Next Article in Journal
Coaxial Electrospinning of PCL-PVA Membranes Loaded with N-Heterocyclic Gold Complex for Antitumoral Applications
Previous Article in Journal
Waste Bombyx Mori Silk Textiles as Efficient and Reuseable Bio-Adsorbents for Methylene Blue Dye Removal and Oil–Water Separation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal, Optical, and Emission Traits of SM3+-Ion-Doped Fluoride/Chloride/Oxide Glass for Red/Orange Laser Fiber Applications

1
Institute of Physics, Cracow University of Technology, ul. Pochorazych 1, 30-084 Cracow, Poland
2
Institute of Low Temperatures and Structure Research, Polish Academy of Sciences, ul. Okolna 2, 50-950 Wroclaw, Poland
3
Faculty of Materials Science and Ceramics, AGH-University of Krakow, al. A. Mickiewicza 30, 30-059 Cracow, Poland
4
Centre for Polymer and Carbon Materials, Polish Academy of Sciences, ul. Marii Curie-Skłodowskiej 34, 41-819 Zabrze, Poland
5
Department of Physics, Faculty of Science, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
6
Department of Radiological Sciences, College of Applied Medical Sciences, King Khalid University, Abha 61421, Saudi Arabia
*
Author to whom correspondence should be addressed.
Fibers 2024, 12(11), 100; https://doi.org/10.3390/fib12110100
Submission received: 22 July 2024 / Revised: 14 October 2024 / Accepted: 11 November 2024 / Published: 15 November 2024

Abstract

:
This study examined spectroscopic, thermal, and other qualities, such as the lasing parameters, of Sm3+-doped glass with the composition 40P2O5–30ZnO–20LiCl–10BaF2. The ellipsometric data were used in a Sellmeier dispersion relation to estimate the refractive index values of the glasses investigated. The measured absorption spectra of the doped glass reveal the presence of various absorption bands assigned to transitions from the 6H5/2 ground state attributed to Sm3+-ion-excited states. We studied the decay of the 4G5/2 level of the Sm3+ ions in the doped glass by analyzing its absorption and emission fluorescence spectra. The Judd–Ofelt hypothesis allowed us to determine that the quantum efficiency of the 4G5/26H7/2 transition is high: 96% and 97% for glass doped with 4.05 × 1019 ions/cm−3 and 11 × 1019 ions/cm−3, respectively. Furthermore, this glass exhibits efficient red/orange enhanced spontaneous emission that matches the excitation band of the photosensitizer material used in medical applications.

1. Introduction

Phosphate glasses possess high thermal stability, relatively low melting temperatures, and unique optical properties, along with a broad range of transparency in the UV-VIS spectral range; however, they have a low refractive index [1]. Therefore, phosphate glasses doped with rare-earth (RE) ions are widely studied for their applications in waveguides, optical detectors, fiber optic amplifiers, and lasers [2]. The addition of various fluorides to RE-ion-doped phosphate glasses enables the creation of host matrices that combine the significant optical properties of fluoride glasses and the high stability of phosphate glasses. Oxides increase the mechanical resistance and thermal stability of the glasses, while low-frequency fluorides reduce the nonradiative decay losses caused by multifamily relaxation [3], improve the quantum efficiency and lifetimes of rare-earth ions’ luminescence, and act as flux agents that reduce the melting temperature of the glasses. When samarium ions are doped into fluorophosphate glass, the resulting glass exhibits distinctive optical behavior due to its emissions in the UV–visible region through various emission channels from the 4G5/2 level of Sm3+ ions [4]. Samarium ions (Sm3+) are regarded as excellent luminescent centers among the rare-earth (RE3+) ions due to their strong emissions at longer wavelengths in the red/orange region. The strong luminescence in the red/orange region, as well as the high quantum efficiency value of Sm3+-doped P2O5–Na2O–BaO–Al2O3 glasses, makes them a potential candidate for laser emission [5].
This characteristic makes them highly suitable for applications in solid-state lasers, LEDs, and display devices. Additionally, the lowest emitting level of Sm3+ ions, 4G5/2, exhibits high quantum efficiency and various quenching emission channels [6,7]. The 4f5 configuration of samarium ions with strong orange-red fluorescence in the visible region is used in underwater communications, color displays, visible semiconductor lasers [8,9,10], and phosphors for white light-emitting diodes (LEDs) [11,12]. Although extensive research has been conducted on phosphate glasses doped with Sm3+ ions [13,14,15,16], there is still a need to explore new and advanced compositions of phosphate glass due to their potential applications.
In [17], the researchers confirmed the formation of defects and the conversion of Sm3+ into Sm2+ ions in Sm-doped fluorophosphate glasses exposed to X-ray irradiation. Additionally, Hamdy et al. [18] synthesized compositions containing 20, 30, and 50 mol% NaF, AlF3, and PF5 doped with Sm3+ to investigate the impact of Sm3+ on optical transmission and photoluminescence (PL) spectroscopy. The authors of [19] discussed the influence of the chemical composition of [(45 − x) P2O5–10AlF3–45 NaF x Sm2O3](where x equal 0 to 1 mol%) glasses on their optical properties, mainly their linear and non-linear parameters. The study performed by [20] focused on examining the thermal, structural, and luminescent characteristics of sodium barium metaphosphate glasses doped with Sm3+. Samarium-doped fluoroaluminate and fluorophosphate glasses studied by Chicilo et al. [21] proved to be excellent candidates for high-resolution, large-dynamic-range microbeam radiation therapy (MRT) dosimetry. It is worth noting that Sm3+ ions are ideal for doping because their excited energy level 4G5/2 has a high quantum efficiency with various quenching emission channels. It emits strong reddish-orange light and possesses sufficient energy to initiate photodynamic reactions. Recently, lasers with excellent directivity and high intensity, as well as LEDs with relatively narrow spectral bandwidths and high fluence rates, have been specifically developed for photodynamic therapy (PDT) treatments. However, scattered or misdirected light from the target area can unintentionally expose large areas of normal tissue to high power densities, potentially leading to side effects such as inflammation, pain, swelling, burns, and scarring. While LEDs can be integrated with optical fibers, their low coupling efficiency limits their use. Additionally, their narrow excitation spectrum restricts the use of multiple photosensitizers, reducing the efficiency of therapy. Amplified spontaneous emission (ASE) fluorescence produced in rare-earth (RE) ion-doped glass channel waveguides offers a broadened bandwidth that can be tuned by adjusting the RE ion concentration and pumping power. This type of light source delivers sufficient intensity, excellent directivity, and high coupling efficiency, making it a promising option for use in photodynamic therapy (PDT) treatments [22].
A fluorophosphate glass channel waveguide produces amplified spontaneous emission (ASE) fluorescence with a broad width of 600–730 nm. This light source has excellent efficiency, strong intensity, and good directivity, thus making it an attractive option for use in photodynamic therapy (PDT) [22,23]. Red light in the 600–730 nm spectral band has 50–200% greater penetrating power than light in the 400–500 nm region, and it possesses more energy to initiate a photodynamic reaction that produces 1O2 according to most tissue models [22]. A diverse range of light sources, both coherent and incoherent, have demonstrated their effectiveness in achieving anti-tumor effects in PDT for various superficial and interstitial treatment sites [24,25,26]. Thus, in this paper, we have focused on the lasing parameters of Sm3+-doped fluorophosphate glasses as a candidate light source for photodynamic therapy (PDT). In the glass compositions, we included BaF2 ions as network modifier ions with a low field strength and ZnO to improve the mechanical strength, chemical durability, and hygroscopic nature, which, in turn alter, the optical, electrical, and magnetic properties of phosphate glasses.

