Next Article in Journal
Multi-Walled Carbon Nanotubes Enhance Methanogenesis from Diverse Organic Compounds in Anaerobic Sludge and River Sediments
Next Article in Special Issue
Defect Processes in Halogen Doped SnO2
Previous Article in Journal
Two-Lane Highways Crest Curve Design. The Case Study of Italian Guidelines
Previous Article in Special Issue
CQDs@NiO: An Efficient Tool for CH4 Sensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impact of Dy2O3 Substitution on the Physical, Structural and Optical Properties of Lithium–Aluminium–Borate Glass System

by
Osama Bagi Aljewaw
1,
Muhammad Khalis Abdul Karim
1,*,
Halimah Mohamed Kamari
1,
Mohd Hafiz Mohd Zaid
1,
Noramaliza Mohd Noor
2,
Iza Nurzawani Che Isa
3 and
Mohammad Hasan Abu Mhareb
4,5
1
Department of Physics, Faculty of Science, Universiti Putra Malaysia, Serdang 43400, Selangor, Malaysia
2
Department of Radiology, Faculty of Medicine, Universiti Putra Malaysia, Serdang 43400, Selangor, Malaysia
3
Diagnostic Imaging and Radiotherapy Programme, Faculty of Health Sciences, Universiti Kebangsaan Malaysia, Kuala Lumpur 50300, Malaysia
4
Department of Physics, College of Science, Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Dammam 31441, Saudi Arabia
5
Basic and Applied Scientific Research Center, Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Dammam 31441, Saudi Arabia
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(22), 8183; https://doi.org/10.3390/app10228183
Submission received: 3 October 2020 / Revised: 4 November 2020 / Accepted: 10 November 2020 / Published: 19 November 2020
(This article belongs to the Special Issue 10th Anniversary of Applied Sciences: Invited Papers in Materials)

Abstract

:
In this study, a series of Li2O-Al2O3-B2O3 glasses doped with various concentrations of Dy2O3 (where x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 mol%) were prepared by using a conventional melt-quenching technique. The structural, physical and optical properties of the glasses were examined by utilising a variety of techniques instance, X-ray diffraction (XRD), UV–Vis-NIR spectrometer, Fourier transform infrared (FTIR) and photoluminescence (PL). The XRD spectra demonstrate the amorphous phase of all glasses. Furthermore, the UV-vis-NIR spectrometers have registered optical absorption spectra a numbers of peaks which exist at 1703, 1271, 1095, 902, 841, 802, 669, 458, 393 and 352 nm congruous to the transitions from the ground of state (6H15/2) to different excited states, 6H11/2, 6F11/2 + 6H9/2, 6F9/2 + 6H7/2, 6F7/2, 6F5/2, 6F3/2, 4F9/2, 4I15/2, 4F7/2 and 6P7/2, respectively. The spectra of emission exhibit two strong emanation bands at 481 nm and 575 nm in the visible region, which correspond to the transitions 4F9/26H15/2 and 4F9/26H13/2. All prepared glass samples doped with Dy2O3 show an increase in the emission intensity with an increase in the concentration of Dy3+. Based on the obtained results, the aforementioned glass samples may have possible applications, such as optical sensor and laser applications.

1. Introduction

Borate glass has been acknowledge as a good host for various rare-earth (RE) oxides among the traditional glass formers due to their strong glass formulation when compared with other conventional systems such as phosphates, germanates, vanadates and tellurite glass [1]. In addition, the glassy system is available easily, inexpensive, simple to prepare and a good host for a variety of elements [2]. The structural, physical, and optical characteristics of the glasses are greatly influenced by the composition and synthesis conditions. Therefore, to accomplish high emission efficiency, most of the glass system is activated using suitable transitional metals and/or rare-earth elements. The incorporation of rare-earth ions to the different glassy systems led to an improvement in the optical properties, such as refractive index, optical band gaps energy and laser amplification [3]. These improvements in optical properties for glassy systems drove them to be a potential candidate for lasers, solar concentrate systems, optical detectors, waveguides and telecommunications optical fibres [4].
Borate glasses have specific properties that make them beneficial for a wide technical application, but their chemical durability is relatively feeble, which limits their utility. Nevertheless, the addendum of oxides such as lithium oxide (Li2O) and aluminium oxide (Al2O3) can enhance the chemical durability and physical properties. Moreover, adding metal oxides as modifiers to the host matrix raises the radiative parameters. Besides, glasses containing metals minimize phonon energy and lead to an increase in the luminescence quantum from excited rare-earth ion states [5]. Therefore, glasses doped with rare-earth elements may be used because of their ion emission efficiencies of 4f–4f and 4f–5d. The 5s and 5p orbits provide shielding effects to electrons determined by 4f, which leads to significant RE spectral absorption and emission lines [6,7]. Due to their unique characteristics, glass systems doped with rare-earth elements have been given considerable attention in recent decades.
Furthermore, between all the RE ions, trivalent dysprosium (Dy3+) ions have been extensively investigated for the production of or improvement in optical amplification systems in telecommunication. It is important to study the luminescence of Dy3+ ions in level 4F9/2 because it occurs in visible and NIR regions [8,9]. Usually, transitions of Dy3+ ion exhibit in yellow or blue regions of 4F9/26H13/2 (electric dipole) and 4F9/26H15/2 (magnetic dipole), respectively [10,11]. It has been established that the transition of 4F9/26H13/2 is hypersensitive, and therefore, its intensity depends heavily upon the existence of the host, while the intensity of transition of 4F9/26H15/2 is not quite sensitive to host conditions. Dy3+ has been acknowledged for its white light production, which is appropriate at an acceptable yellow to blue (Y/B) intensity ratio. Luminescent materials doped by Dy3+ ions are therefore commonly used both in glasses and phosphors to generate white light [5,6,10]. Moreover, Dy3+ ions have a various interesting characteristics such as high mechanical strength, high sensitivity and high thermal neutron absorption. They also can be utilized as the X-ray scintillators and as a yellow laser medium due to their intense yellow emission [12,13,14]. Unfortunately, there is limited information on the structure and physical properties of aluminium borate-bases doped with rare-earth ions, especially with Dy3+ ion. Hence, in this study, we seek to examine the impact of dysprosium oxide (Dy2O3) on the physical, structural and optical properties of 23Li2O-(69.5 − x) B2O3-7.5Al2O3: xDy2O3 glass series (where x = 0, 0.2, 0.4, 0.6, 0.8 and 1 mol%).

