xylene density. The molar volume (Vm) of the glasses is estimated using the relation Vm = Mwt/ρ, where Mwt is the glass molecular weight.

2.3 Spectroscopic measurements

Raman spectra were recorded in the wavenumber range 40–3750 cm−1 using SENTERRA Raman microscope (Bruker, Germany) with a laser excitation wavelength of 785 nm and power of 50 mW. The FTIR absorption spectra were recorded using computerized FTIR spectrometer (Bruker, Vertex 70v, Germany) covers the wavenumber range 4000–400 cm−1. The UV–vis–NIR absorption spectra were recorded using a single monochromatic spectrophotometer (JASCO, V-770, Japan) covers the wavelength range 2500–190 nm. XRD spectra were recorded using a diffractometer (Philips, PW-1390) in 2Θ range from 5° to 70°.

3 Results and discussion

3.1 XRD spectra

Figure 1 shows the XRD spectra of the studied glasses in the range of 5°–70°. The non-presence of a sharp peak in the XRD pattern confirms the amorphous or non-crystalline phase of the synthesized glasses [20].

Fig. 1
figure 1

XRD spectra of Yb3+-doped barium titanium borate glasses

3.2 Density and molar volume

The dependance of the density (ρ) and molar volume (Vm) of the studied glasses on Yb2O3 content is shown in Fig. 2. The addition of Yb2O3 with ascending ratios increases the density of the glasses from 3.652 to 3.761 g/cm3 due to the replacement of the lighter mass content B2O3, TiO2, and BaO by the heavier mass content Yb2O3 [4]. Moreover, replacing the oxides of boron, titanium, and barium with ytterbium oxide increases the number of oxygen ions within the glass matrix, which promotes the conversion of triangular BO3 units to tetrahedral BO4 units, and because the density of BO4 units is greater than that of BO3 units, this conversion promotes an increase in the density of the glass [21, 22]. The molar volume is found to decrease from 27.875 to 27.784 cm3/mol then increase to 28.06 cm3/mol due to the difference between ionic radii of Yb (0.985 Å) and other cations (Ba 1.35 Å, Ti 0.74 Å, B 0.23 Å) within the glass matrix. Thus, both the rigidity and compactness of the glasses diminishes as a result of the creation of more non-bridging oxygens [21]. The obtained values of the density and molar volume are presented in Table 2.

Fig. 2
figure 2

Density and molar volume as a function of Yb2O3 content

Table 2 Density (ρ), molar volume (Vm), ytterbium ions concentration (N), mean ytterbium ions separation (R), oxygen packing density (OPD), and mean boron–boron separation (<dB–B>) of Yb2O3-doped barium titanium borate glasses

The obtained values of the density and molar volume are utilized to calculate some physical parameters such as Yb3+ ions concentration (N), Yb3+ ions separation (R), oxygen packing density (OPD), and mean boron–boron separation (<dB–B>) based on the next relationships [7, 23, 24]:

$$N=2\left(\frac{\rho {x}_{{\mathrm{Yb}}_{2}{\mathrm{O}}_{3}}}{Mwt}\right){N}_{\mathrm{A}}$$
(2)
$$R = \sqrt[3]{\frac{1}{N}}$$
(3)
$$\mathrm{OPD}= \frac{z \rho }{Mwt} \times 1000$$
(4)
$$< d_{{{\text{B}}{-}{\text{B}}}} > \, = \left( {\frac{{V_{m} }}{{2N_{{\text{A}}} \left( {1 - x_{{{\text{B}}_{2} {\text{O}}_{3} }} } \right)}}} \right)^{1/3} ,$$
(5)

where xYb2O3 is the molar fraction of Yb2O3, NA is Avogadro’s number, z is the number of oxygen atoms in the glass composition, and xB2O3 is the molar fraction of B2O3. The N values increase as the concentration of Yb2O3 increases while the R values decrease. The OPD values increase until 0.769 mol% of Yb2O3 then decrease while the <dB–B> values have opposite trend for the same concentrations. The observed increase in the OPD and the observed decrease in the <dB–B> are the result of the compactness of the glass matrix with increasing the ratio of Yb2O3. The values of N, R, OPD, and <dB–B> are presented in Table 2.

