Next Article in Journal
Geometric Progression of Optical Vortices
Next Article in Special Issue
650 W All-Fiber Single-Frequency Polarization-Maintaining Fiber Amplifier Based on Hybrid Wavelength Pumping and Tapered Yb-Doped Fibers
Previous Article in Journal
Drift Artifacts Correction for Laboratory Cone-Beam Nanoscale X-ray Computed Tomography by Fitting the Partial Trajectory of Projection Centroid
Previous Article in Special Issue
A 102 W High-Power Linearly-Polarized All-Fiber Single-Frequency Laser at 1560 nm
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Luminescence and Gamma Spectroscopy of Phosphate Glass Doped with Nd3+/Yb3+ and Their Multifunctional Applications

1
LaMaCoP, Faculty of Sciences of Sfax, University of Sfax, Sfax 3018, Tunisia
2
Department of Radiological Sciences, College of Applied Medical Sciences, King Khalid University, Abha 61421, Saudi Arabia
3
BioImaging Unit, Space Research Centre, Department of Physics and Astronomy, University of Leicester, Leicester LE1 7RH, UK
4
Department of Medical Physics and Instrumentation, National Cancer Institute, University of Gezira, Wad Medani 2667, Sudan
5
Physics Department, Faculty of Science, King Khalid University, Abha 61413, Saudi Arabia
6
Physics Department, College of Arts and Sciences Jouf University, Tabrjal 74713, Saudi Arabia
7
Department of Material Science, Institute of Graduate Studies and Researches, Alexandria University, 163 Horreya Avenue, Shatby, Alexandria 21526, Egypt
8
Faculty of Materials Science and Ceramics, AGH—University of Science and Technology, Al. Mickiewicza 30, 30-059 Cracow, Poland
9
Research Center for Advanced Materials Science (RCAMS), King Khalid University, Abha 61413, Saudi Arabia
*
Author to whom correspondence should be addressed.
Photonics 2022, 9(6), 406; https://doi.org/10.3390/photonics9060406
Submission received: 8 April 2022 / Revised: 26 May 2022 / Accepted: 3 June 2022 / Published: 8 June 2022
(This article belongs to the Special Issue Rare-Earth-Doped Fiber Lasers and Amplifiers)

Abstract

:
A new glass with a composition of 40P2O5-30ZnO-20LiCl-10BaF2 (in mol%), doped with 3.5Nd2O3-3.5Yb2O3, was fabricated by the quenching melt technique. The luminescence (PL) and gamma spectroscopy of the glass were investigated systematically. The spectroscopic parameters of the prepared glass, such as the optical energy gap, Judd–Ofelt parameters Ωk (where k = 2, 4 and 6), lifetimes and branching ratio of the Nd3+/Yb3+ level, were evaluated. Moreover, the shielding parameters, such as the linear and mass attenuation coefficients, mean free path and half-value layer, were evaluated. The prepared glass had a spectroscopic quality factor (Ω46) of 0.84, which is about three-times larger than that of the most standard laser host, Nd3+:YAG. The energy of the 2P1/2 (Nd3+) level (~23,250 cm−1) was twice the energy of the Yb3+ transition (~10,290 cm−1). The value of the emission cross section ( σ e m ( λ ) ) of Nd3+:4F3/24I9/2 and Yb3+:2F5/22F7/2 were 2.23 × 10−24 cm2 and 2.88 × 10−24 cm2, respectively. The fabricated glass had a high emission cross section and low mean free path parameters, which makes the fabricated glass a potential candidate for multifunctional applications, such as laser emissions for medical purposes.

1. Introduction

The codoped ions Nd3+ and Yb3+ are some of the best examined rare-earth elements and are used in fiber lasers. In both glasses and crystals, efficient energy transfer between Nd3+ and Yb3+ ions has been established. Nd3+/Yb3+ glasses are applied as compact sources in non-linear microscope technology in image processing with a THz frequency [1,2].
It is worth exploring the higher excited states of Nd3+ involved in the energy transfer from Nd3+ to Yb3+. The lanthanide ion Yb3+ emits a wavelength of approximately 1000 nm, which may be absorbed without losing any energy. Moreover, the longer lifetime and broader absorption and emission bands of Yb3+ ions allow greater energy-storage efficiency in comparison with other rare earth ions, namely the Nd3+ ions [3,4]. Although the lanthanide ion Yb3+ has just one excited state, around 10,000 cm−1 above the ground state, the pump wavelength range of Yb3+ ions is limited to 980 nm by this simple energy level scheme. Down-conversion may occur if a suitable sensitizer with energy levels of 20,000 cm−1, two times the energy of Yb3+, is used [5].
However, laser glasses have been specifically designed for photodynamic therapy surgery and medical radiation technology. Researchers investigated the optical and shielding properties of such glasses for their usefulness as a safety replacement device in medical workplaces, such as X-ray and atomic projects in the field of inventions, V-ray equipment, gamma camera rooms and examination workplaces for computed tomography (CT) [6,7]. The radiation protecting characteristics of Erbium Zinc–Tellurite glass were inspected to compare a vast range of radiation energies for health and medical imaging applications (20, 30, 40 and 60 keV) [8]. Furthermore, the laser diodes for fiber optical intercommunicating applications exhibit near-infrared (NIR) discharge at a wavelength of 1.54 μm that can be employed in armed forces, detective work and medical investigations facilities [9].
Therefore, we fabricated phosphate laser glass with the composition P2O5-ZnO-LiCl-BaF2 (PZLB) as the host lattice due to the low matrix phonon energy codoped with Nd3+/Yb3+ ions, which decreases the non-radiative transition rate to lower states. This is beneficial for the cross-relaxation energy transfer between Nd3+ and Yb3+. Thus, the spectroscopic properties of this glass were comprehensively investigated; furthermore, the energy transfer efficiency from Nd3+ to Yb3+ is also discussed. In addition to that, we measured attenuation radiation parameters, such as the linear attenuation coefficient, LAC; the mass attenuation coefficient, MAC; half-value layer, HVL; and the mean free path, MFP, of the prepared glass.