2. The Experimental Section

The composition of the glasses was chosen as 40P2O5–30ZnO–20LiCl–10BaF2 with different concentrations of Sm3+ ions: sample 1 (SM1) at 4.05 × 10 19 cm−3 and sample 2 (SM2) at 11 × 1019 cm−3. The following chemicals were used for batch production: phosphorus oxide (P2O5), zinc oxide (ZnO), lithium chloride (LiCl), barium fluoride (BaF2), and samarium oxide (Sm2O3). All of the raw materials were carefully mixed. The fluorophosphate glasses were obtained by melting 50 g batches in a gold crucible in an electric furnace at a temperature of 850°C in an air atmosphere [27]. The densities of the glasses studied were determined using Archimedes’ method (Table 1).
Their refractive indexes were determined with use of a Woollam M-2000 ellipsometer; the spectra were recorded in the 190–1700 nm spectral range. The transmittance and reflectance spectra were recorded using Jasco V-570 spectrophotometers [27]. Luminescence spectra were measured using the Optron Dong Woo fluorometer system [28]. The luminescence decay curves were measured following short-pulse excitation provided by an optical parametric oscillator with the third harmonic of a Nd YAG laser [29,30]. Changes in the thermal behavior of the 40P2O5–30ZnO–20LiCl–10BaF2 glasses were investigated with the DTA/DSC method using the PerkinElmer DTA-7 System.
Under a flowing air environment (80 mL min−1), the glass samples were ground into powders with grain sizes of 0.1 to 0.3 mm and then heated in platinum crucibles at a rate of 10 °C min−1.
Alumina oxide (Al2O3) was used as a reference material. Several thermal parameters characteristic of the glassy state have been established, i.e., the glass transformation temperature (Tg) and the glass crystallization temperature (Tc). Using the midpoint of the corresponding transformation phase as the glass conversion temperature (Tg) and the start and maximum of the glass crystallization peak as the parameters for the glass melting and solidification processes, we were able to obtain the glass heating and solidification temperatures.
While the midpoint of the corresponding transformation step was used to estimate the glass conversion temperature (Tg), the start (Tx) and maximum (Tc) of the glass crystallization peak were used to calculate the glass crystallization temperature.
The characteristic glass temperature values that appeared in DSC were determined using Proteus Thermal Analysis (Version 5.0.0.). The amorphous nature of the glasses studied was confirmed using the X-ray diffraction (XRD) method and a Philips X’Pert X-ray diffractometer with CuKα radiation. All of the research was conducted at room temperature. The refractive index of the bulk glasses was investigated using transmission mode ellipsometric data. The ellipsometric results were registered in the range of 400–1700 nm, and n was modeled using the Sellmeier dispersion function. Therefore, the Sellmeier refractive index (n)’s normal dispersion, which was determined from our ellipsometric data collected as a function of the wavelength λ (see Equation (1)) [31,32], was fitted to the low-absorption region, which is 400–1700 nm.
n2 = A + Bλ2/(λ2 − C2) − Dλ2
where A, B, C, and D are the fitting parameters.
The refractive index dispersions of the glasses studied are quite similar to each other, as are the values of n measured at 633 nm (see Table 1). The addition of Sm2O3 to the glasses led to a decrease in the refractive index compared to the sample SM without a dopant.
It is worth mentioning that ellipsometry is not a direct method for determining the refractive index.