2. Materials and Methods

2.1. Glass Preparation

The new glass formulation was prepared by using the conventional melt-quench technique. Various raw materials such as lithium oxide (Li2O), aluminium oxide (Al2O3), boron oxide (B2O3) and trivalent dysprosium oxide (Dy2O3) were chosen at a specific ratio, as illustrated in Table 1. The chemical powders were weighed and mixed well for 60 min at 90 rpm. Then, the mixture was put in an alumina crucible and inserted in an electrical furnace for 30 min to be melted at 1000 °C to ensure complete melting. After that, the mixture was moved to another furnace for annealing at 400 °C for three hours. Then, the temperature of the furnace decreased progressively to reach the ambient temperature at a cooling rate of 10 °C per minute. Lastly, the samples were divided into two groups, where the first group was grinded to explore the characterisation, and the second group was buffed to investigate the optical of properties.

2.2. Structural and Physical Parameters

The weight of the prepared glass samples was measured in air and distilled water using a sensitive microbalance based on the Archimedes principle. The density, ρ, of the samples was calculated by using the following equation
ρ = ( W A ( W A W B ) ) ρ B
where WA and WB denote the sample weight in the air and distilled water, respectively, and ρB is the density of distilled water, ρ B = 0.999 g c m 3 . Molar volumes (Vm) for all glass samples were determined using the following equation
V m = ( M a v / ρ ) ( c m 3 / m o l )
where Mav represents the average molecular weight. The average boron–boron separation d B B was calculated by applying the following formula
d B B = [ V m b N A ] 1 3
where V m b refers to the volume of boron atoms per mole and is given by
V m b = V m 2 ( 1 X B )
where X B is the mole fraction and V m is Molar volume for the glass samples [15]. The ion concentration can be acquired using the next expression
N = ( X % ρ N A M T ) ( i o n c m 3 )
where X% is the mole percent of dopant, N T is the Avogadro number, and M T is the molecular weight. Based on the ion concentrations, it is also possible to compute three essential physical parameters, such as Polaron radius (rp), inter-nuclear distance (ri) and field strength
r p ( A ) = 1 2 [ π 6 N ] 1 3
r i ( A ) = [ 1 N ] 1 3
F = ( Z m r P 2 )
where Z m is the atomic mass of the dopant [16,17].
The non-crystalline phases for the selected samples were observed using X-ray diffraction (XRD) system PANalytical X’pert PRO (PW3040/60 MPD, Philips, EA, The Netherland). The system was combined with the software of diffraction analysis based on the 2θ range from 10° to 80° with the steps 0.02°. Perkin Elmer Spectrum 100 (Waltham, MA, USA) instrument was utilized to obtain the FTIR spectrum absorption for studied glasses with the size <63 μm in Attenuated-Total Reflectance (ATR) mode within wavenumber range of 400–4000 cm−1.

2.3. Optical Properties

2.3.1. UV-Vis-NIR Absorption Spectra

UV-Vis spectrophotometer for reflective spectroscopy (RSA) (Lambda 35 Perkin Elmer, MA, USA) was used to estimate the absorption spectra of the glass samples within wavelength ranges of 200–2600 nm. Using absorption, the energy gap (Eg) can be obtained by applying the Mott and Davis relation
( α h ν ) = A ( h ν E g ) n / ( h ν )
where α is coefficient of optical absorption, A is a constant, ( h ν ) is the incident photon energy, and Eg is the indirect permitted optical band gap energy. The direct Eg value is obtained from the plot ( α h ν ) 1 / 2 , and h ν by extrapolating the linear-compatible regions to the value ( α h ν ) 1 / 2 = 0 [16]. Urbach energy (Eu) offers knowledge about glass disorder. This Eu can be measured by using the relation [8]
α ( ν ) = c . exp ( h ν E u )
where c is constant and E u is Urbach energy. The oscillator strength ( f exp ) of glasses can be determined using the next equation
f exp = 4.32 10 9 ε ( ν ) d ν
where ε ( ν ) is the coefficient of molar the absorption of the each band at an energy of ν ( cm 1 ) [17,18,19]. The refractive index (n) for the electronic polarization of ions and the local field of materials is among the most important optically dependent material parameters. The next relation can be used to define the refractive index (n)
( n 2 1 ) ( n 2 + 2 ) = 1 E g 20
Reflective loss on the surface of the glass is determined by using the refractive index of the Fresnel formula
R L = ( ( n 1 ) ( n + 1 ) ) 2
Using the Volf and Lorentz–Lorenz formula, the molar refraction ( R m ) for all samples was measured [20,21]
R m = ( ( n 2 1 ) ( n 2 + 2 ) ) V m
Molar refractivity (RM) can be acquired using the following equation [22]
R m = ( n 2 1 ) ( n 2 + 2 ) ( M ρ )
The following relation can be used to estimate molar polarizability (αm) [23]
α m = ( 3 4 π N A ) R m
Metallisation criterion is the prediction of the metal or isolating behaviour of the condensed matter and is determined using the following relation [24]
M = 1 ( R m V m )
The materials are considered metallic when Rm/Vm ˃ 1, and they are considered insulating when Rm/Vm ˂ 1. Further, the polarizability of electrons (αo) and optical basicity ( Λ ) linked electronegativity ( χ ) can be obtained by [25]
χ = 0.2688 E g
where (Eg) is optical band gap. The electronic polarizability is given by
α = 0.9 χ + 3.5
The relation between the electronic oxide polarizability and optical basicity is described by [26].
Λ = 1.67 ( 1 1 α ο 2 )
The dielectric constant (ε) and optical dielectric constant can be calculated using the following formulae [8]. The dielectric constant has calculated using refractive index of the glass
ε = n 2
where (n) is the refractive index. The optical dielectric constant of the glass calculated by the following relation
p d t d p = ( ε 1 ) = n 2 1
where ε is the dielectric constant.