2. Materials and Methods

A homogeneity Nd3+/Yb3+doped glass with the composition (in mol%) 40P2O5-30ZnO -20LiCl-10BaF2-3.5Nd2O3-3.5Yb2O3, referred to as PZLBNdYb glass, was prepared by the melt quenching technique. We mixed the chemical very well in an alumina crucible. We placed it into a muffle furnace to be melted at 1200 °C for 1 h following the melt quenching technique. The molten chemical mixture was then quenched in a preheated copper mold. The quenched glass was heat-treated at 450 °C for 2 h at the beginning of the annealing process to eliminate any strain on sample molecules.
The density of the glass was measured using a gas pycnometer (Model: UltraPyc 1200e). The measured density of the PZLBNdYb glass sample was ρ = 4.426 ± 0.001 g/cm3. Using a UV-VIS-NIR spectrophotometer, the optical absorption spectra of the glasses were measured in the wavelength range of 190–2500 nm (JASCO, V-570).
The concentration of Nd3+ (or Yb3+) ions can be calculated by the following expression:
N N d 3 + = 2 3.5 ρ A υ 100 M
where M is the molecular weight of PZLBNdYb glass and Aν is Avogadro’s number. The concentration of the dopant Nd3+ or Yb3+ in PZLBNdYb glass was NNd3+ = NYb3+ = 1.405 × 1027 ions/m3.
The shielding parameters—the linear attenuation coefficient (LAC), mass attenuation coefficient (MAC), half-value layer (HVL) and mean free path (MFP)—of the proposed sample were measured using the NaI detector system (SPECTECH-NaI 1.5 PX 1.5/2.0 IV, S/N 010723-6) connected to a computer-based multichannel analyzer (MCA). Different gamma sources (Am241-5µCI-59.5 keV, Cs137-5µCI-662 keV, Co60-5µCI-1170 and 1330 keV) were used to produce a collimated beam at the detector level. Our group study, in Ref. [10], detailed the experimental setup utilized for detecting the incident and transmission radiation intensities in Figure 1.
The values of μ m , μ , HVL and MFP parameters can be calculated using the following relations.
{ μ = l n I o I x , μ m = l n I o I ρ x , H V L = 0.693 μ , and   M F P = 1 μ
In Ref. [10], we detailed the experimental setup utilized for detecting the incident and transmission radiation intensities in Figure 1.