3. Results and Discussion

3.1. Thermal Analysis

The broad hump in the XRD profile shows the long-term structural instability (i.e., random atomic arrangements) of the glasses prepared (Figure 1). Therefore, based on XRD analysis of all the glasses, it can be found that they are amorphous. The differential scanning calorimetry (DSC) curves of all of the glasses studied showed distinct endothermic events associated with the glass transition Tg, as well as the melting point Tm. The exothermic event Tc associated with crystallization was observed. During heating, the glasses studied demonstrate several characteristic temperatures. The range of vitreous state transitions (Tg,endset − Tg,onset) decreases progressively when the Sm3+ ion concentration in the glass increases, although the transition temperature increases (Figure 2).
The increase in Tg refers to structural changes in the glass network with increasing Sm3+; furthermore, the narrowing of the transition range of the vitreous state (i.e., a reduction in the relaxation time) supports the theory of increased rigidity in the material’s internal structure (Table 2). According to [33], this process occurs via the breaking of chemical bonds rather than the displacement of structural units when the time required for structural strain relaxation shortens, maintaining the network’s continuity. To determine the thermal behavior of glasses, many parameters have been established, mainly considering the dependencies between values of characteristic temperatures. To evaluate the stability of the glass, the following parameters were utilized: the Hruby (KH) = (Tx − Tg)/(Tm − Tx) [34], the Weinberg (KW) = (Tx − Tg)/Tm), the Lu and Liu (KLL) = Tx/(Tg + Tm)), and the Angell KA = (Tx − Tg) parameters [35]. The resistance of the resulting material to crystallization is expressed by the term (Tx − Tg) in the suggested formula; a narrower gap indicates an increased tendency to crystallize. As the concentration of Sm3+ ions in the glasses increased, their KA was in the 69–73 °C temperature range. Greater Sm3+ doping results in a higher KA value, increasing the glass’s thermal stability.
The stronger thermal stability of glass against devitrification on heating was further verified by the fact that the values of other thermal stability metrics, such as Kw, KH, and KLL, increased with larger Sm3+ concentrations. Not only can one infer a material’s stability from all these variables but one can also infer its glass-forming ability, which is the ease with which a liquid vitrifies upon cooling. Table 3 shows that the stability and glass-forming ability of 40P2O5–30ZnO–20LiCl–10BaF2 glasses are enhanced when the Sm3+ ion concentration in their structure increases. Both Mawlud et al. and Mnjeet et al. [36,37] found similar correlations in their studies.

3.2. Absorption and Excitation Spectra

The transmittance and reflectance spectra allowed us to calculate the dispersion of the absorption coefficient of the fluorophosphate glasses doped with the Sm3+ ions. The results are illustrated in Figure 3 and Figure 4.
As shown in Figure 3 and Figure 4, the absorption spectra of the Sm3+-doped fluorophosphate glasses registered in the UV-VIS-NIR region showed 19 peaks. As can be seen in Figure 3 and Figure 4, these peaks are caused by Sm3+ ions’ transition from the 6H5/2 ground state to various energy levels [38,39].
The absorption spectra allowed us to observe a large number of energy transitions of the samarium ions in the glass from the system P2O5–ZnO–LiCl–BaF2. Above 450–500 nm, in the SM1 and SM2 glasses, one may find the high energy manifolds built from the 4D, 4G, 4I, 4L, and 4M quartets and 6P sextet terms. Because many 2S+1LJ splitters overlap, the UV-VIS bands that are visible are less intense and difficult to identify. Shifts in the absorption edge towards the UV region are observed with an increasing Sm3+ ion content. However, the bands that formed in the near-infrared spectrum are more distinct and strongly differentiated. In the spectral range of 500–250 nm, there is a comparable effect; however, in contrast, intense absorption peaks at 402 nm occurred, and this band is associated with transitions ending in 4M19/2, 6P3/2, 4L15/2, 6P7/2, 4D3/2, 4D7/2, 3P3/2, and 4P5/2 closely spaced multiples.
In addition, the absorption spectra of the glasses tested were supplemented by photoluminescent excitation spectra, which are shown in Figure 5.
Figure 6 shows the absorption cross-section (ACS) of the 6H5/26F7/2 transition, which is fundamental for samarium-doped fiber amplifiers. The ACS is defined as σabs = α/N and is easily obtained from the absorption spectra shown in Figure 3.
The most significant transitions of Sm3+ ions occur in the range of 1000–1300 nm, corresponding to the 6H5/26 F9/2 and 6H5/26F7/2 transitions of Sm3+ ions (Figure 6).
Figure 7 presents the absorption cross-section (σabs) calculated for the SM1 and SM2 glasses, and the maximum value at 402 nm was σabs = 17.0067 × 10−21 cm2 for SM2 and σabs = 14.1206 × 10−21 cm2 for SM1.

3.3. Judd–Ofelt Analysis

Here, we analyzed the absorption data using the conventional Judd–Ofelt (J-–O) theory (Figure 3 and Figure 4). Numerical integration of suitable absorption bands, excluding the background absorption of the glass arrays, was used to establish the experimental oscillator strengths for transitions from the Sm3+ ions’ ground level 6H5/2 to subsequent excited levels.
Many intense transitions can be observed in the PLE spectrum (Figure 5), like those in the absorption spectra. In Table 4, the computed J–O parameters (Ω2, Ω4, and Ω6) can be observed. Judd–Ofelt parameters are a useful tool for determining the radiative properties of Sm3+-doped glasses, which are strongly dependent on the host matrix. The values of the J–O parameters obtained [40,41] are comparable to values found in other phosphate glasses doped with Sm3+ and show the same trend, namely Ω4 > Ω6 > Ω2 [42,43]. The values of Ω4 and Ω6 are associated with structural qualities like the viscosity and stiffness of the material, whereas the value of Ω2 indicates the covalency of oxygen atoms and the asymmetry of the RE ion sites, according to J–O theory. The ion sites exhibit less asymmetry, leading to dominating covalent interactions between oxygen ligands and Sm3+ ions, as shown by the high value of Ω4 compared to Ω2 and Ω6. We may find out more about the luminescence activators by calculating the spectroscopic quality factor using the Ω46 ratio. The glasses examined show promise as luminescence activators, with computed spectroscopic quality factor values of 1.1 (SM1) and 1.5 (SM2), respectively [44,45].
Three phenomenological spectroscopic parameters were calculated. In addition, the value of fitting quality (σrms) esti mated from the root mean square (RMS) deviation [46] was determined and is presented in Table 4. This value implies a good match between the measured Pexp and the calculated Pcal oscillator strengths.
The radiative transition probability (Wr) and the total radiative lifetime (τrad = 1/Wr) of the emission level 4G5/2 were estimated using the J-O parameters determined. In the conventional J-O approach, the Ωt parameter values are derived using an averaging process that takes into account all the absorption bands, known as the least-squares method. The samples’ observed lifetimes of the excited state 4G5/2 (Figure 8a,b) were considerably lower than the expected radiation equivalent of τrad (refer to Table 5) and those predicted by the following relationship [42]:
η = τ e x p τ r a d × 100 %
This gives more than 96% of the quantum efficiency of the excited state for the material under study.
In the case of samples SM1 and SM2, one deals with high-concentration Sm3+ ions, leading to effective cross-relaxation (CR) between them [47]. The CR mechanism causes a decrease in the Sm3+:4G5/2 lifetime for SM1 to 3.134 ms when compared to the J-O radiative lifetime of 3.856 ms; for sample SM2, the lifetime decreases to τexp = 3.758 ms, compared to a radiative lifetime of 4.865 ms.