3.3 Raman spectra

The Gaussian deconvolution of the Raman spectra for the studied glasses are shown in Fig. 3. The deconvoluted spectrum of the base undoped glass (BTiBaYb0) reveals eleven bands centered at 77 and 223 cm−1 due to vibrations of Ba2+ ions and Ti–O bonds, respectively [25, 26], 418 and 1140 cm−1 due to vibrations of diborate groups [27, 28], 540 and 901 cm−1 due to bending and stretching vibrations of orthoborate units, respectively [29, 30], 794 cm−1 due to vibration of planar six membered borate rings containing one BO4 tetrahedron [31], 1007 cm−1 due to asymmetric stretching modes of tetrahedral borate units [5], 1281 cm−1 due to vibration of pyroborate groups [31], 1413 cm−1 due to vibration of non-bridging oxygens from various B–O structural units [32], and 1488 cm−1 due to vibration of BO2O triangles connected with other borate triangular units [30]. The addition of Yb2O3 into host glass matrix with different ratios leads to the following modifications: (i) The formation of new bands in the range 707–736 cm−1 and 1502–1666 cm−1 due to the vibrations of B–O bonds in BO4 units and metaborate rings, respectively [28, 29, 33,34,35]. (ii) The absence of the band at 1281 cm−1 in the BTiBaYb1 and BTiBaYb3 glasses shows the non-presence of pyroborate groups in these glasses [31].

Fig. 3
figure 3

Deconvoluted Raman spectra of Yb3+-doped barium titanium borate glasses

Sekhar et al. [36] studied the deconvoluted Raman spectra of the Na2B4O7–CdO–CuO–BaTiO3 glasses which revealed distinctive bands at 733–768 cm−1 due to the vibrations of B–O–B linkages in the six membered borate rings with BO4 tetrahedral, 842–877 cm−1 due to the vibration of orthoborate units and 1180–1375 cm−1 due to the stretching vibration of B–O with non-bridging oxygens. The present work is in line with the results reported by Sekhar et al.

3.4 FTIR absorption spectra

The FTIR spectra of the studied glasses are shown in Fig. 4. The spectrum of the base undoped glass (BTiBaYb0) reveals a small band in the region 458–501 cm−1 with two peaks at 469 and 483 cm−1 due to vibration of barium ions [37]. Furthermore, a broad band in the region 545–760 cm−1 with its peak at 625 cm−1 is ascribed to vibration of titanium–oxygen bond in TiO6 groups [38]. Moreover, a broad band in the region 776–1167 cm−1 is attributed to stretching vibration of BO4 groups [39]. In addition, a broad band in the region 1195–1569 cm−1 with its center at 1384 cm−1 is attributed to stretching vibration of boron–oxygen bond in BO3 groups [40]. A medium band in the region 1573–1728 cm−1 with two peaks at 1618 and 1638 cm−1 is ascribed to vibration of water and OH groups [39, 41]. The incorporation of Yb2O3 into glass matrix with different ratios leads to the following modifications: (i) The intensity of the first band (458–501 cm−1) decreases with increasing the concentration of Yb3+ ions. (ii) The second band (545–760 cm−1) becomes more broadened and splits into two bands centred at 572 and 721 cm−1 due to the bending modes of various borate groups and the vibrations of TiO4 groups, respectively [38, 41]. (iii) The intensity of the third band (776–1167 cm−1) increases with increasing the concentration of Yb3+ ions. (iv) The fourth band (1195–1569 cm−1) becomes more broadened and splits into two bands with their peaks at 1207 and 1384 cm−1 due to the stretching vibrations of BO4 and BO3 groups, respectively, in the BTiBaYb1 and BTiBaYb5 glasses [39, 40]. (v) The intensity of the fifth band (1573–1728 cm−1) decreases with increasing the concentration of Yb3+ ions. The obtained analysis from FTIR spectra are consistent and in agreement with that of Raman spectra.