3. Results and Discussion

Absorption Spectroscopy, Optical Energy Gap and Judd–Ofelt Analysis

The UV–Vis–NIR absorption spectra for the PZLBNdYb sample are recorded in Figure 2. The absorption spectra displayed seven bands centered around 352, 428, 470, 522, 580, 628, 682, 744, 802, 882 and 980 nm. All bands originated from the absorption transition from the ground state (4I9/2) to different excited states of the Nd3+(4f3) ion, except the 980 nm peak, which corresponds to the 2F7/22F5/2 absorption transition for ytterbium ions.
The Nd3+ ion bands detected are comparable to those previously described [11,12], except for minor changes in the peak positions and relative intensities. This can be attributed to the nature of various ligand fields of different glass matrixes [13]. The maximum absorption coefficient is observed at 590 nm wavelength, corresponding to the 4I9/22G7/2 + 4G5/2 hypersensitive transition. In contrast, the absorption peak of 4I9/22H11/2 (at 628 nm) was the lowest and was weak enough to be considered in the computation.
The Stark structure was poorly resolved for all bands due to the inhomogeneous broadening, and nearby energy levels overlapped—appearing as a single peak in the measured spectra. The number of long orders in the host causes changes in the micro symmetry surrounding the Nd3+ ions, resulting in amplification of the absorption bands. In other words, the linewidth of the various transitions is a measure of the Stark splitting of the J-manifold. The inhomogeneous broadening is due to the site-to-site variation in the local field seen by the rare earth ion.
Extrapolation of the linear relation, Equation (3), of the plotted absorption curves yielded the optical energy gap, Eopt, which reflected the transition from localized states at the top of the valence band into localized ones in the conduction band or vice versa.
The optical energy gap, Eopt, was calculated by using the absorption spectra of the produced glass as follows [7];
[ α ( 𝜈 ) h 𝜈 ] 1 r = A ( h 𝜈 Ε o p t )
where α ( 𝜈 ) is the absorption coefficient, A is a constant, h 𝜈 is the incident radiation’s photon energy, and r is an index that varies depending on the transition type (direct or indirect). The best fit to the result obtaining that r = 2 indicated the indirect allowed transition in the gap. The optical band edge is obtained by extrapolating from the linear range in plots of ( α h 𝜈 ) 1 2 versus h 𝜈 , as shown in Figure 3. The optical energy gap Eopt was 4.41 eV higher than phosphate glasses doped with rare earth [11,12,13,14,15,16,17]. This wide optical energy gap range shows that this glass is superior and a good medium as an acceptor of the rare earth donor from the optical fiber (non-linear/laser waveguide).
To determine the spectroscopic parameters of this glass system, the Judd–Ofelt (JO) analysis was applied to the absorption intensities of Nd3+ doped in PZLBNdYb. The detailed applications of the JO model can be found in the literature [17,18]. A brief outline of the JO analysis is given hereafter.
The measured line strength S exp ( J J ) of a given band is determined by the following expression:
S exp ( J J ) = 9 n ( n 2 + 2 ) 2 · 4 π ε 0 · 3 c · h · ( 2 J + 1 ) 8 π 3 e 2 . 2.303 N · l · λ · Γ ( λ )
where c is the velocity of light, h is the Planck’s constant, e is the elementary charge, J is the angular momentum of the initial state, N is the density of Nd3+ ions, λ is the mean wavelength of the absorption bands, l is the thickness of the studied sample (l = 7.82 mm), and n is the refractive index dispersion. Γ = O D ( λ ) · d λ is the total area under the absorbed band and can be used to calculate the experimental integrated optical density in the wavelength range.
The magnetic-dipole contribution has been ignored in this work as its impact on the measured line strength is relatively negligible for absorption transitions of Nd3+ ion [15,19,20].
The results of the intensity measurements and line strength calculations for transitions of Nd3+ ions are reported in Table 1.
On the other hand, in the Judd–Ofelt theory, the line strength S c a l ( J J ) between initial state J characterized by (S, L, J) and the final state J’ given by (S′, L′, J′) can be calculated by the following expression [17,18]:
S c a l ( J J ) = t = 2 ,   4 ,   6 Ω t | S L J U ( t ) S L J | 2
where Ωt (t = 2, 4, 6) denote the Judd–Ofelt variables and ||U(t)||2 (t = 2, 4, 6) are the doubly reduced matrix elements.
The reduced matrix elements could be found in the literature [13,21,22,23]. For two or more manifolds, the reduced matrix elements are taken as the sum of the corresponding matrix elements. The values of matrix elements for each absorption band are given in Table 1.
Based on Equations (4) and (5), a fitting between experimental and calculated line strengths of the absorption transitions provides the values of the three JO parameters. A least squares fitting of S exp to S c a l values was used to compute the Ωt (t = 2, 4, 6) parameters for the studied PZLBNdYb sample.
According to matrix element and Sexp values, one can notice that, principally, the Ω2 parameter depends on 4I9/24G5/2 + 2G7/2 + 2H11/2 (580 nm) and 4I9/24D3/2 + 4D5/2 + 2I11/2 + 4D1/2 (352 nm) peak intensities, Ω4 also depends on this later peak, while Ω6 depends on the 4I9/24F7/2 + 4S3/2 (744 nm) absorbance. Thus, one can consider that the 4I9/24D3/2 + 4D5/2 + 2I11/2 + 4D1/2 (352 nm) transition is the most influencing line in the Ωt computation for Nd3+ doped glasses.
The obtained JO parameters of Nd3+ in PZLBNdYb glass are given in Table 2 with those for Nd3+ doped in other hosts. For Yb3+, it is not possible to determine the three Ωt intensity parameters because only one transition (corresponding to the 2F7/22F5/2 transition at 980 nm) can be observed.