3.4. Emission Spectra

The photoluminescence (PL) emission spectra of Sm3+-doped fluorophosphate glasses are shown in Figure 9c, with images of the glass emission presented in Figure 9a,b.
Figure 9c displays the photoluminescence spectra produced by the glasses stimulated at 402 nm. The emission bands seen at 561.5 nm, 597.5 nm, 643.5 nm, and 708 nm, respectively, are caused by the 4G5/2  6H5/2, 4G5/2  6H7/2, 4G5/2  6H9/2, and 4G5/2  6H11/2 transitions of Sm3+. The strength of the green and reddish-orange emission bands is significantly affected by the kind of glass host. 4G5/2  6H7/2 and 4G5/2  6H9/2 (red/orange) are the two distinct transition peaks of the Sm3+ ions in the visible emission spectrum. A decrease in the higher energy levels of 4G5/2 may occur when higher sub-levels of energy are near a Sm3+ ion in the ground state because the energy can be absorbed by the Sm3+ ion.
The glasses tested, like glasses in other research, show that the exceptional intensity of the 4G5/26H7/2 transition makes them ideal for various applications, including color displays, high-density optical storage devices, and medical diagnostics. Our data closely align with those on other glasses described in the referenced studies [48,49,50].
Figure 10 shows a simplified diagram of the emission energy of Sm3+. From SM1 and SM2, it can be seen that as the Sm3+ ions are pumped with a 402 nm excitation wavelength, they are excited to the 6P3/2 level, and then some of the Sm3+ ions relax non-radioactively into lower levels of 4G5/2 and 4F3/2 and then decay into 6H9/2.
Figure 11 displays a CIE 1931 chromaticity diagram, which illustrates the effect of the emission transitions at a 402 nm excitation wavelength. For the manufactured glasses SM1 and SM2, the color coordinates (x, y) are (0.551, 0.352) and (0.581, 0.398), respectively. Thus, the glasses produced in this study have the potential to serve as materials for optical gain in orange/red laser applications.

4. Conclusions

In this study, glasses with a base chemical composition of 40P2O5–30ZnO–20LiCl–10BaF2 doped with Sm3+ were successfully prepared using a melt quenching technique. The glasses obtained fulfil the criteria to be considered superior materials for channel waveguide applications. They are compositionally and structurally homogeneous to prevent scattering and ensure consistent guiding properties and are free from significant defects and impurities that could introduce additional loss or alter the refractive index profile. From the DCS curve profiles, the glass transition temperature (Tg), the onset of the glass crystallization peak (Tx), the maximum glass crystallization peak (Tc), glass melting temperature (Tm), and thermal stability parameters were analyzed. The range of the vitreous state transition (Tg_endset − Tg_onset) was observed to decrease, while the glass transition temperature shifted towards higher temperatures with an increase in the Sm3+ ion concentration. The values of the thermal stability parameters calculated (Kw, KH, KLL) increased with a greater Sm3+ content, confirming the higher thermal stability of glass against devitrification on heating. Based on thermal studies, we confirmed very good glass stability. The glasses under investigation showed normal refractive index values for fluorophosphate glasses, according to ellipsometric measurements, and the Sellmeier model provided an excellent description of this index’s dependency on wavelength. The Judd–Ofelt parameters calculated for the 40P2O5–30ZnO–20LiCl–10BaF2 glasses are as follows: SM1 (Ω2 = 0.4327 × 10−20 cm2, Ω4 = 1.1224 × 10−20 cm2, Ω6 = 1.0071 × 10−20 cm2) and SM2 (Ω2 = 0.9897 × 10−20 cm2, Ω4 = 1.9765 × 10−20 cm2, Ω6 = 1.4576 × 10−20 cm2). The large quantum efficiency of Sm3+/SM1/4G5/2 (96%) and SM2 (97%) was reported for all samples. The absorption cross-section data for the 4G5/26H7/2 transition were determined. The CIE chromaticity coordinate (x,y) values corresponded to the orange-red region at the concentrations studied. The prominent transition peaks of the Sm3+ ions in the visible emission spectrum wavelength (red/orange) make our glasses useful as laser fiber sources that could be used in photodynamic therapy (PDT) treatment.

