AnaKely2023 Te Yb

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Ceramics International 49 (2023) 19470–19480

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

A systematic interpretation of the quantum cutting effect by a cooperative


energy transfer mechanism in Te4+/Yb3+ co-doped tellurite glasses
Ana Kely Rufino Souza a, Junior Reis Silva a, Francine Bettio Costa b, Joao Carlos Silos Moraes b,
Luiz Antonio de Oliveira Nunes c, Luis Humberto da Cunha Andrade a, Sandro Marcio Lima a, *
a
Programa de Pós-Graduação em Recursos Naturais, Universidade Estadual de Mato Grosso do Sul, 79804-970, Dourados, MS, Brazil
b
Universidade Estadual Paulista (UNESP), Faculdade de Engenharia, 15385-000, Ilha Solteira, SP, Brazil
c
Instituto de Física de São Carlos, Universidade de São Paulo, 13560-970, São Carlos, SP, Brazil

A R T I C L E I N F O A B S T R A C T

Handling Editor: Dr P. Vincenzini The near-infrared downconversion (DC) mechanism in Te4+/Yb3+ co-doped 75TeO2–25Li2O tellurite glasses
(amounts in mol%) was closely followed using optical and thermal spectroscopy techniques. The glasses were
Keywords: prepared by the conventional melt-quenching method, with a melting temperature of 800 ◦ C, in an ambient
Tellurite glass atmosphere, which were the best synthesis conditions for observing Te4+ in the glass. The results indicated that
Yb3+ ion
excitation in the ultraviolet region led to an intense emission of two NIR photons (at around 978 nm) in the co-
Te4+ ion
doped tellurite glasses. This effect revealed the occurrence of a cooperative energy transfer (CET) mechanism in
Cooperative energy transfer
Thermal lens spectroscopy the system, where a Te4+ ion was responsible for the excitation of two Yb3+ ions. The CET efficiency (ηCET) was
Downconversion calculated from the Te4+ lifetime, obtaining a maximum of 74% for the tellurite glass prepared with the highest
Quantum cutting Yb3+ concentration. The use of thermal lens spectroscopy confirmed the quantum cutting effect, by observing the
Fluorescence quantum efficiency dependence of the thermal properties of the glass on the Yb3+ concentration. A maximum DC efficiency of 137%
was measured for the sample with 4 mol% of Yb3+.

1. Introduction Several materials, such as glasses and crystals, have been investi­
gated as hosts for luminescent ions, aiming at the DC effect, while rare
After Dexter’s demonstration in 1957 [1] of the downconversion earths (REs) have been proposed as emitting ions. In recent years, re­
(DC) effect involving a cooperative energy transfer (CET) process, many searchers worldwide have searched for luminescent materials, beyond
studies have been performed to better understand and quantify this REs, that present efficient CET processes [6,7]. As mentioned above, in
process in luminescent materials, aiming at their application as photonic this mechanism, one ion (activator ion) is used to absorb UV photons,
devices [2]. Among various possible optical applications, the CET with the energy then being transferred in a cooperative way to another
mechanism can make an important contribution to enhancing the pro­ two identical ions (sensitizer ions). The Yb3+ ion has often been used as a
duction of electric current in a solar cell, having shown interesting and sensitizer, mainly due to the strong absorption/emission in the
promising results involving reduction of the spectral mismatch between near-infrared, which overlaps with the silicon-based semiconductor
the emission spectrum of sunlight and the sensitivity of pn junction absorption. The activator/sensitizer pairs that have been studied include
semiconductors used in solar cells [3–5]. This process essentially con­ Tb3+/Yb3+ [8,9], Pr3+/Yb3+ [10–12], Tm3+/Yb3+ [13], Ho3+/Yb3+
sists of the luminescent material absorbing one high-energy ultraviolet [14], and Ce3+/Yb3+ [4,5,15,16].
(UV) photon, which it converts to two half-energy photons in the With the propose to find alternative luminescent activator ions to
near-infrared (NIR). It is expected that part of the NIR emission inside REs for investigating DC effect, more recently, Souza et al. [17] reported
the solar cell can contribute to the photocurrent effect. Consequently, CET from a transition metal (Te4+) to Yb3+ in a tellurite glass, indicating
when CET is observed in a system, the UV energy that was previously that the system could be improved turning it a suitability material for
converted into heat, contributing to a decrease of the photoelectric solar cell applications. In this point is important to mention that CET
process, becomes useful for enhancing the efficiency of the solar cell. mechanism between a transition metal and a RE is not often observed,

* Corresponding author.
E-mail address: smlima@uems.br (S.M. Lima).

https://doi.org/10.1016/j.ceramint.2023.03.079
Received 17 November 2022; Received in revised form 6 March 2023; Accepted 8 March 2023
Available online 11 March 2023
0272-8842/© 2023 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