The calculated Judd–Ofelt parameters are in good agreement with literature values for Nd3+ doped glasses. The JO intensity parameters for this PZLBNdYb glass followed the trend Ω6 > Ω4 > Ω2. This trend is similar to that of commercial laser glasses [25], YAG:Nd3+ [24], as well as LGBaBNd [12], 30,000 ppm Nd3+/Yb3+ phosphate [13], AEBTNd0.1 [16], ZnBBi [27], 75NaPO3-24CaF2-1NdF3 [28] glasses (see Table 2).
The intensity parameter, Ω2, strongly correlated with the local structure of the rare earth ions and the covalency degree lanthanide–O bonds, which is equivalent to the dynamic polarization of the ligands. In comparison, the Ω4 and Ω6 parameters depend upon the rigidity and viscosity of the host glasses [16,31,32]. Similar to the other reported work (mentioned in Table 2), the larger Ω6 in the present glass indicates its high rigidity, and the lower Ω2 indicates the higher asymmetry and lower covalency between the Nd-O group in this PZLBNdYb glass.
The obtained Ωt values are used to recalculate the transition line strengths Scal of the absorption bands from Equation (5) and deduce the rms deviation. The root-mean-square (rms) deviation of the fit between experimental and calculated oscillator strengths is deduced by the expression. δ r m s = ( ( S exp S c a l ) 2 / ( N t r a n s 3 ) ) 1 / 2
The value of the rms deviation calculated in the present work is equal 1.019 × 10−25 m2. This slightly elevated value of rms deviation can be explained by overlaps of absorption bands in the UV region.
The spectroscopic quality factor, χ = Ω46, is an important characteristic in predicting the stimulated emission cross section for the laser active media. In the case of Nd3+, it is indicated that the smaller the ratio, the more intense the laser 4F3/24I11/2 transition. Due to the decreased matrix element’s zero value, this occurs F 3/2 4 U ( t ) 2 I J 4 of Nd3+ ion [33,34,35].
For this PZLBNdYb glass, the spectroscopic quality factor (Ω46) was 0.84, which is about three-times larger than that of the standard laser host Nd3+:YAG. Usually, χ is in the range from 0.22 to 1.5 for Nd3+ in several host materials (see Table 2).
According to the absorption spectrum of Figure 2, the emission band 4F3/24I9/2 of Nd3+ partially overlapped with the absorption band 2F5/22F7/2 of Yb3+, which ensures that the Yb3+ ion absorbs the NIR light very effectively. The wideness and lowness of the Yb3+ absorption peak ascribe to the Nd3+→Yb3+ energy transfer. However, the short separation distance between ions, estimated as d = (3/4πN)1/3 = 7Å, confirms an efficient Nd3+→Yb3+ energy transfer.
It is worth noting that systems based on Nd3+/Yb3+ energy transfers are interesting because they combine the Yb3+ ion’s good laser emission characteristics with the Nd3+ ion’s multiple intense absorption bands, which could be used for pumping with a variety of sources (laser diodes, flash lamps, solar radiation, etc.).
The 4F3/2 level for Nd3+ ions is the only exciting J manifold that did not relax predominantly by a multiphonon process. This level fluoresces in the four bands ascribed to 4F3/24I9/2 at 882 nm, 4F3/24I11/2 around 1056 nm, 4F3/24I13/2 at 1318 nm and 4F3/24I15/2 at 1870 nm. The emission line strengths attributed to the transition from 4F3/2 to 4IJ manifolds were computed through Equation (5) using the Ω2, Ω4 and Ω6 parameters. The Nd3+→Yb3+ energy transfer competes with the self-quenching of Nd3+ emission due to the cross-relaxation of the resonant process (4F3/2 + 4I9/24I15/2 +4I15/2) or phonon-assisted process (4F3/2 + 4I9/24I15/2 +4I13/2) between the Nd3+ ions.
The spontaneous emission probability AJJ′ for a transition from a J-multiplet to a lower J′-multiples is calculated as:
A J J = 1 4 π ε 0 64 π 4   e 2   υ 3   n ( n 2 + 2 ) 2 27 h ( 2 J + 1 ) S c a l
S c a l is the corresponding emission line strength.
The radiative lifetime τ for electric dipole transitions between an excited state (J) and the lower-lying terminal manifolds (J′) could also be estimated as follows:
τ = 1 J A J J
The sum is taken over all final states J′. The fluorescence branching ratio is a critical parameter to the laser designer; it might be calculated by predicting the relative strength of lines from certain excited states and describing the possibility of achieving stimulated emission from a specific transition. It is determined by:
β = A J J . τ
Table 3 shows the radiative transition probabilities ( A J J ), radiative lifetimes (τ) and branching ratios (β) of energy levels 4I9/2,4I11/2, 4I13/2, 4I15/2 and 4F3/2 of Nd3+ ion in the PZLBNdYb glass.
A critical factor in the success of the Nd3+ amplifier is the long lifetime of the 4F3/2 metastable state that permits the required high population inversion to be obtained. The radiative lifetime of the 4F3/2 state was calculated to be 1.644 ms, which is an essential metric to consider when considering the pumping need for the laser action threshold. The trend of lifetimes appears to be decreasing as 4I11/2 > 4I13/2 > 4I15/2 > 4F3/2. The branching ratios for emission transitions are steadily decreasing as follows: 4I13/24I9/2 > 4I15/24I11/2 > 4F3/24I11/2 > 4F3/24I9/2 > 4F3/24I13/2 > 4F3/24I15/2. A similar tendency of branching ratios was noticed by Zamen et al. [12] and James et al. [14].
In order to understand the probability of lasing action between 4F3/24I11/2 and 4F3/24I9/2 of Nd3+ ion, the essential parameters, such as the peak wavelength (λ), transition probability (A) and branching ratio (β) for the 4F3/24I11/2 and 4F3/24I9/2 transitions of different Nd3+-doped glasses are collected in Table 4.
The branching ratios of 4F3/24I9/2 and 4F3/24I11/2 transition in the PZLBNdYb glass was ~41% and 48%, respectively, which is comparable to BSKNLNd10 and PNbKA [36,37], PZLNNd1.0 [38], phosphate [14] and Nd3+-doped P2O5-Li2O3-GdF3 [29] glasses (see Table 4).