Author Contributions

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

Funding

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

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors extend their appreciation to the Ministry of Education in the KSA for funding this research work through project number KKU-IFP2-DA-6.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Silva, G.H.; Anjos, V.; Bell, M.J.V.; Carmo, A.P.; Pinheiro, A.S.; Dantas, N.O. Eu3+ emission in phosphate glasses with high UV transparency. Lumin J. 2014, 154, 294–297. [Google Scholar] [CrossRef]
  2. Ravi Babu, Y.N.C.; Sree Ram Naik, P.; Vijaya Kumar, K.; Rajesh Kumar, N.; Suresh Kumar, A. Spectral investigations of Sm3+ doped lead bismuth magnesium borophosphate glasses. J. Quant. Spectrosc. Radiat. Transf. 2012, 113, 1669–1675. [Google Scholar] [CrossRef]
  3. Jayachandiran, M.; Masilla Moses Kennedy, S. Synthesis and optical properties of Ba3Bi2 (PO4)4: Dy3+ phosphors for white light emitting diodes. J. Alloys Compd. 2019, 775, 353–359. [Google Scholar] [CrossRef]
  4. Annapurna, K.; Das, M.; Kundu, M.; Dwivendhi, R.N.; Budduhudu, S. Spectral properties of Eu3+: ZnO-B2O3-SiO2. J. Mol. Struct. 2005, 741, 53–60. [Google Scholar] [CrossRef]
  5. Bayoudhi, D.; Bouzidi, C.; Matei, E.; Secu, M.; Galca, A.C. Optical characterization of Sm3+ doped phosphate glasses for potential orange laser applications. J. Lumin. 2024, 265, 120204. [Google Scholar] [CrossRef]
  6. Rao, C.S.; Jayasankar, C.K. Spectroscopic and radiative properties of Sm3+-doped K–Mg–Al phosphate glasses. Opt. Commum. 2013, 286, 204–210. [Google Scholar] [CrossRef]
  7. Kesavulu, C.R.; Jayasankar, C.K. Spectroscopic properties of Sm3+ ions in lead fluorophosphate glasses. J. Lumin. 2012, 132, 2802–2809. [Google Scholar] [CrossRef]
  8. Mahato, K.K.; Rai, D.K.; Rasi, S.B. Optical studies of Sm3+ doped oxyfluoroborate glass. Solid State Commun. 1998, 108, 671–676. [Google Scholar] [CrossRef]
  9. Swapna, K.; Mahamuda, S.; Srinivasa Rao, A.; Sasikala, T.; Rama Moorthy, L. Visible luminescnence characteristics of Sm3+ doped zinc alumino bismuth borate glasses. J. Lumin. 2014, 146, 288–294. [Google Scholar] [CrossRef]
  10. Souza Filho, A.G.; Mendes Filho, J.; Melo, F.E.A.; Custodio, M.C.C.; Lebullenger, R.; Hern, A.C. Optical properties of Sm3+ doped lead fluoroborate glasses. J. Phys. Chem. Solids 2000, 61, 1535–1542. [Google Scholar] [CrossRef]
  11. Pan, C.; Kou, K.; Wu, G.; Zhang, Y.; Wang, Y. Fabrication and characterization of AlN/PTFE composites with low dielectric constant and high thermal stability for electronic packaging. J. Mater. Sci. Mater. Electron. 2016, 27, 286–292. [Google Scholar] [CrossRef]
  12. Okasha, A.; Gaafar, M.S.; Marzouk, S.Y. The influence of concentration variationon the spectroscopic behavior of Sm3+ doped zinc–lead-phosphates glasses for orange and reddishorangelight-emitting applications: Experimental and Judd–Ofelt approach. J. Mater. Sci. Mater. Electron. 2023, 34, 354. [Google Scholar] [CrossRef]
  13. Hussain, S.; Amjad, R.J.; Tanveer, M.; Nadeem, M.; Mahmood, H.; Sattar, A.; Iqbal, A.; Hussain, I.; Amjad, Z.; Hussain, S.Z.; et al. Optical Investigation of Sm3+ Doped in Phosphate Glass. Glass Phys. Chem. 2017, 43, 538–547. [Google Scholar] [CrossRef]
  14. Marzouk, M.A.; Elbatal, H.A. Investigation of photoluminescence and spectroscopic properties of Sm3+-doped heavy phosphate glasses before and after gamma irradiation. Appl. Phys. A 2021, 127, 70. [Google Scholar] [CrossRef]
  15. Alzahrani, J.S.; Alrowaili, Z.A.; Alqahtani, M.S.; Eke, C.; Olarinoye, I.O.; Adam, M.; Al-Buriahi, M.S. Influence of Alkaline Earth Metals on the Optical Properties and Radiation-Shielding Effectiveness of Sm3+-Doped Zinc Borophosphate Glasses. J. Electron. Mater. 2023, 52, 7794–7806. [Google Scholar] [CrossRef]
  16. Okasha, A.; Abdelghany, A.M.; Mohamed, S.K.; Marzouk, S.Y.; El-Batal, H.A.; Gaafar, M.S. Gamma ray interactions with samarium doped strontium phosphate glasses. J. Mater. Sci. Mater. Electron. 2018, 29, 20907–20913. [Google Scholar] [CrossRef]
  17. Vahedi, S.; Okada, G.; Koughia, C.; Sammynaiken, R.; Edgar, A.; Kasap, S. ESR study of samarium doped fluorophosphate glasses for high-dose, high-resolution dosimetry. Opt. Mater. Express 2014, 4, 1244–1256. [Google Scholar] [CrossRef]
  18. Hamdy, Y.M.; Bahammam, S.; Abd El All, S.; Ezz-Eldin, F.M. Spectroscopic properties and luminescence behavior of γ irradiated Sm3+ doped oxy-fluoride phosphate glasses. Results Phys. 2017, 7, 1223–1229. [Google Scholar] [CrossRef]
  19. Ahmed, R.M.; Taha, T.A.; Ezz-Eldin, F.M. Investigation of Sm2O3 effect on opto-electrical parameters and dielectric properties of some fluorophosphate glasses. J. Mater. Sci. Mater. Electron. 2021, 32, 28919–28934. [Google Scholar] [CrossRef]
  20. Mrabet, H.; Khattech, I.; Bouzidi, S.; Kechiche, L.; Jbeli, A.; Al Harbi, N.; Bouzidi, C.; Muñoz, F.; Balda, R. Influence of barium substitution on the physical, thermal, optical and luminescence properties of Sm3+ doped metaphosphate glasses for reddish orange light applications. RSC Adv. 2024, 14, 2070–2079. [Google Scholar] [CrossRef]
  21. Chicilo, F.; Okada, G.; Belev, G.; Chapman, D.; Edgar, A.; Curry, R.J.; Kasap, S. Instrumentation for high-dose, high resolution dosimetry for microbeam radiation therapy using samarium-doped fluoroaluminate and fluorophosphate glass plates. Meas. Sci. Technol. 2019, 65, 075010. [Google Scholar] [CrossRef]
  22. Brancaleon, L.; Moseley, H. Laser and non-laser light sources for photodynamic therapy. Lasers Med. Sci. 2002, 17, 173–186. [Google Scholar] [CrossRef] [PubMed]
  23. Brown, S.B. Photodynamic therapy: Two photons are better than one. Nat. Photonics 2008, 2, 394–395. [Google Scholar] [CrossRef]