2.3.2. Photoluminescence (PL) Spectrum

The LS55 Luminescence Spectrophotometer (Perkin Elmer, MA, USA) was used to determine photoluminescence between the wavelengths of 200 and 1300 nm. The luminescence signal was analysed based on excitation and emission methods using a Monk–Gillieson monochromator.

3. Results and Discussion

The pattern of the XRD for all the glass samples did not show any sharp diffraction or peaks, as shown in Figure 1, which confirms the amorphous nature for all the studied glass samples. Figure 2 shows the FTIR spectrum for lithium–aluminium–borate (LAB) glass doped with various Dy3+ ion concentrations. All infrared spectrums revealed several absorption bands, as listed in Table 2.
Density ( ρ ) is a key physical parameter for analysing the physical features of glass samples, as it indicates the relation between the masses and the volume within the glass system. Likewise, the molar volume (Vm) also correlates directly to the oxygen distribution in the glass structure. Figure 3 shows the relation between the density and molar volume of the glass upon adding the different concentrations of Dy2O3.
As illustrated in Table 3, the Vm of these glasses increases slightly with increasing Dy2O3 concentration up to 0.4 mol%, but the Vm values decrease gradually from 29.28 up to 29.17 cm3 with the addition of Dy2O3 up to 1 mol%. This enhancement in Vm value can be related to the decrement in glass compactness. Upon further addition of Dy2O3, the Vm values reduce gradually as a result of the increasing compactness of the glass system [27,28,29]. The replacement with Dy2O3 instead of B2O3 changes the ratio of boron to oxygen, creating BO4 units that contribute to the compactness of the glass structure, thereby increasing the glass density. Furthermore, the molecular weight of Dy2O3 is higher than B2O3, meaning a significant increase in glass density [30].
The calculated rp, ri and F values are listed in Table 3, and Figure 4 displays the behaviour of these parameters. The decrease in rp and ri with increased Dy2O3 is related to the decrease in the Dy–O distance, as a result of which the strength of the Dy–O bond increases, producing stronger field around Dy3+ ions [8]. Besides, the addition of Dy2O3 to the glass network led to overcrowding that decreased the average distance between the RE-oxygen. The significant increment in field strength values is, therefore, due to the appearance of strong linkages in the glass matrix between Dy3+ and B ions. [10]. It is noted from this table that the boron–boron separation ‹dB-B› decreases with an increase in the Dy2O3 concentration due to the stretching force of the binds in the glass network.
Figure 5 displays the optical absorption spectrum for the Dy3+ doped lithium–aluminium–borate (LAB) glass samples within wavelengths in the range 300–1890 nm at room temperature. A spectrum identifies that the intensity of absorption raises with an addendum to Dy2O3 [29,31].
The spectra show ten inhomogeneous of absorption bands existing at the wavelengths 352, 393, 458, 669, 802, 841, 904, 1095, 1271 and 1703 nm due to the transitions of Dy3+ at the ground state (6H15/2) into different excited states (6P7/2), (4F7/2), (4I15/2), (4F9/2), (6F3/2), (6F5/2), (6F7/2), (6F9/2 + 6H7/2), (6F11/2 + 6H9/2) and (6H11/2), respectively [17]. Due to the strong absorption of LAB host glass, some of the absorption bands have disappeared in ultraviolet (UV) regions and are very sparse at 352 nm (6H15/26P7/2), 393 nm (6H15/24F7/2) and 456 nm (6H15/24I15/2), and also they have very low intensity [32]. Besides, the current samples showed a hypersensitive transition at 1270 nm (6H15/26H11/2) with high intensity and were subjected to the rule of selection |ΔS| = 0, |ΔL| ≤ 2, and |ΔJ| ≤ 2, where these transitions are more sensitive than others [33]. The variation between transition levels and their respective oscillator of strength for the glass samples are tabulated in Table 4. Noting that, the influence of Dy3+ ion on difference absorption bands led to their appropriate wavelengths, energies and oscillator strengths [34].
Figure 5 illustrate the optical energy band gaps for direct and indirect based on the curves of the UV-absorption spectrum. Figure 6a and 6b show the indirect and direct bandgap, respectively. The energy band gap value can be obtained by using Equation (9) to plot (αhν)n against photon energy (hv). Then, the linear extrapolating region of the curves extending to the X-axis gives the energy bandgap (Eg) reading. In the recent glass system, the values of the direct Eg exhibit from 3.650 to 3.706 eV, and the values of indirect (Eg) show a decrease from 3.189 to 2.556 eV with increasing dopant contents, as listed in Table 5. The declines in Eg values may be due to structural changes due to the addition of Dy3+ ions. The addition of Dy2O3 can contribute to an increase in electron localisation that increases donor centres in the glass matrix. This increment causes a decrease in Eg values [35]. This is also because a new extrinsic band is formed by Dy3+ on the grid between the boron and oxygen ions. As a consequence, there is an amount of possible reduction in (B–O–B) [36].
Figure 7 indicates that the energy of Urbach decreases with changes in the Dy2O3 concentration. The reduction in the energy of Urbach is attributed to the creation of fewer defects, as reported [37]. Figure 8 and Table 6 show a slight gradual increase in refractive index values from 2.29 to 2.35 with increasing Dy2O3 concentration that can be attributed to the increase in electronic polarizability from 2.67 to 2.75 [17]. The sample has a higher refractive index, as it has a smaller bandgap value that reflects the compactness of the glass network structure. Meanwhile, the increment in molar refractivity and electronic polarisation values with increasing Dy2O3 concentration is indicated to form more non-bridging oxygen (NBOs) in the glass matrix [12].
Table 6 presented the optical properties of the glass samples. It is observed that the molar refractivity (RM), reflection loss (RL) and refractive index (n) has a significant influence on the polarizability (αm) which demonstrates that the refractive index of the glasses does not solely depend on the density. It is known that the samples containing NBOs have great polarizability compared with samples containing bridging oxygen (BOs). The results presented here are in agreement with another work [38].
The theoretical optical basicity (Λ) is a calculation of oxygen’s capacity to contribute a negative charge load in glasses. To classify the covalent/ionic ratios of the glass, the theoretical optical basicity may be used, because the increment in (Λ) values indicate the declining covalence. Table 6 indicates that the values of Λ are within the range 1.240–1.285 and found to increase with increasing Dy2O3 concentrations. Here, the increment in optical basicity values means the ability of oxide ions to transfer electrons in the cations surrounding them [7]. Figure 9 represents the decline in the metallization criterion with an increase in the Dy2O3 concentration. The obtained values confirm the non-metallic nature of the current glass samples [8]. From Table 3, the results show that the refractivity RM, αM, ε and optical dielectric constant increase with increasing Dy2O3, meaning an increase in NBOs inside the glass matrix.
The PL spectra for the current samples of various Dy3+-doped compositions of (LAB) glass registered at room temperature in the wavelength region 420–720 nm below the excitation of the wavelength 375 nm are exhibited in Figure 10.
It was noticed that the emission peaks’ intensity increased gradually as Dy3+ concentrations increased from 0.2 mol% to 1 mol%. The bands obtained in this present work are in agreement with the previous investigations [39]. Five emission peaks were spotted, containing two comparatively intense emission bands at nearly 481 and 575 nm, respectively, for the transitions 4F9/26H15/2 and 4F9/26H13/2, and three considerably feeble bands at almost 458, 661 and 689 nm corresponding to the transitions 4I15/26H15/2, 4F9/26H11/2 and 4F9/26H9/2, respectively [40]. These transitions are similar to other study, where the transition 4I15/2 level excites the Dy3+ ions at band 458 nm [12]. The excited Dy3+ ions populate the 4F9/2 meta-stable state during rapid of non-radiative decay process due to the small energy gap between 4I15/2 and 4F9/2 states [41]. The main higher band at 575 nm (4F9/26H13/2) in the yellow range of the visible spectrum is a supersensitive transition following the selection rules (∆J = ±2 and ∆L = ±2) [42].
In addition, the band at 481nm (4F9/26H11/2), which is in the blue range, has low sensitivity to the host glass and is lower than the band at 575 nm that applies the selection rules (∆J = ±3). Therefore, the intensity of this transition is heavily impacted by the surrounding environment. Further to this, the weak intensity of emission bands appears at 661 nm (4F9/26H11/2) and 689 nm (4F9/26H9/2) is in the red range, which follows the selection rules (ΔJ = ±1) and (ΔJ = 0), 6F11/2, respectively [33,43]. Spectroscopic and luminescence properties at 0.8 mol%-doped dysprosium ion are compared with other reported glass matrices and are found to be useful for yellow lighting applications in the visible spectral region. Those emissions, especially those at the visible light range, may increase the sensitivity of the composition and are apt to be used in several applications such as sensors and solar cells. Figure 11 illustrates a partial energy level diagram for the composition (LAB) of glass doped by several concentrations of Dy3+ ions. At excitation 375 nm, the Dy3+ ions were excited from the lower level of 6H15/2 to higher-level 6P7/2. After that, when the Dy3+ ions were decayed non-radiatively to 6F9/2 level, the Dy3+ moved downward radiatively below the excitation of 4I15/2. Prominent emission bands have been identified: 6H15/2 (blue), 6H13/2 (yellow) and 6H11/26H9/2 (red). The second part of this figure represents the absorption transition from the lower level (6H15/2) into the various energy levels 6H11/2, 6F11/2 + 6H9/2, 6F9/2+ 6H7/2, 6F7/2, 6F5/2, 4F9/2, 6F3/2,4I15/2, 4F7/2 and 6P7/2. Hence the emission, absorption transitions, non-radiative, radiative energy transfer, cross-relaxation and ground state of the study samples are discussed.