but yes, between two REs. However, the main DC effect presents in a charge coupled device (CCD) detector, in the spectral range 238–900
system involving two REs is not CET. As example, Florêncio et al. [18] cm− 1, at 3.0 cm− 1 resolution. Mid-infrared absorption spectra were
concluded that DC effect in a Tb3+/Yb3+-codoped tellurite glass does not recorded in the range from 3510 to 990 cm− 1, using a Fourier transform
occur by a CET mechanism, but through a sub-linear process involving a infrared (FTIR) spectrophotometer (Nexus 670, Thermo-Nicolet), with
virtual level. Besides, van Wijngaarden et al. studied the energy transfer spectral resolution of 8 cm− 1 and 32 scans.
process between Pr3+ and Yb3+ couple in LiYF4 and concluded that the UV–Vis–NIR absorption spectra were acquired using a PerkinElmer
cross relaxation (CR) is more efficient than CET, with CR typically 1000 LAMBDA 1050 spectrometer, in the wavelength range from 320 to 1060
times more probable than CET mechanism [11]. However, research nm, with integration time of 0.2 ms. Photoluminescence spectra were
looking for systems with efficient DC effect through CET mechanism are collected after excitation with an argon laser operating at 457.9 nm,
strongly motivates, mainly considering different metallic ions. with transmission through an optical fiber to a portable spectrometer
In the past, tellurite glasses were the object of many studies, mainly (MAYA 2000 Pro, Ocean Optics). The Yb3+ photoluminescence (PL) and
due to their relatively high linear refractive index (1.9–2.3 in the visible PL excitation (PLE) spectra were obtained using a Xe lamp (150 W) as
region), wide optical transparency window (0.3–5 μm), and low phonon the excitation source, coupled to a monochromator (Newport Corner­
energy (~700 cm− 1) [19]. More recently, these glasses have been stone 260™) and a 300-line diffraction grating with blaze wavelength of
investigated as hosts for different luminescent ions, such as the REs, 500 nm. The samples were exited in the spectral range from 300 to 525
aiming at applications in optical devices [18,20]. This glass family has nm, with steps of 5 nm, and the PL was transmitted through an optical
the advantages of low cost and easy preparation, allowing the incor­ fiber to a portable spectrometer (MAYA 2000 Pro, Ocean Optics), which
poration of REs at high doping concentrations. Although the tellurite had been previously calibrated in terms of its response in the spectral
glasses have been studied for a long time, it was only in 2018 that the emission range from 850 to 1100 nm. For observing the Te4+ emission
Te4+ emission in a tellurite glass was first reported [21]. Since then, our region, PL and PLE spectra were obtained using a spectrofluorometer
research group has explored the factors that influence Te4+ formation in (LS-55, PerkinElmer), with excitation in the range from 300 to 500 nm
tellurite glass, aiming to identify the optimum melting temperature, and emission from 500 to 800 nm, with steps of 5 nm. The photo­
atmosphere, and lithium concentration, particularly for the multiplier was operated at 900 V, with 0.03 μs detection delay and 0.10
tellurium-lithium binary system [22,23]. It was found that Te4+ emis­ μs integration time. Photoluminescence decay curves of the 3T1u level of
sion was most pronounced using a TeO2:Li2O proportion of 75:25 (mol Te4+ in the visible region were obtained using excitation at 460 nm,
%), with the glass being prepared at 800 ◦ C in an ambient atmosphere. employing an OPO laser pumped by a 10 ns pulsed Nd:YAG laser (10
This sample composition and synthesis conditions were selected for this Hz), with the signal being detected by an Andor iCCD.
research. Thermal lens spectroscopy (TLS) in the mode-mismatched dual-beam
In the present study, the DC effect resulting from the CET mechanism configuration was used to measure the fraction of absorbed energy
was systematically investigated in tellurite glasses, with the transition converted into heat by the sample. In this setup, a tunable Ti:sapphire
metal Te4+ as the activator and the rare earth Yb3+ as the sensitizer. laser (700–1040 nm) or a diode laser (375 nm) was used as the excita­
Multiwavelength thermal lens spectroscopy (TLS) was used to determine tion source, while a HeNe laser (632.8 nm) was used as the probe beam.
the absolute value of the fluorescence quantum efficiency (η) for both Further details of the TLS experimental setup have been described pre­
Te4+ and Yb3+ ions, so that CET efficiency from one Te4+ to two Yb3+ viously [25,26]. All the measurements were performed at room
ions could be measured by considering radiative and non-radiative temperature.
transitions present in the system.
3. Results and discussion
2. Materials and methods
3.1. Glass structural characterization
Tellurite glasses with nominal composition (mol%) of (100 − x)
(75TeO2 + 25Li2O) + xYb2O3, where x = 0, 0.5, 1.0, 2.0, 3.0, and 4.0 Fig. 1 shows the differential scanning calorimetry (DSC) curves ob­
mol%, were prepared by the conventional melt-quenching method, tained in a N2 atmosphere for the studied set of samples. The Tg, Toc, and
using a melting temperature of 800 ◦ C, in an ambient atmosphere. The
samples were labeled as 75TeLi: xYb. The reagents Li2CO3 (VETEC),
TeO2 (99% purity, Sigma-Aldrich), and Yb2O3 (99.99% purity, Sigma-
Aldrich) were weighed out, mixed, and melted in a Pt-5% Au crucible
for 30 min. The melt was poured into a stainless-steel mold preheated
close to the glass transition temperature (Tg), in air, and was subse­
quently annealed for 5 h to relieve the mechanical stresses. The samples
were then cut and polished, prior to the structural and spectroscopic
characterizations.
The volumetric density (ρ) was determined by the Archimedes
method, using distilled water as the immersion liquid. Differential
scanning calorimetry (DSC) measurements were carried out with the
samples in a nitrogen atmosphere and heating at a rate of 10 ◦ C/min.
The glass-forming tendency coefficient (Hr) was calculated using the
relation of Hrubý [24]:
Toc − Tg
Hr = (1)
Tom − Toc

where, Tg is the glass transition temperature, Toc is the temperature for


onset of crystallization, and Tom is the temperature for onset of melting.
Raman spectra were obtained with excitation at 785 nm, employing
a 150 mW diode laser. The signal was detected using a confocal mi­
croscope (BX51-Voyage) coupled to an Andor monochromator and a
Fig. 1. DSC of the 75TeLi: 0.5; 1.0; 2.0; 3.0; and 4.0 mol% Yb2O3 glasses.