  24. Kim, M.M.; Darafsheh, A. Light Sources and Dosimetry Techniques for Photodynamic Therapy. Photochem. Photobiol. 2020, 96, 280–294. [Google Scholar] [CrossRef] [PubMed]
  25. Shafirstein, G.; Bellnier, D.; Oakley, E.; Hamilton, S.; Potasek, M.; Beeson, K.; Parilov, E. Interstitial photodynamic therapy—A focused review. Cancers 2017, 9, 12. [Google Scholar] [CrossRef]
  26. Burtan, B.; Reben, M.; Cisowski, J.; Wasylak, J.; Nosidlak, N.; Jaglarz, J.; Jarząbek, B. Influence of Rare Earth Ions on the Optical Properties of Tellurite Glass. Acta Phys. Polon. A 2011, 120, 579–581. [Google Scholar] [CrossRef]
  27. Burtan, B.; Mazurak, Z.; Cisowski, J.; Czaja, M.; Lisiecki, R.; Ryba-Rymanowski, W.; Reben, M.; Wasylak, J. Optical properties of Nd3+ and Er3+ ions in TeO2–WO3–PbO–La2O3 glasses. Opt. Mater. 2012, 34, 2050–2054. [Google Scholar] [CrossRef]
  28. Burtan, B.; Cisowski, J.; Mazurak, Z.; Jarząbek, B.; Czaja, M.; Lisiecki, R.; Ryba Romanowski, W.; Reben, M.; Grelowska, I. Concentration-dependent spectroscopic properties of Pr3+ ions in TeO2-WO3-PbO-La2O3 glass. J. Non-Cryst. Solids 2014, 400, 21–26. [Google Scholar] [CrossRef]
  29. Burtan-Gwizdala, B.; Reben, M.; Cisowski, J.; Lisiecki, R.; Ryba-Romanowski, W.; Jarzabek, B.; Mazurak, Z.; Nosidlak, N.; Grelowska, I. The influence of Pr3+ content on luminescence and optical behavior of TeO2-WO3-PbO Lu2O3 glass. Opt. Mater. 2015, 47, 231–236. [Google Scholar] [CrossRef]
  30. Mito, T.; Fujino, S.; Takebe, H.; Moringa, K.; Todorki, S.; Sakaguchi, S. Refractive index and material dispersions of multi-component oxide glasses. J. Non-Cryst. Solids 1997, 210, 155–162. [Google Scholar] [CrossRef]
  31. Herzinger, C.M.; Johs, B.; McGahan, W.A.; Woollam, J.A.; Paulson, W. Ellipsometric determination of optical constants for silicon and thermally grown silicon dioxide via a multi-sample, multi-wavelength, multi-angle investigation. J. Appl. Phys. 1998, 83, 3323–3336. [Google Scholar] [CrossRef]
  32. Hruby, A. Evaluation of glass-forming tendency by means of DTA. Czechoslov. J. Phys. 1972, 22, 1187–1193. [Google Scholar] [CrossRef]
  33. Lu, Z.P.; Liu, C.T. A new glass-forming ability criterion for bulk metallic glasses. Acta Mater. 2002, 50, 3501–3512. [Google Scholar] [CrossRef]
  34. Jiusti, J.; Cassar, D.R.; Zanotto, E.D. Which glass stability parameters can assess the glass-forming ability of oxide systems. Int. J. Appl. Glass Sci. 2020, 11, 612–621. [Google Scholar] [CrossRef]
  35. Mawlud, S.Q.; Ameen, M.M.; Md Sahar, R.; Ahmed, K.F. Influence of Sm2O3 Ion Concentration on Structural and Thermal Modification of TeO2-Na2O Glasses. J. Appl. Mech. Eng. 2016, 5, 1000222. [Google Scholar] [CrossRef]
  36. Manjeet; Kumar, A.; Anu; Lohan, R.; Deopa, N.; Kumar, A.; Chahal, R.P.; Dahiya, S.; Punia, R.; Rao, A.S. Impact of Sm3+ ions on structural, thermal, optical and photoluminescence properties of ZnO–Na2O–PbO B2O3 glasses for optoelectronice device applications. Opt. Mater. 2023, 139, 113778. [Google Scholar] [CrossRef]
  37. Elkhoshkhany, N.; Marzouk, S.Y.; Khattab, M.A.; Dessouki, S.A. Influence of Sm2O3 addition on Judd-Ofelt parameters, thermal and optical properties of the TeO2-Li2I ZnO Nb2O5 glass system. Mater. Charact. 2018, 144, 274–286. [Google Scholar] [CrossRef]
  38. Mohan Babu, A.; Jaamalaiah, B.C.; Sasikala, T.; Saleem, S.A.; Rama Moorthy, L. Absorption and emission spectral studies of Sm3+-doped lead tungstate tellurite glasses. J. Alloys Comp. 2011, 509, 4743–4747. [Google Scholar] [CrossRef]
  39. Judd, B.R. Optical absorption intensities of rare-earth ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  40. Ofelt, G.S. Intensities of crystal spectra of rare-earth ions. J. Chem. Phys. 1962, 37, 511–520. [Google Scholar] [CrossRef]
  41. Klimesz, B.; Lisiecki, R.; Ryba-Romanowski, W. Sm3+-doped oxyfluorotellurite glasses spectroscopic, luminescence and temperature sensor properties. J. Alloys Compd. 2019, 788, 658–665. [Google Scholar] [CrossRef]
  42. Aboudeif, Y.M.; Moteb Alqahtani, M.; Emara, A.M.; Reben, M.; Yousef, E.-S. Luminescence of Phosphate Glasses: P2O5 -ZnO-BaF2-K2TeO3-Al2O3-Nb2O5 doped with Sm3+ ions for display and laser materials. J. Electron. Mater. 2020, 49, 4144–4153. [Google Scholar] [CrossRef]
  43. Sharma, Y.K.; Surana, S.S.L.; Dubedi, R.P.; Joshi, V. Spectroscopic and radiative properties of Sm3+ doped zinc fluoride borophosphate glasses. Mater. Sci. Eng. B 2005, 119, 131–135. [Google Scholar] [CrossRef]
  44. Shoaib, M.; Rooh, G.; Rajaramakrishna, R.; Chanthima, N.; Kim, H.J.; Tuscharoen, S.; Kaewkhao, J. Physical and luminescence properties of samarium doped oxide and oxyfluoride phosphate glasses. Mater. Chem. Phys. 2019, 229, 514–522. [Google Scholar] [CrossRef]
  45. Swetha, B.N.; Keshavamurthy, K.; Pramod, A.G.; Devarajulu, G.; Roopa, K.P.; Rajeshree Patwari, D.; Kebaili, I.; Ahmed, S.B.; Sayyed, M.I.; Khan, S.; et al. Improved photoluminescence and spectroscopic features of Sm3+-doped alkali borate glasses by embedding silver nanoparticles. J. Non-Cryst. Solids 2022, 579, 121371. [Google Scholar] [CrossRef]
  46. Yousef, E.S.; Elokr, M.M.; Aboudeif, Y.M. Raman spectroscopy and raman gain coefficient of telluroniobium-zinc-lead oxyglasses doped with rare earth. Chalcogenide Lett. 2015, 12, 597. [Google Scholar]
  47. Jlassi, I.; Mnasri, S.; Elhouichet, H. Concentration dependent spectroscopic behavior of Sm3+-doped sodium fluoro-phosphates glasses for orange and reddish-orange light emitting applications. J. Lumin. 2018, 199, 516–527. [Google Scholar] [CrossRef]
  48. Timoshenko, A.D.; Matvienko, O.O.; Doroshenko, A.G.; Parkhomenko, S.V.; Vorona, I.O.; Kryzhanovska, O.S.; Safronova, N.A.; Vovk, O.O.; Tolmachev, V.; Baumer, V.N.; et al. Highly-doped YAG:Sm3+ transparent ceramics: Effect of Sm3+ ions concentration. Ceram. Int. 2023, 49, 7524–7533. [Google Scholar] [CrossRef]