Fig. 4
figure 4

FTIR absorption spectra of Yb3+-doped barium titanium borate glasses

3.5 Optical absorption spectra

The optical absorption spectra of the studied glasses are displayed in Fig. 5. The base undoped glass (BTiBaYb0) exhibits an absorption band in the UV region due to the existence of inevitable iron impurities within the raw materials [42]. The addition of Yb2O3 with different ratios leads to the formation of distinctive broad absorption band in the NIR region at 977 nm due to the electronic transition from the 2F7/2 ground state to the 2F5/2 higher excited state [43]. The intensity of this band increases as the ratio of Yb2O3 increases.

Fig. 5
figure 5

UV–vis–NIR absorption spectra of Yb3+-doped barium titanium borate glasses

3.6 Optical bandgap and Urbach energy

Moot and Davis theory is utilized to calculate the optical bandgap energy (Eg) from the optical absorption edge. The absorption coefficient (α) relevant to the absorption edge is determined using the relationship α = 2.303 A/d, where A is the optical absorption and d is the glass thickness. The relationship between α and Eg is given by [44]:

$$\alpha h\nu =B{(h\nu -{E}_{\mathrm{g}})}^{r},$$
(6)

where is the incident photon energy, B is constant, and r is the electronic transition index which holds the value of 2 for indirect allowed transition. Figure 6 shows Tauc’s plots, i.e., (αhυ)1/2 versus for all glasses. The values of Eg are determined by extrapolating the linear parts of the curves to intercept the -axis at zero absorption coefficient. The values of Eg are recorded in Table 3 and found to be 3.198, 3.168, 3.140, and 3.115 eV for BTiBaYb0, BTiBaYb1, BTiBaYb3, and BTiBaYb5 glasses, respectively. The observed decrease in the values of Eg is ascribed to the excess of the non-bridging oxygens inside the glass network with increasing the ratio of Yb2O3 [45, 46].

Fig. 6
figure 6

(αhυ)1/2 vs. for Yb3+-doped barium titanium borate glasses

Table 3 Optical bandgap (Eg & \({E}_{\mathrm{g}}^{\mathrm{ASF}}\)), average optical bandgap (\({E}_{\mathrm{g}}^{\mathrm{avg}}\)), Urbach energy (∆E), and linear refractive index (\({n}_{\mathrm{D}-\mathrm{S}}\)) of Yb2O3 doped barium titanium borate glasses

Also, the absorption spectrum fitting (ASF) method is utilized to calculate the optical bandgap energy (\({E}_{\mathrm{g}}^{\mathrm{ASF}}\)). According to the ASF method, Eq. (6) can be written in terms of optical absorbance (A) and photon wavelength (λ) as follow [47, 48]:

$$A=D{\lambda ({\lambda }^{-1}-{\lambda }_{\mathrm{g}}^{-1})}^{r},$$
(7)

where \(D=[B{d\left(hc\right)}^{r-1}/2.303]\), d is the glass thickness, h is Planck’s constant, c is the light speed, and λg is the photon wavelength corresponding to the \({E}_{\mathrm{g}}^{\mathrm{ASF}}\). Figure 7 shows the plots of (A/λ)1/2 versus λ−1 for all glasses. The values of \({\lambda }_{\mathrm{g}}^{-1}\) are determined by extrapolating the linear parts of the curves to intercept the λ−1-axis at zero absorbance. The values of \({E}_{\mathrm{g}}^{\mathrm{ASF}}\) are calculated from Eq. (8) and recorded in Table 3.