19471
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

Tom values were determined, as indicated in the figure. The crystalliza­


tion and liquid phase were defined as the beginning of the first
exothermic, Toc, and the first endothermic, Tom, peaks, respectively. The
values obtained for Tg, Toc, Tc, the thermal stability range ΔT = (Toc - Tg),
Tom, Tm, the glass-forming tendency coefficient, Hr, and the volumetric
density, ρ, for the 75TeLi: xYb glasses are shown in Table 1. It should be
noted that the Tc values correspond to the first more pronounced crys­
tallization peak observed in the DSC curves. A systematic study would
need to be performed to obtain a complete identification of the crystal
phases. It can be seen that the ρ, Toc, Tom, ΔT, and Tg values increased
with the amount of Yb2O3, as observed in other recent work concerning
RE-doped tellurite glasses [27]. The observed increase of density with
addition of Yb2O3 could be explained by its high molecular weight. The
75TeLi glass allows greater incorporation of Yb2O3 into the vitreous
network, with the ability to form more than one electric dipolar envi­
ronment for Yb3+ ions, consequently increasing the order of solubility
[27]. This is responsible for the increases in Tg, density, and molar
volume, mainly because the formation of bridged oxygen sites results in
better packing of the 3D structure in the glass. In addition, among the
75TeLi glasses studied, the sample with 3.0 mol% of Yb2O3 presented a
greater thermal stability range (ΔT = 94 ◦ C) and higher glass-forming
tendency coefficient (Hr = 1.03). These features indicated that the
glass was the most resistant to temperature variations and was easier to
form [24].
Fig. 2 shows the Raman spectra for the different samples. It can be
seen that Yb2O3 addition led to a reduction of the maximum intensity of
the scattering bands, reflecting a change in the Te atom coordination
state from TeO4 trigonal bi-pyramidal (tbp) to TeO3 trigonal pyramidal
(tp), by means of the TeO3+1 polyhedron, with increase of the non-
bridging oxygens (NBOs) [28,29]. The detailed plots in Fig. 2(a) and
(b) show overlap of bands between 550 and 850 cm− 1, with maximum
scattering at around 661 cm− 1 attributed to the continuous network
structure of the TeO4 and TeO3+1 units. The addition of Yb3+ led to an
apparent small shift to higher energy of the band centered at around 745
cm− 1. This was related to the Te–O- stretching vibrations of TeO3+1 or
TeO3 units, where the NBO atoms interacted with adjacent Te atoms.
The shift was due to a third band centered at approximately 800 cm− 1,
which was related to the Te–O- stretching vibrations of TeO3+1 or TeO3
Fig. 2. (a) Raman spectra of the 75TeLi: 0.5; 1.0; 2.0; 3.0; and 4.0 mol% Yb2O3
units where the NBO atoms had little interaction with adjacent Te atoms glasses, with (inset) magnification of the region from 240 to 540 cm− 1. (b)
[30]. The scattering band located from 370 to 540 cm− 1 (Fig. 2(a)) could Selected 550-900 cm− 1 region to deconvolute using symmetric Gaussian func­
be attributed to the symmetric stretching (and bending) vibrations of tions (c) for the 75TeLi: 4 Yb glass.
Te–O–Te linkages, formed by vertex-sharing of TeO4 tbps, TeO3+1
polyhedra, and TeO3 tp. This scattering region is related to the bending four Gaussian functions, as plotted in Fig. 2(c). The center positions of
vibrations of TeO3 tp with two or three NBOs. Comparison of the Raman the Gaussian functions were located at 580, 661 760, and 800 cm− 1. The
spectra with those obtained for the 80TeLi matrix melted at 850 ◦ C [17] values for the areas under the four bands were used to calculate the
showed that the TeO4 proportion was lower in 75TeLi, suggesting that coordination number of Te ions with oxygen (NTe–O), using the Kaur
the glass underwent important structural changes, due to the variations et al. model [31]. The values obtained are plotted in Fig. 3, together with
in the proportions of former and modifier oxides, besides the lower the NBO fraction calculated using the model proposed by Himei et al.
melting temperature used in the present study (800 ◦ C). [32]. This model represents a change from TeO4 tbp to TeO3 tp with
In order to investigate the relations of the TeO3 and TeO4 bands, the NBO atoms, using the molar ratio of [TeO3]/[TeO4] and the atomic ratio
Raman spectrum of the 75TeLi: 4.0 Yb sample was deconvoluted using [O]/[Te]. The observed increase in the [TeO3]/[TeO4] ratio, as a func­
tion of Yb3+ concentration, showed a strong correlation to the increase
Table 1 in Tg (from 258 to 290 ◦ C) [33,34]. Fig. 3 shows an inverse behavior for
Nominal composition, glass transition temperature (Tg), initial crystallization the two parameters, with NTe–O decreasing and the NBO fraction
temperature (Toc), crystallization temperature (Tc), thermal stability range ΔT increasing, as a function of the Yb3+ concentration. The NTe–O change
= (Toc – Tg), initial melting temperature (Tom), melting temperature (Tm), was more marked when 0.5 mol% of Yb2O3 was added in the matrix
coefficient of glass-forming tendency (Hr), and density (ρ) of the 75TeLi: xYb (without ytterbium), with the Te–O–Te bridges being replaced by
glasses. The error in the density values is ±0.01. Yb-O-Te bridges, which have higher bond dissociation energy (Te–O:
Yb2O3 Tg Toc Tc ΔT Tom Tm Hr ρ (g/ 391 kJ mol− 1; Yb–O: 397.7 kJ mol− 1).
(mol%) (◦ C) (◦ C) (◦ C) (◦ C) (◦ C) (◦ C) cm3) The NBOs fraction shown in Fig. 3 presented an increase, indicating
0 258 292 354 34 422 435 0.26 4.81 that Yb2O3 promoted a cleavage of the continuous network, changing
0.5 263 316 387 53 422 438 0.50 4.90 the Te atom coordination state from TeO4 tbp to TeO3 tp by means of the
1.0 266 322 408 56 430 437 0.52 4.95 TeO3+1 polyhedron, which induced isolated structural units in the glass.
2.0 273 335 388 62 455 467 0.52 4.95
The mid-infrared absorption spectra (Fig. 4) showed a well-known
3.0 282 376 396 94 467 470 1.03 5.00
4.0 290 372 391 82 456 466 0.98 5.10
OH− absorption band centered at around 3000 cm− 1, identified as a

19472
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

Fig. 3. Coordination number of Te and oxygen (NTe–O) and the NBO/(BO +


NBO) relation as a function of the Yb2O3 concentration.

Fig. 5. UV–Vis absorption coefficient and NIR spectra regions for the Yb3+-
doped 75TeLi glasses (a), and partial energy degeneration levels diagram for
Yb3+ (b). The inset shows the 4F7/2 → 4F5/2 transition area as a function of Yb3+
concentration.

Stark manifolds of the ion (b). At higher energy of the absorption spectra
(Fig. 5(a)), a blue shift occurred when Yb2O3 was added in the glass
matrix (indicated by a black arrow in the figure), as already observed in
other RE3+-doped tellurite glasses [17]. This feature was related to the
structural rearrangement discussed above. The blue shift, together with
the observed shift to lower energy in the IR spectra (Fig. 4), led to a
broader optical transparency window of the glasses.
Fig. 4. Mid-infrared absorption spectra for the 75TeLi: xYb glasses in the range The Yb3+ absorption band with maximum at around 978 nm corre­
from 990 to 3600 cm− 1.
sponded to the well-known 2F7/2 → 2F5/2 transition of the ion. Although
the spectra were acquired at room temperature, the four Stark manifolds
combination of weakly H-bonded OH− and free OH− . This band of the 2F7/2 state and the three of the 2F5/2 state (as shown in Fig. 5(b))
decreased when Yb2O3 was added in the glassy matrix, because the OH− could also be seen in the spectra. The inset of Fig. 5(a) shows the Yb3+
content was influenced by the free volume of the glass, due to the NBO absorption area plotted as a function of the Yb3+ concentration, where
created by the alkaline insertion [35]. The OH− content was calculated the linear relationship indicates that Yb3+ was effectively incorporated
using the method proposed by Massera et al. [36], with the values ob­ into the analyzed set of glasses. In other words, this indicated that the
tained varying from 6.4 × 1021 ions/cm3 (sample without Yb) to 1.2 × nominal Yb composition was strongly correlated with the incorporated
1021 ions/cm3 (sample with 4.0% of Yb), representing a reduction of Yb concentration.
81%. This substantial variation in OH− concentration could reduce the Fig. 6(a)-(d) show the PL and PLE contour plots for the 75TeLi: xYb
nonradiative transitions in the doped ion, as reported for other systems glasses, with the excitation wavelength on the x-axis, in the range from
[37,38]. 300 to 500 nm, and the emission wavelength on the y-axis, from 525 to
The band at around 2283 cm− 1 was due to strong H-bonding in the 800 nm. The intensity is plotted using a logarithm color scale, which is
OH− unit [36], while the band at around 1490 cm− 1 was identified as the same for all the maps. The intense emission evident in (a) could be
being due to the combination of two fundamental bands (TeO4 and attributed to the Te4+ emission [21]. The contours show that the Te4+
TeO3, 661 + 745 = 1406 cm− 1) observed in the Raman results. This emission intensity decreased when Yb2O3 was added, disappearing
band was evidenced with increase of the dopant concentration and was completely for the 2 mol% Yb3+ co-doped tellurite glass. This suggested
due to a shift of the glass valence band to lower energy. that the absorbed energy was transferred from Te4+ to Yb3+.
The relation between the Yb3+ NIR emission and the UV–Vis exci­
3.2. Cooperative energy transfer efficiency (ηCET) tation was investigated using PL and PLE measurements of the 75TeLi:
xYb co-doped glasses. Fig. 7(a)-(e) show the PL and PLE contour plots for
Fig. 5 shows the UV–Vis–NIR absorption coefficient spectra of the the studied samples, with excitation and emission wavelengths in the
75TeLi: xYb glasses (a) and the Yb3+ partial energy levels diagram with ranges from 300 to 525 nm and from 900 to 1100 nm, respectively. The