  49. Abdullahi, T.I.; Hashim, S.; Ghoshal, S.K.; Sayyed, M.I. Tailored spectroscopic characteristics of a new type of CuO nanoparticles inserted borate glass system: Samarium concentration tuning effect. Helyon 2023, 9, 20262. [Google Scholar] [CrossRef]
  50. Roopa; Eraiah, B. Impact of samarium ion concentration on the physical, structural and optical properties of multi-component borate glasses. J. Non-Cryst. Solids 2022, 596, 121866. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of the reference glass SM.
Figure 1. XRD pattern of the reference glass SM.
Fibers 12 00100 g001
Figure 2. DSC curves of Sm3+-doped fluorophosphate glasses.
Figure 2. DSC curves of Sm3+-doped fluorophosphate glasses.
Fibers 12 00100 g002
Figure 3. Absorption spectra of Sm3+-doped fluorophosphate glass samples in the range of 800–1600 nm (NIR).
Figure 3. Absorption spectra of Sm3+-doped fluorophosphate glass samples in the range of 800–1600 nm (NIR).
Fibers 12 00100 g003
Figure 4. Absorption spectra of Sm3+-doped fluorophosphate glass samples in the range of 300–500 nm (UV-VIS).
Figure 4. Absorption spectra of Sm3+-doped fluorophosphate glass samples in the range of 300–500 nm (UV-VIS).
Fibers 12 00100 g004
Figure 5. The absorption spectra of phosphate glasses doped with Sm3+ were observed at 644 nm in the photoluminescence excitation (PLE) mode.
Figure 5. The absorption spectra of phosphate glasses doped with Sm3+ were observed at 644 nm in the photoluminescence excitation (PLE) mode.
Fibers 12 00100 g005
Figure 6. Absorption cross-section σabs (λ) of investigated glasses for 6H5/26F7/2 transition.
Figure 6. Absorption cross-section σabs (λ) of investigated glasses for 6H5/26F7/2 transition.
Fibers 12 00100 g006
Figure 7. The calculated absorption cross-section (σabs) for the investigated glasses for the 6H5/26P3/2 transition of Sm3+ ions.
Figure 7. The calculated absorption cross-section (σabs) for the investigated glasses for the 6H5/26P3/2 transition of Sm3+ ions.
Fibers 12 00100 g007
Figure 8. (a) Luminescence lifetimes of the 4G5/2 state of Sm3+ ions for sample SM1. (b) Luminescence lifetimes of the 4G5/2 state of Sm3+ ions for sample SM2.
Figure 8. (a) Luminescence lifetimes of the 4G5/2 state of Sm3+ ions for sample SM1. (b) Luminescence lifetimes of the 4G5/2 state of Sm3+ ions for sample SM2.
Fibers 12 00100 g008
Figure 9. (a,b) Photos of glass emissions; (c) emission spectra of the glasses doped with Sm3+ recorded at room temperature upon 402 nm excitation.
Figure 9. (a,b) Photos of glass emissions; (c) emission spectra of the glasses doped with Sm3+ recorded at room temperature upon 402 nm excitation.
Fibers 12 00100 g009
Figure 10. Schematic energy level diagram of Sm3+ ion shows ground state absorption and excited state absorption mechanisms.
Figure 10. Schematic energy level diagram of Sm3+ ion shows ground state absorption and excited state absorption mechanisms.
Fibers 12 00100 g010
Figure 11. The CIE 1931 chromaticity diagram for the SM1 and SM2 glasses under an excitation wavelength of 402 nm.
Figure 11. The CIE 1931 chromaticity diagram for the SM1 and SM2 glasses under an excitation wavelength of 402 nm.
Fibers 12 00100 g011
Table 1. Volume concentration (N), glass matrix composition (mol%), density, and refractive index at 633 nm.
Table 1. Volume concentration (N), glass matrix composition (mol%), density, and refractive index at 633 nm.
SampleComposition (mol%)N (1019 cm−3)Density (g/cm3)n
SM40P2O5–30ZnO–20LiCl–10BaF2 ----3.8721.588
SM140P2O5–30ZnO–20LiCl–10BaF2–XSm2O34.053.7921.585
SM240P2O5–30ZnO–20LiCl–10BaF2–YSm2O311.003.8211.579
Table 2. Thermal characteristics of Sm3+-doped fluorophosphate glasses.
Table 2. Thermal characteristics of Sm3+-doped fluorophosphate glasses.
Glass IDTgTg rangeTxTcTm
SM35723426523720
SM136914440526725
SM23725445540733
Abbreviations: Tg—glass transition temperature, Tx—endothermic onset of the maximum crystallization peak in glass; Tc—maximum crystallization peak in glass, Tm—endo thermic peak (glass melting temperature).
Table 3. Stability parameters of the Sm3+-doped fluorophosphate glasses.
Table 3. Stability parameters of the Sm3+-doped fluorophosphate glasses.
Glass IDHruby/KHWeinberg/KWLu and Liu/KLLKA
SM0.2340.0950.39569
SM10.2490.0970.40271
SM20.2530.0990.40373
Abbreviations: KA—Angell parameter, KH—Hruby parameter, KW—Weinberg parameter, KLL—Lu and Liu parameter.
Table 4. Judd–Ofelt intensity parameters (Ωt) for fluorophosphate glasses doped with Sm3+ ions.
Table 4. Judd–Ofelt intensity parameters (Ωt) for fluorophosphate glasses doped with Sm3+ ions.
Glass IDΩ2 (10−20 cm2)Ω4 (10−20 cm2)Ω6 (10−20 cm2)σrms (10−6)
SM10.43271.12241.00710.354
SM20.98971.97651.45760.327
Table 5. This study compares the experimentally obtained lifetimes of the 4G5/2 emission band in fluorophosphate glasses (τexp) with the lifetimes estimated using the J-O methods (τrad J-O) and quantum efficiencies (η).
Table 5. This study compares the experimentally obtained lifetimes of the 4G5/2 emission band in fluorophosphate glasses (τexp) with the lifetimes estimated using the J-O methods (τrad J-O) and quantum efficiencies (η).
Glass ID4G5/2 Lifetime (ms)
τexpτrad J-Oη (%)
SM13.1343.85696
SM23.7584.86597
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Burtan-Gwizdala, B.; Cisowski, J.; Lisiecki, R.; Kowalska, K.J.; Jarzabek, B.; Nosidlak, N.; Reben, M.; Alshehri, A.M.; Hussein, K.I.; Yousef, E.S. Thermal, Optical, and Emission Traits of SM3+-Ion-Doped Fluoride/Chloride/Oxide Glass for Red/Orange Laser Fiber Applications. Fibers 2024, 12, 100. https://doi.org/10.3390/fib12110100