Fig. 7
figure 7

(A/λ)1/2 vs. (λ−1) for Yb3+-doped barium titanium borate glasses

$${E}_{\mathrm{g}}^{\mathrm{ASF}}=1239.83\times {\lambda }_{\mathrm{g}}^{-1}$$
(8)

The values of \({E}_{\mathrm{g}}^{\mathrm{ASF}}\) are found to be 3.194, 3.170, 3.137, and 3.113 eV for BTiBaYb0, BTiBaYb1, BTiBaYb3, and BTiBaYb5 glasses, respectively. From Table 3, it can be seen the values of Eg and \({E}_{\mathrm{g}}^{\mathrm{ASF}}\) are matched and in agreement with each other which enhances the formation of non-bridging oxygens. The average values of the optical bandgaps \({(E}_{\mathrm{g}}^{\mathrm{avg}}={E}_{\mathrm{g}}+ {E}_{\mathrm{g}}^{\mathrm{ASF}}/2)\) are calculated and recorded in Table 3.

Urbach energy (∆E) is a measure of the disorders inside the glassy matrix and related to the logarithm of the absorption coefficient through the relationship [49]:

$$\mathrm{Ln}(\alpha )=\mathrm{constant}+\frac{h\upsilon }{\Delta E}$$
(9)

The values of ∆E are calculated by obtaining the reciprocal of the slope of the linear parts of ln (α) versus plots and are recorded in Table 3. The plots of ln (α) vs. for all glasses are presented in Fig. 8. The values of ∆E are found to be 0.210, 0.279, 0.228, and 0.255 eV for for BTiBaYb0, BTiBaYb1, BTiBaYb3, and BTiBaYb5 glasses, respectively. The BTiBaYb1 glass has the highest value of ∆E, and thus the highest degree of disorders. The BTiBaYb0 glass has the lowest value of ∆E, and thus the lowest degree of disorders which means more homogeneity and stable structure with minimal defects [49, 50].

Fig. 8
figure 8

ln (α) vs. for Yb3+-doped barium titanium borate glasses

3.7 Refractive index and dispersion parameters

The average values of the optical bandgaps \({(E}_{\mathrm{g}}^{\mathrm{avg}})\) are employed to calculate the corresponding linear refractive indices (\({n}_{\mathrm{D}-\mathrm{S}}\)) at high frequencies using Dimitrov–Sakka relationship [45]:

$${n}_{\mathrm{D}-\mathrm{S}}={\left(6\sqrt{\frac{5}{{E}_{\mathrm{g}}^{\mathrm{avg}}}}-2\right)}^{1/2}$$
(10)

The values of \({n}_{\mathrm{D}-\mathrm{S}}\) are recorded in Table 3 and observed to increase with increasing the ratio of Yb3+ ions because of the increase in polarization within the glass matrix at high frequencies as mentioned in the next section.

Wemple–DiDomenico (WDD) model is used to calculate the static refractive index (no) at low frequencies, i.e., the refractive index at zero frequency \((h\upsilon \cong 0)\). The mathematical equation that expresses WDD model is as follows [51]:

$${\left({n}^{2}-1\right)}^{-1}=\frac{{E}_{o}}{{E}_{d}}-\frac{1}{{E}_{d}{E}_{o}}{\left(h\upsilon \right)}^{2},$$
(11)

where n is the refractive index, Eo is the single oscillator energy, Ed is the dispersion energy, and is the incident photon energy. The values of Eo and Ed are determined by graphing \({\left({n}^{2}-1\right)}^{-1}\) versus \({\left(h\upsilon \right)}^{2}\) as shown in Fig. 9. The slope of the curves equals \((-1/E\)d \(E\)o\()\) and the intercept equals \((E\)o\(/E\)d\()\). The values of Eo and Ed are recorded in Table 4 and found to decrease as the ratio of Yb3+ ions increases. The observed decrease in Ed values indicates the increase in the degree of structural disorder of the glasses as the ratio of Yb3+ ions increase which is in line with the Urbach energy results [52]. The observed decrease in Eo values is in line with the optical bandgap results. When tends to zero, Eq. (11) can be written in the next form:

Fig. 9
figure 9

\({\left({n}^{2}-1\right)}^{-1}\) versus \({\left(h\upsilon \right)}^{2}\) for Yb3+-doped barium titanium borate glasses