19473
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

Fig. 6. Excitation-emission contour plots for the 75TeLi: xYb glasses in the Te4+ emission region: (a) x = 0, (b) x = 0.5, (c) x = 1.0, and (d) x = 2.0.

intensity is shown using a logarithmic color scale, which is the same for energy migration between them, consequently increasing the probabil­
each map. The contours evidenced that the Yb3+ emission intensity ity of nonradiative relaxation, including energy trapping by defects such
increased when the Yb3+ concentration changed from 0.5 up to 2 mol% as OH− , impurities, and ion-host interactions [26].
(Fig. 7(a)–(c)), with a slight decrease for higher Yb3+ content (Fig. 7(d)– Based on the absorption and emission results described above, a
(e)). The PL-PLE maps indicated that the broad Yb3+ excitation region schematic energy level diagram was prepared to indicate all the tran­
from 350 to 500 nm was characteristic of the Te4+ ion. This supported sitions involving Te4+ and Yb3+ ions, as shown in Fig. 10. When a UV
the hypothesis that Te4+ transferred its energy to Yb3+, which is an (360 nm) or blue (460 nm) photon is absorbed by Te4+, one electron is
important optical feature for the use of tellurite material to improve the promoted from the 1A1g state to the 3A1u or 3T1u state, respectively. After
efficiency of c-Si solar cells. that, the electron decays very rapidly and nonradiatively to the bottom
The white dashed lines in Figs. 6(d) and 7(c) indicate the PL and PLE of the 3T1u state, from where it can return to the 1A1g ground state,
spectra for the 75TeLi: 2 Yb glass that are plotted in Fig. 8. All the spectra nonradiatively or radiatively. In the presence of Yb3+, there is the pos­
were normalized to make the band shapes clear. The PLE data suggested sibility of the CET process occurring from one Te4+ ion to two neigh­
that the absorption region responsible for the Te4+ and Yb3+ emissions boring Yb3+ ions, promoting them to their excited 2F5/2 state. This
was the same. In other words, this reinforced that the glassy system process results in an emission of two photons in the NIR at around 978
studied here exhibited cooperative energy transfer (CET) from Te4+ to nm, which is required for application in Si solar cells. Considering the
Yb3+, as reported previously by Souza et al. [17] for 80TeLi:Yb2O3 heat created by nonradiative transitions, in the presence of the CET
co-doped glasses prepared at a melting temperature of 850 ◦ C. It is well mechanism in the system, a reduction in the heat would be expected if a
known that Te4+ in tellurite glass can exhibit two excitation bands, higher Yb3+ concentration was added in the sample.
centered at around 360 nm (1A1g → 3A1u) and 460 nm (1A1g → 3T1u), It is important to note that the electron in the 3T1u state could also be
with absorption intensities strongly affected by the glass network [15], transferred to the 2F5/2 metastable level of Yb3+ by the phonon-assisted
mainly due to the overlap between the glass charge transfer band and energy transfer mechanism [39]. In this case, an increase in the non­
the Te4+ absorption bands. radiative rate would be expected, with a consequent decrease in the
The photoluminescence (PL) spectra of the 75TeLi: xYb glasses under radiative transition from the 3T1u state of Te4+. However, as will be
excitation at 457.9 nm were acquired in the range 450–1100 nm (Fig. 9 discussed later, the fraction of heat created in the sample decreased with
(a)). The broad and intense emission band assigned to Te4+ (3T1u → 1A1g, the addition of Yb3+, indicating that CET was more probable than the
maximum at 675 nm) and the Yb3+ (2F5/2 → 2F7/2 transition) charac­ phonon-assisted mechanism in the glasses studied.
teristic emission band were observed for all the 75TeLi: xYb glasses In order to determine the CET efficiency (ηCET), luminescence decay
studied. A continuous decrease in the Te4+ emission was observed, while curves of Te4+ were measured at 650 nm, under excitation at 460 nm
the Yb3+ emission increased up to sample 75TeLi: 2 Yb (see Fig. 9(b)), (Fig. 11). The CET process occurring between Te4+ and Yb3+ could be
suggesting that there was CET from Te4+ to Yb3+. The observed reduc­ confirmed by the non-exponential behavior of the decay curves,
tion in the Yb3+ emission for the 75TeLi: 3 Yb and 75TeLi: 4 Yb samples becoming faster as Yb3+ was incorporated into the glass. As usually
could be explained by the Yb3+ concentration quenching effect, due to performed for non-exponential curves, the lifetime was interpreted by
the proximity between Yb3+ ions in the glass, causing a successive its average value (<τ>), and the choice of proper formula was based in

19474
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

Fig. 7. Excitation-emission contour plots for the 75TeLi: xYb glasses in the Yb3+ emission region: (a) x = 0.5, (b) x = 1.0, (c) x = 2.0, (d) x = 3.0, and (e) x = 4.0.

the distribution of radiative rates in the glass system, as stated by Zatryb effect, such as in solar cells.
and Klak [40]. So, the average lifetime for the studied glasses were
∫∞ ∫∞
determined using the expression < τ > = 0 t × I(t)dt/ 0 I(t)dt, where I
(t) is the photoluminescence intensity as a function of time, t [41]. The 3.3. Te4+ and Yb3+ fluorescence quantum efficiencies
<τ> values obtained for the Te4+ lifetime are listed in the inset of
Fig. 11. A decrease of <τ> was observed when Yb3+ was added in the Thermal lens spectroscopy (TLS) was used to indirectly determine
glassy matrix (from 6.5 μs for the undoped glass to 1.7 μs for the 75TeLi: the Te4+ and Yb3+ fluorescence quantum efficiencies (ηTe and ηYb,
4 Yb glass). The ηCET values were determined using the average lifetime respectively), measuring the fraction of the absorbed energy that was
values for the samples, according to the following adapted equation converted into heat by the system. These parameters were fundamental
[42]: for calculation of the downconversion efficiency (described below).
Firstly, all the Yb3+-doped systems were excited at different wave­
< τ >x%Yb
ηCET = 1 − (2) lengths in the NIR absorption region (915–1005 nm range), using the
< τ >Te
multi-wavelength thermal lens method [26,44]. In this experimental
The calculated ηCET values are plotted in Fig. 12, together with the strategy, the energy used to excite the glasses is absorbed by Yb3+ (from
2
average lifetime, as a function of the Yb3+ concentration. An increase to F7/2 to 2F5/2), after which radiative and nonradiative transitions can
74% was observed for the sample with the highest Yb3+ concentration. occur from the 2F5/2 metastable level. The radiative transition was dis­
This value was equal to or higher than those reported in the literature for cussed in the last section, while here the complementary nonradiative
Ce3+/Yb3+ co-doped YAG crystal and borate glass [43], demonstrating part is investigated by TLS. Considering that the energy due to the
the relevance of tellurite glasses for applications involving the CET nonradiative rate is converted into heat, and that the radiative transition
always starts from the bottom of the 2F5/2 state, it is then evident that