Table 4. Peak position (λ, nm), radiative transition probabilities (A, s−1), calculated branching ratios (β) and radiative lifetime (τ, ms) for 4F3/24I9/2 and 4F3/24I11/2 transitions of Nd3+-doped phosphate glasses.
Table 4. Peak position (λ, nm), radiative transition probabilities (A, s−1), calculated branching ratios (β) and radiative lifetime (τ, ms) for 4F3/24I9/2 and 4F3/24I11/2 transitions of Nd3+-doped phosphate glasses.
HostTransitionWavelength (nm)A (s−1)Τ (ms)Β (%)
PZLBNdYb [PW]4F3/24I9/2882253.241.64441
4F3/24I11/21051294.29 48
P2O5-Li2O3-GdF3-Nd2O3 [29] 4F3/24I9/287429170.15144
4F3/24I11/210653091 47
PZLNNd1.0 [38]4F3/24I9/2873398 41
4F3/24I11/21052342 53
3000 ppm Nd3+/Yb3+ in phosphate glasses [14] 4F3/24I9/2896403.80.7731
4F3/24I11/21056716.2 55
30,000 ppm Nd3+/Yb3+ in phosphate glasses [13] 4F3/24I9/2896280.20.1336
4F3/24I11/21056395.1 51
The proposed glasses have suitable spectroscopic quality factors, radiative lifetime and branching ratio values for lasing materials in the infrared region. In both glasses and crystals, efficient energy transfer between Nd3+ and Yb3+ ions has been established. The Nd3+ energy diagram, presented in Figure 4, shows the grouping of levels. The gap between gathering levels guarantees a large radiative probability of transition among both groups. By contrast, the small energy gap between the levels inside each group favors the multiphonon relaxation process.
According to Figure 4, the energy of the 4F3/2 emitting level of Nd3+ is located slightly higher than the 2F5/2 emitting level of Yb3+; thus, an Nd3+→Yb3+ energy transfer could take place via a phonon-assisted process (4F3/2, 2F7/24I9/2, 2F5/2 exothermic nonresonant transfer) and (4F3/2, 2F7/24I11/2, 2F5/2 endothermic nonresonant transfer). Furthermore, it was suggested that only a negligible resonant Yb3+→Nd3+ back transfer occurred. However, the phonon-assisted energy transfer from Nd3+ to Yb3+ as the way of quantum cutting is noticed. The energy of the 2P1/2(Nd3+) level (~23,250 cm−1) was approximately twice the energy of the Yb3+ transition (~10,290 cm−1) and the phonon-assisted energy transfer can be described as follows: Nd3+ emission: 2P1/2(Nd3+)→4I9/2(Nd3+); Yb3+ absorption: 2 2F7/2 (Yb3+)→2 2F5/2(Yb3+).
Normally, the 2F7/2 and 2F5/2 Stark levels of Yb3+ ions split into several sublevels due to the crystal field effect [36]. Here, the absorption spectrum was fitted by Lorentz fitting, shown in Figure 5.
From Figure 5, the spectrum fits two absorption bands attributed to transitions between the ground state of 2F7/2 and two Stark-splitting levels of 2F5/2. In principle, I974/I928 3; however, in our case, this ratio is reduced to ~1.5, which is related to the Nd3+→Yb3+ energy transfer.
The absorption and emission cross sections must be calculated to determine the lasing performance. The absorption cross section of a transition may be calculated as:
σ a b s ( λ ) = 2.303 N l × D ( λ )
where D(λ) is the optical density, l is the thickness of the sample, and N is the ion concentration in the sample. Furthermore, the emission cross section, σ e m ( λ ) , of Nd3+: 4F3/24I9/2 and Yb3+:2F5/22F7/2 can be calculated from the absorption cross section by:
σ e m ( λ ) = σ a b s ( λ ) · Z l Z u · e x p [ h c · ( k T ) 1 · [ ( λ Z L ) 1 ( λ ) 1 ] ]
where the lower and upper levels of the optical transition are Zl and Zu. T is the temperature, k is the Boltzmann constant, and ZZL is the wavelength at which the lower Stark sublevels of emitting multiplets and receiving multiplets intersect at this transition (zero-phonon line). However, at high temperatures, the partition function ratios of the lower and higher states Zl/Zu simply degenerate into a weighting of the two states. The precise Zl/Zu glass value is not known. The zero-phonon line was considered to be 882 nm for Nd3+ and 972 nm for Yb3+ in the following calculation, which assumes a Zl/Zu ratio of 10/4 for Nd3+ and 8/6 for Yb3+.
Figure 6 shows the Absorption and emission cross section of prepared glass PZLBNdYb. The computed σYb abs at (λp = 980 nm) of Yb3+:2F5/22F7/2 was 2.01 × 10−24 cm2, and that of Nd3+ for 4I9/24F3/2 at σNd absp = 882 nm) was 1.71 × 10−24 cm2. The σYb abs was larger than that of Nd3+ σNd abs in the NIR region, confirming that the Nd3+→Yb3+ dominates the Nd3+→Nd3+ energy transfer. Since the absorption cross section of Yb3+ is larger than that of Nd3+ at equal densities, it can be considered that Yb3+ acts as an acceptor, while Nd3+ is the donor in this NIR region. The values of σ e m ( λ ) of Nd3+: 4F3/24I9/2 and Yb3+:2F5/22F7/2 are equal to 2.23 × 10−24 cm2 and 2.88 × 10−24 cm2, respectively. The larger cross sections of the emission transitions indicate that the intense NIR emitting can be conditioned.
Figure 7a,b shows the measured MAC and LAC at 59.5, 622, 1170 and 1330 keV. The recorded values of MAC and LAC for the proposed sample at 662 keV were 0.081 ± 0.03 and 0.225 ± 0.13 in cm2/g, respectively. These values were better than commercially available glass shielding materials, such as RS253 and G18 [39]. The measured half-value layer, HVL; and mean free path, MFP; at the same specific energies 59.5, 622, 1170 and 1330 keV were obtained; Figure 7c,d. The values for HVL and MFP of the present glass at 662 keV were 2.20 ± 0.82 and 3.17 ± 1.18 cm, respectively.
The HVL signifies the material thickness that reduces the intensity of radiation to half. HVLs are reported for commercial materials, such as windows glass (4.73 cm), serpentine (4.07 cm), concrete (3.87 cm), SCHOOT glass RS253 (3.65 cm), hematite serpentine (3.6 cm), Ilmenite (2.63 cm) and SCHOOT glass RS323 (2.48 cm) [40,41]. Thus, our glass had a lower HVL, which is better than commercial material. It was correlated with codoped heavy compounds of Nd3+/Yb3+ ions in its structure of the host glass 40P2O5-30ZnO-20LiCl-10BaF2 to increase interaction probability, and more electrons are effectively available at low energy levels.