4. Conclusions

In this study, new glass samples of the 23Li2O-7.5Al2O3-(69.5 − x) B2O3: xDy2O3 system were prepared using the melt-quenching technique. The amorphous nature of glass samples was confirmed by the XRD analysis. Notably, the density and optical basicity increased with the presence of BO4 tetrahedral units and, due to the structural changes, led to the decline in the direct and indirect energy gaps. From PL results, five emission bands were observed around at 458, 481, 575,661 and 689 nm, which are attributable to Dy3+ transitions of 4I15/26H15/2, 4F9/26H15/2, 4F9/26H13/2, 4F9/26H11/2 and 4F9/26H9/2, respectively. The 576 nm (4F9/26H13/2) band is the largest. Hence, optical properties and other physical parameters, such as refractive index, density, molar volume, molar refractive, electrical polarisation and optical basicity, show a strong connexion with the speciation of dysprosium ions.

Author Contributions

Conceptualization, M.K.A.K.; Data curation, O.B.A.; Formal analysis, O.B.A., M.K.A.K.; Funding acquisition, M.K.A.K. and I.N.C.I.; Investigation, M.K.A.K.; Project administration, I.N.C.I.; Resources, M.H.M.Z.; Supervision, M.K.A.K. and H.M.K.; Validation, N.M.N.; Visualization, M.H.A.M.; Writing—original draft, O.B.A.; Writing—review & editing, M.K.A.K. and M.H.A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Geran Putra IPM with the grant number [GP/IPM/9619800] and the APC was funded by Research Management Centre of Universiti Putra Malaysia.