Table 4 Single oscillator energy (Eo), dispersion energy (Ed), static refractive index (no), static dielectric constant (ε), and optical spectra moments (\({M}_{-1}\& {M}_{-3}\)) of Yb2O3 doped barium titanium borate glasses
$${n}_{o}=\sqrt{1+\frac{{E}_{d}}{{E}_{o}}}$$
(12)

The value of no depends on the values of Ed and Eo. Since the decrease in the values of Ed is greater than the decrease in the values of Eo, the static refractive index decreases accordingly. The values of no are recorded in Table 4. The values of Eo and Ed are employed to calculate the optical spectrum moments \(({M}_{-1}\&{ M}_{-3})\) and static dielectric constant \(({\varepsilon }_{\infty })\) from the next relations [51]:

$${M}_{-1}=\frac{{E}_{d}}{{E}_{o}} \& {M}_{-3}=\frac{{E}_{d}}{{E}_{o}^{3}}$$
(13)
$${\varepsilon }_{\infty }={n}_{o}^{2}=1+\frac{{E}_{d}}{{E}_{o}}$$
(14)

The values of \({M}_{-1},{ M}_{-3}, \mathrm{and} {\varepsilon }_{\infty }\) are recorded in Table 4. The values of \({M}_{-1}\mathrm{ and} {\varepsilon }_{\infty }\) are found to decrease with increasing the concentration of Yb3+ ions while the values of \({M}_{-3}\) increase.

The average oscillator strength (So) and average oscillator wavelength (λo) are calculated using Sellmeier dispersion formula [53]:

$${n}^{2}-1=\frac{{S}_{o}{\lambda }_{o}^{2}}{1-\left(\frac{{\lambda }_{o}^{2}}{{\lambda }^{2}}\right)}$$
(15)

Figure 10 shows the graphs of \({\left({n}^{2}-1\right)}^{-1}\) versus \({\lambda }^{-2}\) for all glasses. The slope of the curves equals \((-1/S\)o\()\) and the intercept equals\((1/{{S}_{o} \lambda }_{o}^{2})\). The values of So are found to decrease with increasing the ratio of Yb3+ ions while the values of λo increase. The values of So and λo are recorded in Table 5.

Fig. 10
figure 10

\({\left({n}^{2}-1\right)}^{-1}\) versus \({\lambda }^{-2}\) for Yb3+-doped barium titanium borate glasses

Table 5 Average oscillator strength (So), average oscillator wavelength (λo), linear optical susceptibility (\({\chi }^{(1)}\)), third order non-linear optical susceptibility (\({\chi }^{(3)}\)), and non-linear refractive index (n2) of Yb2O3-doped barium titanium borate glasses

The linear optical susceptibility \({\chi }^{(1)}\), third order non-linear optical susceptibility \({\chi }^{(3)}\), and non-linear refractive index n2 are calculated from the relationships [53]:

$${\chi }^{(1)}=\frac{1}{4\pi }\frac{{E}_{d}}{{E}_{o}}$$
(16)
$${\chi }^{(3)}=6.82\times {10}^{-15}{\left(\frac{{E}_{d}}{{E}_{o}}\right)}^{4} (\mathrm{esu})$$
(17)
$${n}_{2}=12\pi \frac{{\chi }^{(3)}}{{n}_{o}}$$
(18)

The values of \({\chi }^{(1)}\), \({\chi }^{(3)}\), and n2 are recorded in Table 5 and found to decrease with increasing the concentration of Yb3+ ions due to the great decrease in the values of Ed compared to that of Eo. Figure 11 shows the variations of linear, static, and non-linear refractive indices as a function of Yb2O3 content. The observed linear behavior of \({n}_{\mathrm{D}-\mathrm{S}}\) in Fig. 11 is represented by an empirical equation as follows:

$${n}_{\mathrm{D}-\mathrm{S}}=0.01603 {x}_{{\mathrm{Yb}}_{2}{\mathrm{O}}_{3}}+2.34752$$
(19)

where xYb2O3 is the molar fraction of Yb2O3. The linear refractive index of the studied glasses strongly depends on the concentration of Yb3+ ions. Figure 12 shows the variations of third order non-linear optical susceptibility and static dielectric constant as a function of of Yb2O3 content.