4. Conclusions

The incorporated double ions of Nd3+/Yb3+ into 40P2O5-30ZnO-20LiCl-10BaF2 glass increase the optical energy gap (4.41eV), which is a suitable active medium for laser glasses. We found that the JO intensity parameters of the produced glass follow the trend Ω6 > Ω4 > Ω2, which has a high spectroscopic quality factor (Ω46) equal to 0.84. This is larger than standard laser host Nd3+:YAG.
The fabricated glass had a large value of σ e m ( λ ) , 2.23 × 10−24 cm2, with the corresponding transition level Nd3+: 4F3/24I9/2 and was 2.88 × 10−24 cm2, attributable to Yb3+:2F5/22F7/2. Moreover, the gamma spectroscopic properties of the present glass showed a low half-value layer, which increases the interaction probability and creates more effective electrons at low energy. We conclude that the investigated glass has unique luminescence/gamma spectroscopic properties; hence, it can be used in photodynamic therapy surgery as a laser material in radiology rooms.

Author Contributions

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

Funding

This research was funded by the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia, for funding this research work through the project number IFP-KKU-2020/7.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research and Innovation, Ministry of Education, Saudi Arabia, for funding this research work through project number IFP-KKU-2020/7. The authors thank Ramzi Maâlej for review and scientific editing the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dorosz, D.; Żmojda, J.; Kochanowicz, M.; Dorosz, J. Nd3+/Yb3+ Doped Phosphate and Antimony Glasses for Optical Fibre Source. Acta Phys. Pol. A 2010, 118, 1108. [Google Scholar] [CrossRef]
  2. Afef, B.; Alqahtani, M.M.; Hegazy, H.H.; Yousef, E.; Damak, K.; Maâlej, R. Green and near-infrared emission of Er3+ doped PZS and PZC glasses. J. Lumin. 2018, 194, 706–712. [Google Scholar] [CrossRef]
  3. Lachheb, R.; Damak, K.; Assadi, A.A.; Herrmann, A.; Yousef, E.; Rüssel, C.; Maâlej, R. Characterization of Tm3+ doped TNZL glass laser material. J. Lumin. 2015, 161, 281–287. [Google Scholar] [CrossRef]
  4. Damak, K.; Yousef, E.; AlFaify, S.; Rüssel, C.; Maâlej, R. Raman, green and infrared emission cross-sections of Er3+ doped TZPPN tellurite glass. Opt. Mater. Express 2014, 4, 597–612. [Google Scholar] [CrossRef]
  5. Deguil, N.; Mottay, E.; Salin, F.; Legros, P.; Choquet, D. Novel diode-pumped infrared tunable laser system for multi-photon microscopy. Microsc. Res. Technol. 2004, 63, 23. [Google Scholar] [CrossRef]
  6. Singh, V.P.; Badiger, N.M.; Kaewkhao, J. Radiation shielding competence of silicate and borate heavy metal oxide glasses: Comparative study. J. Non-Cryst. Solids 2014, 404, 167–173. [Google Scholar] [CrossRef]
  7. Singh, K.; Singh, S.; Dhaliwal, A.S.; Singh, G. Gamma radiation shielding analysis of lead-flyash concretes. Appl. Radiat. Isot. 2015, 95, 174–179. [Google Scholar] [CrossRef]
  8. Tijani, S.A.; Kamal, S.M.; Al-Hadeethi, Y.; Arib, M.; Hussein, M.A.; Wageh, S.; Dim, L.A. Radiation shielding properties of transparent erbium zinc tellurite glass system determined at medical diagnostic energies. J. Alloy. Compd. 2018, 741, 293–299. [Google Scholar] [CrossRef]
  9. Qi, F.; Huang, F.; Wang, T.; Ye, R.; Lei, R.; Tian, Y.; Zhang, J.; Zhang, L.; Xu, S. Highly Er3+ doped fluorotellurite glass for 1.5 µm broadband amplification and 2.7 µm microchip laser applications. J. Lumin. 2018, 202, 132–135. [Google Scholar] [CrossRef]
  10. Hussein, K.I.; Alqahtani, M.S.; Alzahrani, K.J.; Alqahtani, F.F.; Zahran, H.Y.; Alshehri, A.M.; Yahia, I.S.; Reben, M.; Yousef, E.S. The Effect of ZnO, MgO, TiO2, and Na2O Modifiers on the Physical, Optical, and Radiation Shielding Properties of a TeTaNb Glass System. Materials 2022, 15, 1844. [Google Scholar] [CrossRef]
  11. Lalla, E.A.; Konstantinidis, M.; De Souza, I.; Daly, M.G.; Martín, I.R.; Lavín, V.; Rodríguez-Mendoza, U.R. Judd-Ofelt parameters of RE3+-doped fluorotellurite glass (RE3+ = Pr3+, Nd3+, Sm3+, Tb3+, Dy3+, Ho3+, Er3+, and Tm3+). J. Alloys Compd. 2020, 845, 156028. [Google Scholar] [CrossRef]
  12. Judd, B.R. Optical absorption intensities of rare-earth ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  13. Ofelt, G.S. Intensities of crystal spectra of rare−earth ions. J. Chem. Phys. 1962, 37, 511–520. [Google Scholar] [CrossRef]
  14. Zaman, F.; Srisittipokakun, N.; Rooh, G.; Khattak, S.A.; Singkiburin, N.; Kim, H.J.; Sangwaranatee, N.; Kaewkhao, J. Investigation of Li2O–Gd2O3–MO–B2O3–Nd2O3 (MO=Ba/Bi) glasses for laser applications by Judd–Oflet (J–O) theory. J. Lumin. 2019, 215, 116639. [Google Scholar] [CrossRef]
  15. Afef, B.; Hegazy, H.H.; Algarni, H.; Yang, Y.; Damak, K.; Yousef, E.; Maâlej, R. Spectroscopic analysis of trivalent Nd3+/Yb3+ ions codoped in PZS host glasses as a new laser material at 1.06 μm. J. Rare Earths 2017, 35, 361–367. [Google Scholar] [CrossRef]
  16. James, J.T.; Jose, J.K.; Manjunatha, M.; Suresh, K.; Madhu, A. Structural, luminescence and NMR studies on Nd3+-doped sodium–calcium-borate glasses for lasing applications. Ceram. Int. 2020, 46, 27099–27109. [Google Scholar] [CrossRef]