Acknowledgments

The authors wish to acknowledge support from to Biotechnology Research Center, Libya for scholarship and Universiti Putra Malaysia for providing research facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anjaiah, J.; Laxmikanth, C.; Veeraiah, N.; Nalluri, V. Spectroscopic properties and luminescence behaviour of europium doped lithium borate glasses. Phys. B Condens. Matter 2014, 454, 148–156. [Google Scholar] [CrossRef]
  2. Alajerami, Y.S.M.; Hashim, S.; Ghoshal, S.K.; Saleh, M.A.; Kadni, T.; Saripan, M.I.; Saripand, K.; Alzimamie, Z.; Bradley, D.A. The Effect of TiO2 and MgO on the Thermoluminescence Properties of a Lithium Potassium Borate Glass System. J. Phys. Chem. Solids 2013, 74, 1816–1822. [Google Scholar] [CrossRef] [Green Version]
  3. Okasha, A.; Abdelghany, A.; Marzouk, S. The influence of Ba2+ and Sr2+ ions with the Dy3+ ions on the optical properties of lead borate glasses: Experimental and Judd–Ofelt comparative study. J. Mater. Res. Technol. 2020, 9, 59–66. [Google Scholar] [CrossRef]
  4. Ramteke, D.D.; Gedam, R.S. Study of Li2O–B2O3–Dy2O3 glasses by impedance spectroscopy. Solid State Ion. 2014, 258, 82–87. [Google Scholar] [CrossRef]
  5. Damodaraiah, S.; Prasad, V.R.; Babu, S.; Ratnakaram, Y. Structural and luminescence properties of Dy3+ doped bismuth phosphate glasses for greenish yellow light applications. Opt. Mater. 2017, 67, 14–24. [Google Scholar] [CrossRef]
  6. Pawar, P.; Munishwar, S.; Gautam, S.; Gedam, R. Physical, thermal, structural and optical properties of Dy3+ doped lithium alumino-borate glasses for bright W-LED. J. Lumin. 2017, 183, 79–88. [Google Scholar] [CrossRef]
  7. Chimalawong, P.; Kirdsiri, K.; Kaewkhao, J.; Limsuwan, P. Investigation on the Physical and Optical Properties of Dy3+ Doped Soda-Lime-Silicate Glasses. Procedia Eng. 2012, 32, 690–698. [Google Scholar] [CrossRef] [Green Version]
  8. Yusof, N.N.; Ghoshal, S.K.; Omar, M.F. Physical, structural, optical and thermoluminescence behavior of Dy2O3 doped sodium magnesium borosilicate glasses. Results Phys. 2019, 12, 827–839. [Google Scholar]
  9. Anwar, A.; Zulfiqar, S.; Yousuf, M.A.; Ragab, S.A.; Khan, M.A.; Shakir, I.; Warsi, M.F. Impact of rare earth Dy+3 cations on the various parameters of nanocrystalline nickel spinel ferrite. J. Mater. Res. Technol. 2020, 9, 5313–5325. [Google Scholar] [CrossRef]
  10. Mhareb, M.; Hashim, S.; Ghoshal, S.; Alajerami, Y.; Bqoor, M.J.; Hamdan, A.I.; Saleh, M.A.; Karim, M.K.A. Effect of Dy2O3 impurities on the physical, optical and thermoluminescence properties of lithium borate glass. J. Lumin. 2016, 177, 366–372. [Google Scholar] [CrossRef]
  11. Mhareb, M.; Hashim, S.; Ghoshal, S.K.; Alajerami, Y.S.M.; Saleh, M.A.; Azizan, S.A.B.; Razak, N.A.B.; Karim, M.K.A. Influences of dysprosium and phosphorous oxides co-doping on thermoluminescence features and kinetic parameters of lithium magnesium borate glass. J. Radioanal. Nucl. Chem. 2015, 305, 469–477. [Google Scholar] [CrossRef]
  12. Pawar, P.P.; Munishwar, S.R.; Gautam, S.; Gedam, R.S. Structural and photoluminescence properties of Dy3+ doped different modifier oxide-based lithium borate glasses. J. Lumin. 2012, 132, 2984–2991. [Google Scholar]
  13. Zelati, A.; Amirabadizadeh, A.; Hosseini, A. A facile approach to synthesize dysprosium oxide nanoparticles. Int. J. Ind. Chem. 2014, 5, 69–75. [Google Scholar] [CrossRef] [Green Version]
  14. Khan, I.; Rooh, G.; Rajaramakrishna, R.; Srisittipokakun, N.; Kim, H.J.; Kaewkhao, J.; Ruangtaweep, Y. Photoluminescence Properties of Dy3+ Ion-Doped Li2O-PbO-Gd2O3-SiO2 Glasses for White Light Application. Braz. J. Phys. 2019, 49, 605–614. [Google Scholar] [CrossRef]
  15. Mhareb, M.; Almessiere, M.; Sayyed, M.; Alajerami, Y. Physical, structural, optical and photons attenuation attributes of lithium-magnesium-borate glasses: Role of Tm2O3 doping. Optik 2019, 182, 821–831. [Google Scholar] [CrossRef]
  16. Kamaruddin, W.; Rohani, M.; Sahar, M.; Liu, H.; Sang, Y. Synthesis and characterization of lithium niobium borate glasses containing neodymium. J. Rare Earths 2016, 34, 1199–1205. [Google Scholar] [CrossRef]
  17. Ichoja, A.; Hashim, S.; Ghoshal, S.; Hashim, I.H.; Omar, R. Physical, structural and optical studies on magnesium borate glasses doped with dysprosium ion. J. Rare Earths 2018, 36, 1264–1271. [Google Scholar] [CrossRef]