Fig. 11
figure 11

Linear, static, and non-linear refractive indices as a function of Yb2O3 content

Fig. 12
figure 12

Third order non-linear optical susceptibility and static dielectric constant as a function of of Yb2O3 content

3.8 Metallization criterion and optical basicity

Lorentz–Lorentz equation correlates the molar refraction (Rm) of the studied glasses with the linear refractive index (\({n}_{\mathrm{D}-\mathrm{S}}\)) and molar volume (Vm) as follows [54]:

$${R}_{m}=\left(\frac{{n}_{\mathrm{D}-\mathrm{S}}^{2}-1}{{n}_{\mathrm{D}-\mathrm{S}}^{2}+2}\right){V}_{m}$$
(20)

The molar electronic polarizability (αm) is proportional to the molar refraction (Rm) and given by:

$${\alpha }_{m}=\left(\frac{3}{4\pi {N}_{\mathrm{A}}}\right){R}_{m}$$
(21)

The values of Rm and αm are calculated and listed in Table 6. As seen from Table 6, the Rm and αm tend to increase with increasing the ratio of Yb3+ ions. The metallization criterion (MC) is utilized to obtain information about the non-metallic behavior of solids based on the relation [45]:

$${M}_{\mathrm{C}}=1-\frac{{R}_{m}}{{V}_{m}}$$
(22)

when the value of MC approaches one, the material is classified as an insulator while when the value of MC approaches zero, the material classified as a metal. The values of MC of the studied glasses are calculated and found to decrease from 0.400 to 0.394 with increasing the concentration of Yb3+ ions. The values of MC (0.394–0.400) lie in between insulators and metals, i.e., the studied glasses behave as semiconductors. The dependance of the molar electronic polarizability and metallization criterion on the Yb2O3 content is presented in Fig. 13.

Table 6 Molar refraction Rm, molar electronic polarizability αm, metallization criterion MC, oxide ion polarizability \({\alpha }_{{\mathrm{O}}^{2-}}\), and optical basicity \(\Lambda\) of Yb2O3-doped barium titanium borate glasses
Fig. 13
figure 13

Molar electronic polarizability and metallization criterion as a function of Yb2O3 content

The oxide ion polarizability \({\alpha }_{{\mathrm{O}}^{2-}}\left({n}_{\mathrm{D}-\mathrm{S}}\right)\) and optical basicity \(\Lambda ({n}_{\mathrm{D}-\mathrm{S}})\) of the studied glasses can be calculated based on the linear refractive index using the next relations [55]:

$${\alpha }_{{\mathrm{O}}^{2-}}\left({n}_{\mathrm{D}-\mathrm{S}}\right)=\left[ \frac{{R}_{m} }{2.52} - \sum {\alpha }_{\mathrm{cat} }\right]{\left({N}_{{\mathrm{O}}^{2-}}\right)}^{-1}$$
(23)
$$\Lambda \left({n}_{\mathrm{D}-\mathrm{S}}\right)=1.67 \left(1- \frac{1}{{\alpha }_{{\mathrm{O}}^{2-}}}\right),$$
(24)

where \(\sum {\alpha }_{\mathrm{cat}}\) is the molar cation polarizability and \({N}_{{\mathrm{O}}^{2-}}\) is the number of oxide ions in the chemical formula of the glass composition. The dependance of the \({\alpha }_{{\mathrm{O}}^{2-}}\left({n}_{\mathrm{D}-\mathrm{S}}\right)\) and \(\Lambda \left({n}_{\mathrm{D}-\mathrm{S}}\right)\) on Yb2O3 content is presented in Fig. 14. As seen from Fig. 14, the \({\alpha }_{{\mathrm{O}}^{2-}}\left({n}_{\mathrm{D}-\mathrm{S}}\right)\) and \(\Lambda \left({n}_{\mathrm{D}-\mathrm{S}}\right)\) tend to increase as the ratio of Yb3+ ions increases due to the increase in the molar refraction (Rm). The values of \({\alpha }_{{\mathrm{O}}^{2-}}\left({n}_{\mathrm{D}-\mathrm{S}}\right)\) and \(\Lambda \left({n}_{\mathrm{D}-\mathrm{S}}\right)\) are listed in Table 6.