19475
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

Fig. 8. PL (open circles) and PLE (filled circles) spectra for the 75TeLi: 2 Yb
sample. The Yb3+ and Te4+ ions were excited at 468 nm for the PL spectrum.
The PLE was carried out observing the emissions at 978 and 600 nm for the
Yb3+ and Te4+ ions, respectively.

decreasing the excitation wavelength from 1005 to 915 nm will lead to


an increase in the fraction of the absorbed energy converted into heat, φ,
as reported by Lima et al. [44]. Hence, by measuring φ as a function of
the excitation wavelength, it is possible to understand the radiative and
nonradiative phenomena simultaneously in the Yb3+-doped system.
A detailed description can enable a better understanding of the
procedure used. Fig. 13(a) shows a typical thermal lens (TL) curve, I(t)/
Io, obtained for the 75TeLi: 2.0 Yb glass with excitation at 915 nm and
Pin = 129 mW. By fitting the transient with the theoretical model pro­
posed by Shen et al. [45], the parameters θ = -(0.0669 ± 0.0005) rad
and tc = (6.0 ± 0.2) ms were determined. Since the thermal diffusivity D
= w2oe/4tc, and woe = (7.5 ± 0.1) μm for the excitation used, a value of
(2.34 ± 0.02) x 10− 3 cm2/s was calculated for the 75TeLi: 2.0 Yb glass,
Fig. 9. (a) PL in the visible and infrared region for the 75TeLi: xYb glasses with
in good agreement with the literature [46]. The parameter θ is propor­
excitation at 457 nm, highlighting (inset) the Te4+ emission. (b) Integrated
tional to the amplitude of the TL transient signals and is related to the areas of the Te4+ emission (centered at 680 nm) and the Yb3+ emission
thermo-optical parameters of the sample [44]. Values of θ were obtained (centered at 978 nm), as a function of the Yb2O3 concentration.
using different excitation powers, in order to verify its linear depen­
dence on the absorbed power. Fig. 13(b) shows the experimental θ data
plotted as a function of the incident power, together with the linear
curve fit, for the 75TeLi: 2.0 Yb glass. The angular coefficient of the
curve was obtained as θ/Pin = -(0.514 ± 0.001) W− 1. Using the
expression Θ = θ/Pabs, with Pabs = PinαL, where α = (2.7 ± 0.1) cm− 1 is
the absorption coefficient and L = (0.034 ± 0.001) cm is the sample
thickness, Θ = (5.78 ± 0.09) W− 1 was determined for the 75TeLi: 2.0 Yb
glass with excitation at 915 nm. This same procedure was applied using
the other excitation wavelengths, λexc, so that Θ ≡ Θ(λexc ). The described
procedure was also used for each Yb3+-doped sample.
A linear behavior of Θ vs. λexc is expected when the sample exhibits
one way to depopulate the metastable state, with the following theo­
retical expression relating them [44]:
( )
1 ds λexc
Θ(λexc ) = 1− η (3)
Kλp dT < λemi >

where, λp is the laser probe wavelength, K is the thermal conductivity,


ds/dT is the temperature coefficient of the optical path length, η is the
fluorescence quantum efficiency, and <λemi> is the average emission
wavelength. As noted from the equation, the linear coefficient (1/λpK) Fig. 10. Partial energy level diagram of Te4+ and Yb3+ ions in the 75TeLi
ds/dT corresponds to the Θ value for the undoped sample, so Θundoped = glasses. The solid arrows indicate radiative transitions (absorption and emis­
sion) and the solid wave arrows are the nonradiative transitions. The dashed
(1/λpK)ds/dT. Since <λemi> = (978 ± 5) nm is a known experimental
arrows indicate the CET mechanism.
parameter, ηYb could be determined by dividing the angular coefficient

19476
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

Fig. 11. Decay curves for luminescence from the Te4+ ion (3T1u → 1A1g tran­
sition at 650 nm), with excitation at 460 nm.

Fig. 13. (a) Thermal lens transient signal for the 75TeLi glass doped with 2.0
mol% of Yb2O3. Excitation at 915 nm with laser power at 421 mW. (b)
Amplitude of the transient thermal lens signal (θ), as a function of incident
excitation power (Pin).
Fig. 12. Calculated average lifetime <τ> and CET efficiency values, as a
function of Yb2O3 concentration.

by the linear one [26,44]. Fig. 14(a) shows the absorption coefficient
spectrum for the 75TeLi: 2.0 Yb glass (solid brown line), together with
the absorption coefficient values determined by the Beer-Lambert law
(closed balls), at the sample position used for the TLS measurements.
Fig. 14(b) shows the Θ values obtained for different excitation wave­
lengths (open balls) and the linear curve fit (solid red line) from Eq. (3).
The best fit was obtained with the linear coefficient Θundoped = (33 ± 3)
W− 1 and, consequently, η = (0.89 ± 0.05), for the 75TeLi: 2.0 Yb glass.
The same methodology described above was applied for the other
Yb3+ concentrations, obtaining absolute values for ηYb of 95, 95, 89, 91,
and 83% for the 75TeLi: 0.5 Yb, 75TeLi: 1 Yb, 75TeLi: 2 Yb, 75TeLi: 3
Yb, and 75TeLi: 4 Yb samples, respectively (plotted in Fig. 15). The
observed Yb3+ concentration quenching (~10%) could be explained by
the energy transfer between ions (including impurities), ion-host in­
teractions, or energy trapping due to defects, such as OH− [26]. The
observed decrease of ηYb was lower than that measured for the Fig. 14. (a) Absorption coefficient measured by the spectrophotometer and by
measuring the incident and transmitted excitation power (Beer-Lambert law).
Yb3+-doped 80TeLi glass (~25%) [26]. This suggested that there was
(b) Θ parameter as a function of the excitation wavelength.
greater spacing in the glassy network of the 75TeLi system, avoiding
quenching of the luminescence of the ytterbium that could occur due to
close proximity of the ions.

19477
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

Fig. 16. Measured and calculated fraction of the absorbed energy converted
into heat, φ, for the 75TeLi: xYb glasses with excitation at 375 nm, as a function
Fig. 15. Yb3+ fluorescence quantum efficiency as a function of Yb2O3 con­ of the Yb3+ concentration (left y-axis), and the downconversion efficiency
centration. The solid line is a guide for the eyes. values calculated using Eq. (5) (right y-axis).