  18. Effendy, N.; Wahab, Z.A.; Aziz, S.A.; Matori, K.A.; Zaid, M.H.M.; Rashid, S.S.A.; Nuraidayani, E.; Abdul, W.Z.; Hj., A.A.S.; Amin, M.K.; et al. Characterization and optical properties of erbium oxide doped ZnO–SLS glass for potential optical and optoelectronic materials. Mater. Express 2017, 7, 59–65. [Google Scholar] [CrossRef]
  19. Shaaban, K.; El-Maaref, A.; Abdelawwad, M.; Saddeek, Y.B.; Wilke, H.; Hillmer, H. Spectroscopic properties and Judd-Ofelt analysis of Dy3+ ions in molybdenum borosilicate glasses. J. Lumin. 2018, 196, 477–484. [Google Scholar] [CrossRef]
  20. Babu, M.R.; Babu, A.M.; Moorthy, L.R. Structural and optical properties of Nd3+-doped lead borosilicate glasses for broadband laser amplification. Int. J. Appl. Eng. Res. 2018, 13, 7692–7700. [Google Scholar]
  21. Bhatia, B.; Meena, S.L.; Parihar, V.; Poonia, M. Optical Basicity and Polarizability of Nd3+-Doped Bismuth Borate Glasses. New J. Glas. Ceram. 2015, 5, 44–52. [Google Scholar] [CrossRef] [Green Version]
  22. Gedam, R.; Ramteke, D. Influence of CeO2 addition on the electrical and optical properties of lithium borate glasses. J. Phys. Chem. Solids 2013, 74, 1399–1402. [Google Scholar] [CrossRef]
  23. Hasnimulyati, L.; Halimah, M.K.; Zakaria, A.; Halim, S.A.; Ishak, M.; Eevon, C. Structural and optical properties of Tm2O3-doped zinc borotellurite glass system. J. Ovonic Res. 2016, 12, 291–299. [Google Scholar]
  24. Kundu, V.; Dhiman, R.L.; Maan, A.S.; Goyal, D.R.; Garg, A.B.; Mittal, R.; Mukhopadhyay, R. Optical Properties of Alkaline Earth Ions Doped Bismuth Borate Glasses. AIP Conf. Proc. 2011, 1349, 545. [Google Scholar] [CrossRef]
  25. Kaur, P.; Singh, K.; Kurudirek, M.; Thakur, S. Study of environment friendly bismuth incorporated lithium borate glass system for structural, gamma-ray and fast neutron shielding properties. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 223, 117309. [Google Scholar] [CrossRef]
  26. Algradee, M.A.; Sultan, M.; Samir, O.M.; Alwany, A.E.B. Electronic polarizability, optical basicity and interaction parameter for Nd2O3 doped lithium–zinc–phosphate glasses. Appl. Phys. A Mate. Sci. Process. 2017, 123, 524. [Google Scholar] [CrossRef]
  27. Pawar, P.; Munishwar, S.R.; Gedam, R. Eu2O3 doped bright orange-red luminescent lithium alumino-borate glasses for solid state lighting. J. Lumin. 2018, 200, 216–224. [Google Scholar] [CrossRef]
  28. Deepa, A.V.; Priya, M.; Suresh, S. Influence of Samarium Oxide ions on structural and optical properties of borate glasses. Sci. Res. Essays 2016, 11, 57–63. [Google Scholar] [CrossRef] [Green Version]
  29. Chandrasekhar, M.; Nagabhushana, H.; SudheerKumar, K.; Dhananjaya, N.; Sharma, S.; Kavyashree, D.; Shivakumara, C.; Nagabhushana, B. Comparison of structural and luminescence properties of Dy2O3 nanopowders synthesized by co-precipitation and green combustion routes. Mater. Res. Bull. 2014, 55, 237–245. [Google Scholar] [CrossRef]
  30. Ramteke, D.; Ganvir, V.; Munishwar, S.R.; Gedam, R. Concentration Effect of Sm3+ Ions on Structural and Luminescence Properties of Lithium Borate Glasses. Phys. Procedia 2015, 76, 25–30. [Google Scholar] [CrossRef] [Green Version]
  31. Pawar, P.; Munishwar, S.; Gedam, R. Intense white light luminescent Dy3+ doped lithium borate glasses for W-LED: A correlation between physical, thermal, structural and optical properties. Solid State Sci. 2017, 64, 41–50. [Google Scholar] [CrossRef]
  32. Shamshad, L.; Rooh, G.; Kirdsiri, K.; Srisittipokakun, N.; Damdee, B.; Kim, H.; Kaewkhao, J. Photoluminescence and white light generation behavior of lithium gadolinium silicoborate glasses. J. Alloys Compd. 2017, 695, 2347–2355. [Google Scholar] [CrossRef]
  33. Deopa, N.; Rao, A. Photoluminescence and energy transfer studies of Dy3+ ions doped lithium lead alumino borate glasses for w-LED and laser applications. J. Lumin. 2017, 192, 832–841. [Google Scholar] [CrossRef]
  34. Dawaud, R.S.E.S.; Hashim, S.; Alajerami, Y.; Mhareb, M.; Tamchek, N. Optical and structural properties of lithium sodium borate glasses doped Dy3+ ions. J. Mol. Struct. 2014, 1075, 113–117. [Google Scholar] [CrossRef]
  35. Pawar, P.; Munishwar, S.; Ramteke, D.; Gedam, R. Physical, structural, thermal and spectroscopic investigation of Sm2O3 doped LAB glasses for orange LED. J. Lumin. 2019, 208, 443–452. [Google Scholar] [CrossRef]
  36. Azizan, S.A.; Hashim, S.; Razak, N.A.; Mhareb, M.H.A.; Alajerami, Y.S.M.; Tamchek, N. Physical and optical properties of Dy3+: Li2O–K2O–B2O3 glasses. J. Mol. Struct. 2014, 1076, 20–25. [Google Scholar] [CrossRef]