  17. 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]
  18. Siva Rama Krishna Reddy, 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]
  19. Elbashar, Y.H.; Rayan, D.A. Judd Ofelt Study of Absorption Spectrum for Neodymium Doped Borate Glass. Int. J. Appl. Chem. 2016, 12, 59–66. [Google Scholar]
  20. De Sa, G.F.; Malta, O.L.; de Mello Donegá, C.; Simas, A.M.; Longo, R.L.; Santa-Cruz, P.A.; da Silva, E.F., Jr. Spectroscopic properties and design of highly luminescent lanthanide coordination complexes. Coord. Chem. Rev. 2000, 196, 165–195. [Google Scholar] [CrossRef]
  21. Carnall, W.T.; Hessler, J.P.; Wagner, F., Jr. Transition probabilities in the absorption and fluorescence spectra of lanthanides in molten lithium nitrate-potassium nitrate eutectic. J. Phys. Chem. 1978, 82, 2152. [Google Scholar] [CrossRef]
  22. 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]
  23. Kaminskii, A.A. Laser Crystals: Their Physics and Properties; Springer: Berlin, Germany, 1981; pp. 157–158. [Google Scholar]
  24. Krupke, W.F. Induced-emission cross sections in neodymium laser glasses. IEEE J. Quant. Electron. 1971, 7, 153. [Google Scholar] [CrossRef]
  25. Campbell, J.H.; Suratwala, T.I.; Suratwala, T.I. Continuous melting of phosphate laser glasses. J. Non-Cryst. Solids 2000, 263–264, 342–357. [Google Scholar] [CrossRef]
  26. Jayasankar, C.K.; Balakrishnaiah, R.; Venkatramu, V.; Joshi, A.S.; Speghini, A.; Bettinelli, M. Luminescence characteristics of Nd3+ -doped K-Ba-Al-fluorophosphate laser glasses. J. Alloy Comp. 2008, 451, 697–701. [Google Scholar] [CrossRef]
  27. Gupta, G.; Sontakke, A.D.; Karmakar, P.; Biswas, K.; Balaji, S.; Saha, R.; Sen, R.; Annapurna, K. Influence of bismuth on structural, elastic and spectroscopic properties of Nd3+ doped zinc–boro-bismuthate glasses. J. Lumin. 2014, 149, 163–169. [Google Scholar] [CrossRef]
  28. Binnemans, K.; Van Deun, R.; Gorller-Walrand, C.; Adam, J.L. Spectroscopic properties of trivalent lanthanide ions in fuorophosphate glasses. J. Non-Cryst. Solids 1998, 238, 11–29. [Google Scholar] [CrossRef]
  29. Shoaib, M.; Rooh, G.; Chanthima, N.; Kim, H.J.; Kaewkhao, J. Luminescence properties of Nd3+ ions doped P2O5-Li2O3-GdF3 glasses for laser applications. Opt.-Int. J. Light Electron. Opt. 2019, 199, 163218. [Google Scholar] [CrossRef]
  30. Prasad, V.R.; Dhamodaraiah, S.; Babu, S.; Ratnakaram, Y.C. Spectroscopic investigation of Nd3+ doped zinc phosphate glasses for NIR emission at 1.05 μ-m. Mater Today Proc. 2016, 3, 3805–3809. [Google Scholar] [CrossRef]
  31. Jorgensen, C.K.; Reisfeld, R. Judd-Ofelt parameters and chemical bonding. J. Common Met. 1983, 93, 107–112. [Google Scholar] [CrossRef]
  32. Reisfeld, R. Luminescence and prediction of transition probabilities for solar energy and lasers. J. Common. Met. 1985, 112, 9–18. [Google Scholar] [CrossRef]
  33. Jacobs, R.; Weber, M. Dependence of the 4F3/24I11/2 induced-emission cross section for Nd3+ on glass composition. IEEE J. Quant. Electron. 1976, 12, 102–111. [Google Scholar] [CrossRef]
  34. Riseberg, L.A.; Weber, M.J. Relaxation phenomena in rare-earth luminescence. Prog. Opt. 1977, 31, 89–159. [Google Scholar]
  35. Fares, H.; Jlassi, I.; Hraiech, S.; Elhouichet, H.; Ferid, M. Radiative parameters of Nd3+-doped titanium and tungsten modified tellurite glasses for 1.06 μm laser materials. J. Quant. Spectrosc. Radiat. Transf. 2014, 147, 224–232. [Google Scholar] [CrossRef]
  36. Manasa, P.; Srihari, T.; Basavapoornima, C.; Joshi, A.S.; Jayasankara, C.K. Spectroscopic investigations of Nd3+ ions in niobium phosphate glasses for laser applications. J. Lumin. 2019, 211, 233–242. [Google Scholar] [CrossRef]
  37. Prasad, R.N.A.; Praveena, R.; Vijaya, N.; Babu, P.; Krishna Mohan, N. Neodymiumdoped magnesium phosphate glasses for NIR laser applications at 1.05 μm. Mater. Res. Express 2019, 6, 096204. [Google Scholar]
  38. Ramteke, D.D.; Kroon, R.E.; Swart, H.C. Infrared emission spectroscopy and upconversion of ZnO-Li2O-Na2O-P2O5 glasses doped with Nd3+ ions. J. Non-Cryst. Solids 2017, 457, 157–163. [Google Scholar] [CrossRef]
  39. Chanthima, N.; Kaewkhao, J. Investigation of radiation shielding parameters of bismuth borosilicate glass from 1 keV to 100 GeV. Ann. Nucl. Energy 2013, 55, 23–28. [Google Scholar] [CrossRef]
  40. Chanthima, N.; Kaewkhao, J.; Limkitjaroenporn, P.; Tuscharoen, S.; Kothan, S.; Tungjai, M.; Kaewjaeng, S.; Sarachai, S.; Limsuwan, P. Development of BaO–ZnO–B2O3 glasses as a radiation shielding material. Radiat. Phys. Chem. 2017, 137, 72–77. [Google Scholar] [CrossRef]
  41. Cheewasukhanont, W.; Limkitjaroenporn, P.; Sayyed, M.I.; Kothan, S.; Kim, H.J.; Kaewkhao, J. High density of tungsten gadolinium borate glasses for radiation shielding material: Effect of WO3 concentration. Radiat. Phys. Chem. 2022, 192, 109926. [Google Scholar] [CrossRef]
Figure 1. The experimental setup used for measuring the shielding parameters for the proposed sample [10].