To determine the Te4+ fluorescence quantum efficiency, ηTe, for the


λex λex
studied set, the 75TeLi: 0 Yb sample was excited at 375 nm, since Te4+ φdoped = 1 − (ηTe + ηCET ) + 4ηCET (1 − ηYb ) . (4)
〈λemi 〉Te 〈λemi 〉Yb
has a strong absorption band. Typical transient curves (I(t)/Io) for
different excitation powers were measured and fitted, from which θ vs. All the parameters on the right side of the equation were determined
Pin was plotted (not shown), giving θ/Pin = -(9.83 ± 0.14) W− 1. The tc before, so they could be used in this theoretical model to determine
value (and consequently D) obtained was in agreement with that re­ φdoped. The calculated φdoped values are also plotted in Fig. 16, showing
ported in the literature for 80TeLi glass [46]. Since α = (7.9 ± 0.3) cm− 1 good agreement with the measured values.
at 375 nm and L = (0.033 ± 0.001) cm for the studied sample, Θ = (23 When CET is the dominant mechanism in a co-doped system, the
± 1) W− 1 was determined for the 75TeLi: 0 Yb sample. Dividing this downconversion efficiency (ηDC) can be rewritten as follows [42]:
value by the Θundoped = (33 ± 3) W− 1 value determined above gave the
ηDC = ηTe (1 − ηCET ) + 2ηYb ηCET . (5)
fraction of the absorbed energy converted into heat, φ = 0.695 ± 0.002.
Considering the Te4+ average emission wavelength <λemi> = 711.5 nm, In the present study, the ηDC values were calculated considering ηTe
λexc = 375 nm, and the relation φ = 1− η(λexc/<λemi>), it was then and ηYb found above by TLS and ηCET obtained by lifetime data. The ηDC
possible to obtain ηTe = (0.58 ± 0.04). This value was of the same order values (Fig. 16) showed an increase with the Yb concentration, reaching
as that determined for the 80TeLi glass [21]. a maximum value of ~137% for the sample with 4 mol% Yb2O3 con­
centration. It should be noted that ηDC greater than unity is expected in
3.4. Downconversion efficiency (ηDC) the process of changing visible photons into NIR radiation [1].

As noted in Fig. 10, the CET mechanism from Te4+ to Yb3+ resulted in 4. Conclusions
a reduction in the population of the 3T1u state. Consequently, there was a
reduction in the nonradiative relaxation from 3T1u to 1A1g in Te4+. This The optical and thermal properties of 75Te–25Li (mol%) tellurite
behavior suggested that the fraction of the absorbed energy converted glass doped with Yb3+ at different concentrations, synthesized at 800 ◦ C
into heat, φdoped, by the Te4+/Yb3+ co-doped tellurite glasses decreased in an ambient atmosphere, were studied to confirm the Te4+ → Yb3+
when the CET mechanism became more pronounced. To check this CET mechanism. Measurement of the radiative transitions, using the
hypothesis, φdoped values were determined with the TLS configured to average lifetime, enabled determination of CET efficiency of 74% for the
excite the glasses in the UV region (375 nm), directly in the Te4+ ion. As sample with the highest Yb3+ concentration. The nonradiative effect in
described above, the TL characteristic curves were determined for each the co-doped samples was monitored using thermal lens spectroscopy.
sample using different excitation powers, with the amplitude of the The experimental results demonstrated that the CET reduced the frac­
recorded TL signal (θ) presenting a linear relation with the incident tion of the absorbed energy in the UV region that was converted into
power, Pin. Since the parameters α and L were known for each sample, heat, in agreement with the theoretical model. A maximum DC effi­
Θdoped could be determined as a function of the Yb3+ concentration. ciency of 137% was determined for the sample co-doped with 4 mol% of
Calculation of Θdoped/Θundoped, where Θundoped is the measured value Yb3+. The systematic interpretation reported here indicated that tel­
determined above, with φundoped = 1, resulted in the measured values for lurite glass is a potential material suitable for coupling in a silicon-based
φdoped, which are plotted in Fig. 16. The observed reduction confirmed, solar cell to increase energy generation.
as expected, that CET was the mechanism responsible for transferring
energy from Te4+ to Yb3+. Declaration of competing interest
Recently, Bento et al. [47] solved a rate equation system for the
population distribution of two ions, where one is considered to be the The authors declare that they have no known competing financial
activator and the other the sensitizer. Appling this model to the studied interests or personal relationships that could have appeared to influence
system, considering Te4+ and Yb3+ to be the activator and sensitizer, the work reported in this paper
respectively, φdoped could be given by Ref. [47]:

19478
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

Acknowledgments of a Te4+center with broad red emission band and high fluorescence quantum
efficiency in TeO2-Li2O glass, J. Lumin. 198 (2018), https://doi.org/10.1016/j.
jlumin.2018.02.002.
Financial support for this work was provided by Coordenação de [22] F.B. Costa, A.K.R. Souza, J.R. Silva, J.C.S. Moraes, L.A. de Oliveira Nunes, L.H. da
Aperfeiçoamento de Pessoal de Nível Superior (CAPES), Conselho Cunha Andrade, R. El-Mallawany, S.M. Lima, Effect of lithium addition on Te4+
Nacional de Desenvolvimento Científico e Tecnológico (CNPq), and emission in TeO2-Li2O glasses, J. Non-Cryst. Solids 524 (2019), https://doi.org/
10.1016/j.jnoncrysol.2019.119609.
Fundação de Apoio ao Desenvolvimento do Ensino, Ciência e Tecnologia [23] A. do C. Capiotto, A.K. Rufino Souza, F.B. Costa, J.C. Silos Moraes, L.A. de Oliveira
do Estado de Mato Grosso do Sul (FUNDECT). Nunes, J.R. Silva, L.H. da Cunha Andrade, S.M. Lima, Influence of synthesis
temperature and atmosphere on Te4+ ion formation in lithium tellurite glass,
Ceram. Int. 47 (2021) 32195–32201, https://doi.org/10.1016/j.
References ceramint.2021.08.112.
[24] A. Hrubý, Evaluation of glass-forming tendency by means of DTA, Czech. J. Phys.
[1] D. Dexter, Possibility of luminescent quantum yields greater than unity, Phys. Rev. 22 (1972) 1187–1193, https://doi.org/10.1007/BF01690134.
108 (1957) 630–633, https://doi.org/10.1103/PhysRev.108.630. [25] S.M. Lima, J.A. Sampaio, T. Catunda, A.C. Bento, L.C.M. Miranda, M.L. Baesso,
[2] R.T. Wegh, H. Donker, K.D. Oskam, A. Meijerink, Visible quantum cutting in Mode-mismatched thermal lens spectrometry for thermo-optical properties
LiGdF4:Eu3+ through downconversion, Science 283 (1999) 663–666, https://doi. measurement in optical glasses: a review, J. Non-Cryst. Solids 273 (2000).
org/10.1126/science.283.5402.663, 80-. [26] S.M. Lima, A.K.R. Souza, A.P. Langaro, J.R. Silva, F.B. Costa, J.C.S. Moraes, M.
[3] B.S. Richards, Enhancing the performance of silicon solar cells via the application S. Figueiredo, F.A. Santos, M.L. Baesso, L.A.O. Nunes, L.H.C. Andrade,
of passive luminescence conversion layers, Sol. Energy Mater. Sol. Cells 90 (2006) Fluorescence quantum yield of Yb3+-doped tellurite glasses determined by thermal
2329–2337, https://doi.org/10.1016/j.solmat.2006.03.035. lens spectroscopy, Opt. Mater. 63 (2017), https://doi.org/10.1016/j.
[4] Z. Liu, J. Li, L. Yang, Q. Chen, Y. Chu, N. Dai, Efficient near-infrared quantum optmat.2016.08.042.
cutting in Ce3+–Yb3+ codoped glass for solar photovoltaic, Sol. Energy Mater. Sol. [27] E.S. Yousef, M.M. Elokr, Y.M. Aboudeif, Optical, elastic properties and DTA of
Cells 122 (2014) 46–50, https://doi.org/10.1016/j.solmat.2013.10.030. TNZP host tellurite glasses doped with Er3+ ions, J. Mol. Struct. 1108 (2016)
[5] D. Chen, Y. Wang, Y. Yu, P. Huang, F. Weng, Quantum cutting downconversion by 257–262, https://doi.org/10.1016/j.molstruc.2015.11.066.
cooperative energy transfer from Ce3+ to Yb3+ in borate glasses, J. Appl. Phys. 104 [28] J.A. Duffy, M.D. Ingram, An interpretation of glass chemistry in terms of the optical
(2008) 88–90, https://doi.org/10.1063/1.3040005. basicity concept, J. Non-Cryst. Solids 21 (1976) 373–410, https://doi.org/
[6] M.B. de la Mora, O. Amelines-Sarria, B.M. Monroy, C.D. Hernández-Pérez, J. 10.1016/0022-3093(76)90027-2.
E. Lugo, Materials for downconversion in solar cells: perspectives and challenges, [29] Y. Liu, Z. Lu, J. Xu, T. Guo, Studies on the influence of structure units on the state
Sol. Energy Mater. Sol. Cells 165 (2017) 59–71, https://doi.org/10.1016/j. of ytterbium ions in TeO2-based glasses, J. Mater. Res. (2020), https://doi.org/
solmat.2017.02.016. 10.1557/jmr.2020.28.
[7] R. Datt, S. Bishnoi, D. Hughes, P. Mahajan, A. Singh, R. Gupta, S. Arya, V. Gupta, [30] T. Sekiya, N. Mochida, A. Ohtsuka, M. Tonokawa, Raman spectra of MO1/2‒TeO2
W.C. Tsoi, Downconversion materials for perovskite solar cells, Sol. RRL. 6 (2022), (M = Li, Na, K, Rb, Cs and Tl) glasses, J. Non-Cryst. Solids 144 (1992) 128–144,
https://doi.org/10.1002/solr.202200266. https://doi.org/10.1016/S0022-3093(05)80393-X.
[8] X. Liu, S. Ye, Y. Qiao, G. Dong, B. Zhu, D. Chen, G. Lakshminarayana, J. Qiu, [31] A. Kaur, A. Khanna, L.I. Aleksandrov, Structural, thermal, optical and photo-
Cooperative downconversion and near-infrared luminescence of Tb3+-Yb3+ luminescent properties of barium tellurite glasses doped with rare-earth ions,
codoped lanthanum borogermanate glasses, Appl. Phys. B Laser Opt. 96 (2009) J. Non-Cryst. Solids 476 (2017) 67–74, https://doi.org/10.1016/j.
51–55, https://doi.org/10.1007/s00340-009-3478-z. jnoncrysol.2017.09.025.
[9] I.A.A. Terra, L.J. Borrero-Gonzalez, J.M. Carvalho, M.C. Terrile, M.C.F.C. Felinto, [32] Y. Himei, A. Osaka, T. Nanba, Y. Miura, Coordination change of Te atoms in binary
H.F. Brito, L.A.O. Nunes, Spectroscopic properties and quantum cutting in Tb3+- tellurite glasses, J. Non-Cryst. Solids 177 (1994) 164–169, https://doi.org/
Yb3+ co-doped ZrO2 nanocrystals, J. Appl. Phys. 113 (2013), https://doi.org/ 10.1016/0022-3093(94)90526-6.
10.1063/1.4792743. [33] M. Udovic, P. Thomas, A. Mirgorodsky, O. Durand, M. Soulis, O. Masson, T. Merle-
[10] L.J. Borrero-Gonzalez, L.A.O. Nunes, J.L. Carmo, F.B.G. Astrath, M.L. Baesso, Méjean, J.C. Champarnaud-Mesjard, Thermal characteristics, Raman spectra and
Spectroscopic studies and downconversion luminescence in OH− -free Pr3+–Yb3+ structural properties of new tellurite glasses within the Bi2O3-TiO2-TeO2 system,
co-doped low-silica calcium aluminosilicate glasses, J. Lumin. 145 (2014) J. Solid State Chem. 179 (2006) 3252–3259, https://doi.org/10.1016/j.
615–619, https://doi.org/10.1016/j.jlumin.2013.08.036. jssc.2006.06.016.
[11] J.T. Van Wijngaarden, S. Scheidelaar, T.J.H. Vlugt, M.F. Reid, A. Meijerink, Energy [34] N. Gupta, A. Kaur, A. Khanna, F. Gonzàlez, C. Pesquera, R. Iordanova, B. Chen,
transfer mechanism for downconversion in the (Pr3+, Yb3+) couple, Phys. Rev. B Structure-property correlations in TiO2-Bi2O3-B2O3-TeO2 glasses, J. Non-Cryst.
Condens. Matter 81 (2010) 1–6, https://doi.org/10.1103/PhysRevB.81.155112. Solids 470 (2017) 168–177, https://doi.org/10.1016/j.jnoncrysol.2017.05.021.
[12] L.J. Borrero-Gonzalez, G. Galleani, D. Manzani, L.A.O. Nunes, S.J.L. Ribeiro, [35] I. Savelii, F. Desevedavy, J.C. Jules, G. Gadret, J. Fatome, B. Kibler, H. Kawashima,
Visible to infrared energy conversion in Pr3+-Yb3+ co-doped fluoroindate glasses, Y. Ohishi, F. Smektala, Management of OH absorption in tellurite optical fibers and
Opt. Mater. 35 (2013) 2085–2089, https://doi.org/10.1016/j. related supercontinuum generation, Opt. Mater. 35 (2013) 1595–1599, https://
optmat.2013.05.024. doi.org/10.1016/j.optmat.2013.04.012.
[13] X. Liu, Y. Qiao, G. Dong, S. Ye, B. Zhu, G. Lakshminarayana, D. Chen, J. Qiu, [36] J. Massera, A. Haldeman, J. Jackson, C. Rivero-Baleine, L. Petit, K. Richardson,
Cooperative downconversion in Yb3+–RE3+ (RE=Tm or Pr) codoped lanthanum Processing of tellurite-based glass with low OH content, J. Am. Ceram. Soc. 94
borogermanate glasses, Opt. Lett. 33 (2008) 2858, https://doi.org/10.1364/ (2011) 130–136, https://doi.org/10.1111/j.1551-2916.2010.04031.x.
ol.33.002858. [37] F.B. Costa, K. Yukimitu, L.A.O. Nunes, L.H. Da Cunha Andrade, S.M. Lima, J.C.
[14] P. Babu, I.R. Martin, V. Lavin, U.R. Rodriguez-Mendoza, H.J. Seo, K. S. Moraesa, Characterization of Nd3+-doped tellurite glasses with low OH content,
V. Krishanaiah, V. Venkatramu, Quantum cutting and near-infrared emissions in Mater. Res. 18 (2015), https://doi.org/10.1590/1516-1439.320614.
Ho3+/Yb3+ codoped transparent glass-ceramics, J. Lumin. 226 (2020), 117424, [38] P.F. Wang, W.N. Li, B. Peng, M. Lu, Effect of dehydration techniques on the
https://doi.org/10.1016/j.jlumin.2020.117424. fluorescence spectral features and OH absorption of heavy metals containing
[15] S.K. Karunakaran, C. Lou, G.M. Arumugam, C. Huihui, D. Pribat, Efficiency fluoride tellurite glasses, J. Non-Cryst. Solids 358 (2012) 788–793, https://doi.
improvement of Si solar cells by down-shifting Ce3+-doped and down-conversion org/10.1016/j.jnoncrysol.2011.12.029.
Ce3+-Yb3+ co-doped YAG phosphors, Sol. Energy 188 (2019) 45–50, https://doi. [39] F. Rivera-López, P. Babu, C. Basavapoornima, C.K. Jayasankar, V. Lavin, Efficient
org/10.1016/j.solener.2019.05.076. Nd3+→Yb3+ energy transfer processes in high phonon energy phosphate glasses for
[16] D.C. Yu, F.T. Rabouw, W.Q. Boon, T. Kieboom, S. Ye, Q.Y. Zhang, A. Meijerink, 1.0 μm Yb3+ laser, J. Appl. Phys. 109 (2011), https://doi.org/10.1063/1.3580475,
Insights into the energy transfer mechanism in Ce3+-Yb3+ codoped YAG phosphors, 0–10.
Phys. Rev. B Condens. Matter 90 (2014) 1–7, https://doi.org/10.1103/ [40] G. Zatryb, M.M. Klak, On the choice of proper average lifetime formula for an
PhysRevB.90.165126. ensemble of emitters showing non-single exponential photoluminescence decay,
[17] A.K. Rufino Souza, A.P. Langaro, J.R. Silva, F.B. Costa, K. Yukimitu, J.C. Silos J. Phys. Condens. Matter 32 (2020), https://doi.org/10.1088/1361-648X/ab9bcc.
Moraes, L. Antonio de Oliveira Nunes, L. Humberto da Cunha Andrade, S.M. Lima, [41] X. Liu, Y. Teng, Y. Zhuang, J. Xie, Y. Qiao, G. Dong, D. Chen, J. Qiu, Broadband
On the efficient Te4+→Yb3+ cooperative energy transfer mechanism in tellurite conversion of visible light to near-infrared emission by Ce3+, Yb3+-codoped
glasses: a potential material for luminescent solar concentrators, J. Alloys Compd. yttrium aluminum garnet, Opt. Lett. 34 (2009) 3565, https://doi.org/10.1364/
781 (2019), https://doi.org/10.1016/j.jallcom.2018.12.038. ol.34.003565.
[18] L. de A. Florencio, L.A. Gomez-Malagon, B.C. Lima, A.S.L. Gomes, J.A.M. Garcia, L. [42] P. Vergeer, T.J.H. Vlugt, M.H.F. Kox, M.I. Den Hertog, J.P.J.M. Van Der Herden,
R.P. Kassab, Efficiency enhancement in solar cells using photon down-conversion A. Meijerink, Quantum cutting by cooperative energy transfer in YbxY 1-xPO4:Tb3+,
in Tb/Yb-doped tellurite glass, Sol. Energy Mater. Sol. Cells 157 (2016) 468–475, Phys. Rev. B Condens. Matter 71 (2005) 1–11, https://doi.org/10.1103/
https://doi.org/10.1016/j.solmat.2016.07.024. PhysRevB.71.014119.
[19] R.A.H. El-Mallawany, Tellurite Glasses Handbook, Second, CRC Press, Boca Raton, [43] A.D. Sontakke, J. Ueda, Y. Katayama, Y. Zhuang, P. Dorenbos, S. Tanabe, Role of
2016, https://doi.org/10.1201/b11295. electron transfer in Ce3+ sensitized Yb3+ luminescence in borate glass, J. Appl.
[20] M.J.V. Bell, V. Anjos, L.M. Moreira, R.F. Falci, L.R.P. Kassab, D.S. da Silva, J. Phys. 117 (2015), https://doi.org/10.1063/1.4905317.
L. Doualan, P. Camy, R. Moncorgé, Laser emission of a Nd-doped mixed tellurite [44] S.M. Lima, A.A. Andrade, R. Lebullenger, A.C. Hernandes, T. Catunda, M.L. Baesso,
and zinc oxide glass, J. Opt. Soc. Am. B 31 (2014) 1590, https://doi.org/10.1364/ Multiwavelength thermal lens determination of fluorescence quantum efficiency of
josab.31.001590. solids: application to Nd3+-doped fluoride glass, Appl. Phys. Lett. 78 (2001),
[21] F.B. Costa, A.K.R. Souza, A.P. Langaro, J.R. Silva, F.A. Santos, M.S. Figueiredo, https://doi.org/10.1063/1.1375000.
K. Yukimitu, J.C.S. Moraes, L.A.O. Nunes, L.H.C. Andrade, S.M. Lima, Observation