Fig. 14
figure 14

Oxide ion polarizability and optical basicity as a function of Yb2O3 content

3.9 Absorption–emission cross sections and gain spectra

The absorption σabs(λ) and emission σemis(λ) cross-sections for 2F5/2 → 2F7/2 transition of Yb3+ ions are calculated based on the the optical absorbance (A), Yb3+ ions concentration (N), and glass thickness (L) by the next relationships [56]:

$${\sigma }_{\mathrm{abs}}\left(\lambda \right)=2.303 A\left(\lambda \right)/NL$$
(25)
$${\sigma }_{emis}\left(\lambda \right)={\sigma }_{abs}\left(\lambda \right)\frac{{Z}_{l}}{{Z}_{u}}\mathrm{exp}\left[ \left({E}_{zl}-\frac{hc}{\lambda }\right)/{K}_{\mathrm{B}}T\right],$$
(26)

where Zl and Zu are the partitioning functions of the lower and upper states, respectively, EZl is the energy difference between the ground state and upper state manifold, and T is the room temperature. The obtained values of the absorption and emission cross-sections for 2F5/2 → 2F7/2 transition are displayed in Fig. 15. The peaks of the emission cross-section (σemis) have the values of 16.284 × 10–21, 20.062 × 10–21, and 16.315 × 10–21 cm−2 for BTiBaYb1, BTiBaYb3, and BTiBaYb5 glasses, respectively. The high values of σemis make the present glasses strong candidate for laser and amplifier applications. Moreover, the wide emission is helpful for the creation of tunable and ultra-short pulse lasers [57].

Fig. 15
figure 15

Absorption and emission cross-sections for 2F5/2 → 2F7/2 transition of Yb3+-doped barium titanium borate glasses

The optical gain coefficient G(λ) can be calculated based on the values of the absorption and emission cross-sections using the next relation [56]:

$$G\left(\lambda \right)={\sigma }_{\mathrm{emis}}\left(\lambda \right)NP-{\sigma }_{\mathrm{abs}}\left(\lambda \right)N\left(1-P\right),$$
(27)

where P is the rate of population inversion of 2F5/2 → 2F7/2 transition and taken as: 0, 0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, 0.9, and 1.0. The obtained values of G(λ) are displayed in Fig. 16, and it is noted that the gain coefficient and gain band increase as the value of P increases, and the gain will be completely flat when the value of P is greater than 0.6. Furthermore, the gain spectra exhibit positive values in the wavelength range 971–1010 nm when the values of P are greater than 0.3.

Fig. 16
figure 16

Gain coefficient for 2F5/2 → 2F7/2 transition of Yb3+-doped barium titanium borate glasses

4 Conclusion

The density of the studied glasses increased with increasing the concentration of Yb2O3 from 3.652 to 3.761 g/cm3 while the molar volume decreased from 27.875 to 27.784 cm3/mol then increased to 28.06 cm3/mol. The distinctive bands of FTIR and Raman spectra were described and identified. The optical bandgap and metallization criterion decreased with increasing the ratio of Yb2O3 while the linear refractive index increased. The average oscillator strength, linear and non-linear optical susceptibilities, and non-linear refractive index decreased with increasing the ratio of Yb2O3. The absorption and emission cross-sections and gain coefficients for 2F5/2 → 2F7/2 transition of Yb3+ ions were calculated. The high values of emission cross-sections of the studied glasses made them strong candidates for laser and amplifier applications. Furthermore, the wide emission is helpful for creating tunable and ultra-short pulse lasers.