Figure 1. The experimental setup used for measuring the shielding parameters for the proposed sample [10].
Photonics 09 00406 g001
Figure 2. Absorption spectra of PZLBNdYb glass. The upper transition state is identified for each peak.
Figure 2. Absorption spectra of PZLBNdYb glass. The upper transition state is identified for each peak.
Photonics 09 00406 g002
Figure 3. Relation between (αhν) and () of the prepared glass PZLBNdYb.
Figure 3. Relation between (αhν) and () of the prepared glass PZLBNdYb.
Photonics 09 00406 g003
Figure 4. Energy level diagrams of Nd3+ and Yb3+ ions.
Figure 4. Energy level diagrams of Nd3+ and Yb3+ ions.
Photonics 09 00406 g004
Figure 5. Deconvolution of the Yb3+ absorption peak.
Figure 5. Deconvolution of the Yb3+ absorption peak.
Photonics 09 00406 g005
Figure 6. Absorption and emission cross section of prepared glass PZLBNdYb.
Figure 6. Absorption and emission cross section of prepared glass PZLBNdYb.
Photonics 09 00406 g006
Figure 7. The measured and theoretical shielding parameters for TeTaNbZn glass samples at 59.5, 622, 1170 and 1330 keV: (a) MAC; (b) LAC; (c) HVL; and (d) MFP.
Figure 7. The measured and theoretical shielding parameters for TeTaNbZn glass samples at 59.5, 622, 1170 and 1330 keV: (a) MAC; (b) LAC; (c) HVL; and (d) MFP.
Photonics 09 00406 g007
Table 1. The results of the intensity measurements and line strength calculations for absorption transitions of Nd3+ doped in PZLBNdYb glass, δrms = 1.019 × 10−25 m2.
Table 1. The results of the intensity measurements and line strength calculations for absorption transitions of Nd3+ doped in PZLBNdYb glass, δrms = 1.019 × 10−25 m2.
Transitions: 4I9/2λ (nm)ν (cm−1)||U2||2||U4||2||U6||2Γ
(nm·cm−1)
Sexp
(10−26 m2)
Scal
(10−26 m2)
4D3/2 + 4D5/2 + 2I11/2 + 4D1/235228,4090.00500.52560.04787.851124.48922.244
2P1/2 + 2D5/243023,2550.00000.03690.00210.51171.29931.4960
2K15/2 + 2G9/2 + 2D3/2 + 4G11/247021,2760.00100.04720.03643.15057.31693.4488
2K13/2 + 4G7/2 +4G9/252419,0830.06650.21820.127116.2234.05315.772
4G5/2 + 2G7/2 + 2H11/258017,2410.97370.59680.077727.29851.29452.438
4F9/268214,6620.00090.00920.04172.60654.17852.2393
4F7/2 + 4S3/274413,4400.00100.0450.659825.62437.60031.252
4F5/2 + 2H9/280212,4680.01020.24510.512714.85320.26932.523
4F3/288211,4670.00000.22930.05482.09652.609811.164
Table 2. Comparison of Judd–Ofelt parameters (Ωt, ×10−20 cm2) of PZLBPr glass along with other systems.
Table 2. Comparison of Judd–Ofelt parameters (Ωt, ×10−20 cm2) of PZLBPr glass along with other systems.
SystemsΩ2Ω4Ω6Trend χ
PZLBNdYb [Present Work]:0.2690.3790.447Ω2 < Ω4 < Ω60.84
LGBaBNd05 [12]6.10 6.859.83Ω2 < Ω4 < Ω60.69
YAG:Nd3+ [24]0.20 2.70 5.00 Ω2 < Ω4 < Ω60.54
LHG-750 [25] 4.60 4.80 5.60 Ω2 < Ω4 < Ω60.85
PKFBAN10 [26] 4.923.67 5.26Ω4 < Ω2 < Ω60.70
ZnBBi [27] 3.56 4.30 4.87 Ω2 < Ω4 < Ω60.88
Nd3+:fluorotellurite(n = cst) [11]4.215.975.45Ω2 < Ω6 < Ω41.09
Nd3+:fluorotellurite(n ≠ cst) [11]4.51 6.346.16Ω2 < Ω6 < Ω41.02
Nd3+ doped Y2O3 ceramic [22]8.84 9.824.44Ω6 < Ω2 < Ω42.21
75NaPO3-24CaF2-1NdF3 [28]2.78 4.165.56Ω2 < Ω4 < Ω60.74
3000 ppm Nd3+/Yb3+ in phosphate [13]1.8970.8201.834Ω4 < Ω6 < Ω20.44
30,000 ppm Nd3+/Yb3+ in phosphate [13]0.23390.64370.9598Ω2 < Ω4 < Ω60.67
NCB:Nd glasses [14] 1.50 0.932.39Ω4 < Ω2 < Ω60.39
PMZ1.5 Nd [15]4.69 4.72 2.98Ω6 < Ω2 < Ω41.58
AEBTNd0.1 [16]3.694 2.865 2.548Ω2 < Ω4 < Ω61.12
P2O5-Li2O3-GdF3-Nd2O3 [29]8.55 11.5410.25Ω2 < Ω6 < Ω41.13
Nd3+ doped zinc phosphate [30]4.67 5.53 5.77 Ω2 < Ω4 < Ω60.95
Table 3. Spectroscopic parameters of the PZLBNdYb glass system.
Table 3. Spectroscopic parameters of the PZLBNdYb glass system.
TransitionEnergy (cm−1)A (s−1)Τ (ms)Β (%)
4F3/24I9/2882253.241.64441.639
4F3/24I11/21051294.29 48.389
4F3/24I13/2132957.675 9.4834
4F3/24I15/218272.9709 0.4884
4I15/24I9/216261.6995118.7820.187
4I15/24I11/224784.5382 53.905
4I15/24I13/248782.1812 25.909
4I13/24I9/224405.935126.4275.028
4I13/24I11/250401.9754 24.972
4I11/24I9/247302.5817387.34100
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Charfi, B.; Damak, K.; Alqahtani, M.S.; Hussein, K.I.; Alshehri, A.M.; Elkhoshkhany, N.; Assiri, A.L.; Alshehri, K.F.; Reben, M.; Yousef, E.S. Luminescence and Gamma Spectroscopy of Phosphate Glass Doped with Nd3+/Yb3+ and Their Multifunctional Applications. Photonics 2022, 9, 406. https://doi.org/10.3390/photonics9060406

AMA Style

Charfi B, Damak K, Alqahtani MS, Hussein KI, Alshehri AM, Elkhoshkhany N, Assiri AL, Alshehri KF, Reben M, Yousef ES. Luminescence and Gamma Spectroscopy of Phosphate Glass Doped with Nd3+/Yb3+ and Their Multifunctional Applications. Photonics. 2022; 9(6):406. https://doi.org/10.3390/photonics9060406

Chicago/Turabian Style

Charfi, Bilel, Kamel Damak, Mohammed S. Alqahtani, Khalid I. Hussein, Ali M. Alshehri, Nehal Elkhoshkhany, Abdullah L. Assiri, Khaled F. Alshehri, Manuela Reben, and El Sayed Yousef. 2022. "Luminescence and Gamma Spectroscopy of Phosphate Glass Doped with Nd3+/Yb3+ and Their Multifunctional Applications" Photonics 9, no. 6: 406. https://doi.org/10.3390/photonics9060406

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