19479
A.K. Rufino Souza et al. Ceramics International 49 (2023) 19470–19480

[45] J. Shen, R.D. Lowe, R.D. Snook, A model for cw laser induced mode-mismatched methods, J. Non-Cryst. Solids 352 (2006), https://doi.org/10.1016/j.
dual-beam thermal lens spectrometry, Chem. Phys. 165 (1992) 385–396, https:// jnoncrysol.2006.02.116.
doi.org/10.1016/0301-0104(92)87053-C. [47] A.C. Bento, N. Cella, S.M. Lima, L.A.O. Nunes, L.H.C. Andrade, J.R. Silva, V.
[46] S.M. Lima, W.F. Falco, E.S. Bannwart, L.H.C. Andrade, R.C. de Oliveira, J.C. S. Zanuto, N.G.C. Astrath, T. Catunda, A.N. Medina, J.H. Rohling, R.F. Muniz, J.
S. Moraes, K. Yukimitu, E.B. Araujo, E.A. Falcao, A. Steimacher, N.G.C. Astrath, A. W. Berrar, L.C. Malacarne, W.R. Weinand, F. Sato, M.P. Belancon, G.J. Schiavon,
C. Bento, A.N. Medina, M.L. Baesso, Thermo-optical characterization of tellurite J. Shen, L.C.M. Miranda, H. Vargas, M.L. Baesso, Photoacoustic and photothermal
glasses by thermal lens, thermal relaxation calorimetry and interferometric and the photovoltaic efficiency of solar cells: a tutorial, J. Appl. Phys. 131 (2022),
https://doi.org/10.1063/5.0088211.

19480

You might also like