AMA Style

Burtan-Gwizdala B, Cisowski J, Lisiecki R, Kowalska KJ, Jarzabek B, Nosidlak N, Reben M, Alshehri AM, Hussein KI, Yousef ES. Thermal, Optical, and Emission Traits of SM3+-Ion-Doped Fluoride/Chloride/Oxide Glass for Red/Orange Laser Fiber Applications. Fibers. 2024; 12(11):100. https://doi.org/10.3390/fib12110100

Chicago/Turabian Style

Burtan-Gwizdala, Bozena, Jan Cisowski, Radoslaw Lisiecki, Kinga J. Kowalska, Bozena Jarzabek, Natalia Nosidlak, Manuela Reben, Ali M. Alshehri, Khalid I. Hussein, and El Sayed Yousef. 2024. "Thermal, Optical, and Emission Traits of SM3+-Ion-Doped Fluoride/Chloride/Oxide Glass for Red/Orange Laser Fiber Applications" Fibers 12, no. 11: 100. https://doi.org/10.3390/fib12110100

APA Style

Burtan-Gwizdala, B., Cisowski, J., Lisiecki, R., Kowalska, K. J., Jarzabek, B., Nosidlak, N., Reben, M., Alshehri, A. M., Hussein, K. I., & Yousef, E. S. (2024). Thermal, Optical, and Emission Traits of SM3+-Ion-Doped Fluoride/Chloride/Oxide Glass for Red/Orange Laser Fiber Applications. Fibers, 12(11), 100. https://doi.org/10.3390/fib12110100

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