  37. Obayes, H.K.; Hussin, R.; Wagiran, H.; Saeed, M. Strontium ion concentration effects on structural and spectral properties of Li4Sr(BO3)3 glass. J. Non-Cryst. Solids 2015, 427, 83–90. [Google Scholar] [CrossRef] [Green Version]
  38. Yusof, N.N.; Ghoshal, S.; Omar, M.F. Modified absorption attributes of neodymium doped magnesium-zinc-sulfophosphate glass. Malays. J. Fundam. Appl. Sci. 2017, 13, 258–262. [Google Scholar] [CrossRef]
  39. Bulus, I.; Dalhatu, S.A.; Isah, M.; Hussin, R.; Soje, E.A. Structural and Luminescence Characterization of Lithium-Borosulfophosphate Glasses Containing Dysprosium Ions. Sci. World J. 2017, 12, 98–101. [Google Scholar]
  40. Rani, P.R.; Venkateswarlu, M.; Mahamuda, S.; Swapna, K.; Deopa, N.; Rao, A. Spectroscopic studies of Dy3+ ions doped barium lead alumino fluoro borate glasses. J. Alloys Compd. 2019, 787, 503–518. [Google Scholar] [CrossRef]
  41. Babu, A.M.; Jamalaiah, B.; Kumar, J.S.; Sasikala, T.; Moorthy, L.R. Spectroscopic and photoluminescence properties of Dy3+-doped lead tungsten tellurite glasses for laser materials. J. Alloys Compd. 2011, 509, 457–462. [Google Scholar] [CrossRef]
  42. Shasmal, N.; Karmakar, B. White light-emitting Dy3+-doped transparent chloroborosilicate glass: Synthesis and optical properties. J. Asian Ceram. Soc. 2018, 7, 42–52. [Google Scholar] [CrossRef] [Green Version]
  43. Deopa, N.; Rao, A.; Gupta, M.; Prakash, G.V. Spectroscopic investigations of Nd3+ doped Lithium Lead Alumino Borate glasses for 1.06 μm laser applications. Opt. Mater. 2018, 75, 127–134. [Google Scholar] [CrossRef]
Figure 1. XRD pattern for all studied glasses.
Figure 1. XRD pattern for all studied glasses.
Applsci 10 08183 g001
Figure 2. FTIR of spectra for the various concentration of Dy3+ doped in Li2O–Al2O3–B2O3 glasses.
Figure 2. FTIR of spectra for the various concentration of Dy3+ doped in Li2O–Al2O3–B2O3 glasses.
Applsci 10 08183 g002
Figure 3. Variation in density ( ρ ) and molar volume (Vm) concerning Dy2O3.
Figure 3. Variation in density ( ρ ) and molar volume (Vm) concerning Dy2O3.
Applsci 10 08183 g003
Figure 4. Polaron radius, inter-nuclear distance and field strength as functions of Dy2O3 content of prepared glasses.
Figure 4. Polaron radius, inter-nuclear distance and field strength as functions of Dy2O3 content of prepared glasses.
Applsci 10 08183 g004
Figure 5. UV–Vis-NIR optical absorption spectra of different concentrations of Dy3+-doped LAB glasses.
Figure 5. UV–Vis-NIR optical absorption spectra of different concentrations of Dy3+-doped LAB glasses.
Applsci 10 08183 g005
Figure 6. Bang gap energy: (a) Tauc’s plot for allowed direct transitions (n = 2); (b) Tauc’s plot for allowed indirect transitions (n = 1/2).
Figure 6. Bang gap energy: (a) Tauc’s plot for allowed direct transitions (n = 2); (b) Tauc’s plot for allowed indirect transitions (n = 1/2).
Applsci 10 08183 g006aApplsci 10 08183 g006b
Figure 7. Variation of Urbach energy (eV) with mol% of Dy2O3.
Figure 7. Variation of Urbach energy (eV) with mol% of Dy2O3.
Applsci 10 08183 g007
Figure 8. Variation in refractive index (n) and Electron polarizability (αo) with concerning Dy2O3.
Figure 8. Variation in refractive index (n) and Electron polarizability (αo) with concerning Dy2O3.
Applsci 10 08183 g008
Figure 9. Variation in metallisation criterion with mol% Dy2O3.
Figure 9. Variation in metallisation criterion with mol% Dy2O3.
Applsci 10 08183 g009
Figure 10. Emission spectra of LAB glasses doped with different concentrations (in mol%) of Dy3+ ions.
Figure 10. Emission spectra of LAB glasses doped with different concentrations (in mol%) of Dy3+ ions.
Applsci 10 08183 g010
Figure 11. Partial energy levels of the RE ion showing distinct transitions.
Figure 11. Partial energy levels of the RE ion showing distinct transitions.
Applsci 10 08183 g011
Table 1. Composition ratios for all studied glass samples.
Table 1. Composition ratios for all studied glass samples.
Glass Samplesmol%
Li2OAl2O3B2O3Dy2O3
LABD-0.0237.569.50.0
LABD-0.2237.569.30.2
LABD-0.4237.569.10.4
LABD-0.6237.568.90.6
LABD-0.8237.568.70.8
LABD-1237.568.51
Table 2. FTIR assignment bands of Li2O–Al2O3–B2O3: Dy2O3 glasses.
Table 2. FTIR assignment bands of Li2O–Al2O3–B2O3: Dy2O3 glasses.
Positions of Band (cm−1) for LAB Glasses with Diverse Dy3+ Ions Contents (mol%)Band Assignments
0.00.20.40.60.81.0
412.7, 526.5416.6, 586.3426.2, 540.06433.9, 578.6406.6, 588.3406.9, 563.2Vibrations of Li+ ions [6].
692.4696.3694.3700.1696.3692.4Bending vibration B-O-B linkages with BO3 units altogether with B-O-B bending vibration of bridging oxygen’s in BO3 and bonds in AlO6 groups [27].
925.8, 1051.2918.1, 1056.9925.8, 1043.4923.9, 1055.06933.5, 1033.8920.05, 1024.2B-O stretching of tetrahedral BO4 bond [10].
1227.9, 1334.7, 1375.21244.09, 1332.8, 1396.41236.3, 1340.5, 1375.21246.02, 1328.9, 1390.61246.2, 1377.11247.9, 1382.9B–O–, stretching in pyroborate units Stretching of the trigonal BO3 units [8].
1687.71687.71687.7, 1801.51687.71687.7, 1870.9B-O-H bridge, OH bending vibration [28].
3269.3----------Stretching of OH groups or O-H (H2O bond) [29].
Table 3. Physical properties for all series of glasses.
Table 3. Physical properties for all series of glasses.
Physical Parameters *UnitsDoping (mol% Dy2O3)
0.00.20.40.60.81
Density(g.cm−3)2.1442.1622.1822.2102.2382.260
Molecular weight, MTg62.90363.50964.11664.72365.33065.937
Molar   volume   V m (cm3/mol)29.33929.37529.38429.28629.19129.175
The volume of boron atoms per mole V m b 16.03111.95911.88611.05811.65611.576
Ion concentration (N) × 1020(ions/cm3) 0.4100.8191.2331.6502.064
Polaron radius
(rp) × 10−8
11.6879.2798.0967.3476.819
Inter-nuclear distance (ri) × 10−8 29.000423.02720.09118.23216.921
Field strength (F) × 1016(cm−2) 1.1891.8872.4793.0103.494
Average of boron-boron distance ‹dB-Bnm0.2980.2700.2700.2630.2680.267
* With consideration of ± 0.01% error.
Table 4. The difference at transition levels and oscillator strengths.
Table 4. The difference at transition levels and oscillator strengths.
Absorption TransitionWavelength
(nm)
Energy
(×10−3cm−1)
Oscillator Strength fexp (×10−6)
6H15/26P7/235229.3210.451
6H15/24F7/239325.9450.520
6H15/24I15/245822.2340.696
6H15/24F9/266913.5230.554
6H15/26F3/280212.6440.615
6H15/26F5/284111.1130.478
6H15/26F7/290210.1451.305
6H15/26F9/2 + 6H7/210959.2671.821
6H15/26F11/2 + 6H9/212717.8343.218
6H15/26H11/217033.5333.046
Table 5. Direct and indirect band gaps, Urbach energy (Eu) and cut-off wavelength (λc) of the samples studied.
Table 5. Direct and indirect band gaps, Urbach energy (Eu) and cut-off wavelength (λc) of the samples studied.
SampleEdir (eV)Eindir (eV)Eu (eV)λc [nm]
LABD-0.03.5803.1893.520362.36
LADB-0.23.7063.1663.337368.83
LABD-0.43.5662.9193.272376.44
LABD-0.63.5852.8863.075385.81
LABD-0.83.4732.70053.352394.15
LABD-13.6502.5563.058401.81
Table 6. Optical properties of prepared glass samples.
Table 6. Optical properties of prepared glass samples.
MeasurementDy3+ Doped Concentration, mol%
0.00.20.40.60.81.0
Refractive index (n)2.2952.3052.3172.3312.3442.355
Reflection loss (RL)0.1540.1550.1570.1590.1610.163
Molar refraction (Rm) (cm−3)5.7414.3304.3554.3664.3754.393
Oxygen packing density (OPD)86.5786.4686.4486.7287.0187.81
Optical basicity (Λ)1.2401.2481.2571.2681.2771.285
Optical electronegativity (χ)0.9190.9030.8850.8630.8450.829
Metallisation criterion (M)0.4120.4100.4070.4030.4000.397
Molar refractivity (RM) (cm−3)17.22717.32417.42017.46717.50517.576
Molar polarizability (αm) ×10−24 (cm−3)2.2751.7161.7161.7301.7341.741
Dielectric constant (ε)5.2675.3135.3685.4335.4945.546
Optical dielectric constant4.2674.3134.3684.4334.4944.546
Electron polarizability (αo)2.6722.6872.7032.7232.7392.753
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bagi Aljewaw, O.; Karim, M.K.A.; Mohamed Kamari, H.; Mohd Zaid, M.H.; Mohd Noor, N.; Che Isa, I.N.; Abu Mhareb, M.H. Impact of Dy2O3 Substitution on the Physical, Structural and Optical Properties of Lithium–Aluminium–Borate Glass System. Appl. Sci. 2020, 10, 8183. https://doi.org/10.3390/app10228183

AMA Style

Bagi Aljewaw O, Karim MKA, Mohamed Kamari H, Mohd Zaid MH, Mohd Noor N, Che Isa IN, Abu Mhareb MH. Impact of Dy2O3 Substitution on the Physical, Structural and Optical Properties of Lithium–Aluminium–Borate Glass System. Applied Sciences. 2020; 10(22):8183. https://doi.org/10.3390/app10228183

Chicago/Turabian Style

Bagi Aljewaw, Osama, Muhammad Khalis Abdul Karim, Halimah Mohamed Kamari, Mohd Hafiz Mohd Zaid, Noramaliza Mohd Noor, Iza Nurzawani Che Isa, and Mohammad Hasan Abu Mhareb. 2020. "Impact of Dy2O3 Substitution on the Physical, Structural and Optical Properties of Lithium–Aluminium–Borate Glass System" Applied Sciences 10, no. 22: 8183. https://doi.org/10.3390/app10228183

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