[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (16)

Search Parameters:
Keywords = scintillation quenching

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 42919 KiB  
Article
Effect of Eu Ions Concentration in Y2O3-Based Transparent Ceramics on the Electron Irradiation Induced Luminescence and Damage
by Wenhui Lou, Yang Tang, Haohong Chen, Yisong Lei, Hui Lin, Ruijin Hong, Zhaoxia Han and Dawei Zhang
Materials 2024, 17(20), 4954; https://doi.org/10.3390/ma17204954 (registering DOI) - 10 Oct 2024
Viewed by 243
Abstract
Eu3+-doped Y2O3-based luminescent materials can be used as a scintillator for electron or high energy β-ray irradiation, which are essential for applications such as electron microscopy and nuclear batteries. Therefore, it is essential to understand their defect [...] Read more.
Eu3+-doped Y2O3-based luminescent materials can be used as a scintillator for electron or high energy β-ray irradiation, which are essential for applications such as electron microscopy and nuclear batteries. Therefore, it is essential to understand their defect mechanisms and to develop materials with excellent properties. In this paper, Y2O3-based transparent ceramics with different Eu3+ doping concentrations were prepared by solid-state reactive vacuum sintering. This series of transparent ceramic samples exhibits strong red emission under electron beam excitation at the keV level. However, color change appears after the high-energy electron irradiation due to the capture of electrons by the traps in the Y2O3 lattice. Optical transmittance, laser-excited luminescence, X-ray photoelectron spectroscopy (XPS), and other analyses indicated that the traps, or the color change, mainly originate from the residual oxygen vacancies, which can be suppressed by high Eu doping. Seen from the cathodoluminescence (CL) spectra, higher doping concentrations of Eu3+ showed stronger resistance to electron irradiation damage, but also resulted in lower emissions due to concentration quenching. Therefore, 10% doping of Eu was selected in this work to keep the high emission intensity and strong radiation resistance both. This work helps to enhance the understanding of defect formation mechanisms in the Y2O3 matrix and will be of benefit for the modification of scintillation properties for functional materials systems. Full article
(This article belongs to the Section Advanced and Functional Ceramics and Glasses)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD <span class="html-italic">θ</span>-2<span class="html-italic">θ</span> scanning patterns (<b>a</b>); and enlarged view in the range of 28–30° (<b>b</b>) of the (Eu<sub>x</sub>Y<sub>0.97−x</sub>Zr<sub>0.03</sub>)<sub>2</sub>O<sub>3</sub> (x = 0.001, 0.01, 0.1, and 0.2) ceramic samples.</p>
Full article ">Figure 2
<p>SEM surface morphology of the (Eu<sub>x</sub>Y<sub>0.97−x</sub>Zr<sub>0.03</sub>)<sub>2</sub>O<sub>3</sub> (x = 0.001, 0.01, 0.1, and 0.2) ceramic samples: (<b>a</b>) x = 0.001; (<b>b</b>) x = 0.01; (<b>c</b>) x = 0.1; and (<b>d</b>) x = 0.2.</p>
Full article ">Figure 3
<p>Optical transmission spectra (<b>a</b>); and the calculated optical bandgap energy (<b>b</b>) of the (Eu<sub>x</sub>Y<sub>0.97−x</sub>Zr<sub>0.03</sub>)<sub>2</sub>O<sub>3</sub> (x = 0.001, 0.01, 0.1, and 0.2) ceramic samples.</p>
Full article ">Figure 4
<p>PLE (<b>a</b>); and PL (<b>b</b>) spectra of the (Eu<sub>x</sub>Y<sub>0.97−x</sub>Zr<sub>0.03</sub>)<sub>2</sub>O<sub>3</sub> (x = 0.001, 0.01, 0.1, and 0.2) ceramic samples.</p>
Full article ">Figure 5
<p>CL (<b>a</b>); and TL (<b>b</b>) spectra of (Eu<sub>x</sub>Y<sub>0.97−x</sub>Zr<sub>0.03</sub>)<sub>2</sub>O<sub>3</sub> (x = 0.001, 0.01, 0.1, and 0.2) ceramic samples.</p>
Full article ">Figure 6
<p>Transmittance (<b>a</b>); and PL (<b>b</b>) spectra of Eu<sup>3+</sup>:Y<sub>2</sub>O<sub>3</sub> ceramic samples under the 450 nm excitation after accelerated electron irradiation.</p>
Full article ">Figure 7
<p>Schematic diagram to illustrate the mechanism of the electron beam irradiation induced color center as well as the red emission.</p>
Full article ">Figure 8
<p>XPS comparison for 0.1% Eu<sup>3+</sup>:Y<sub>2</sub>O<sub>3</sub> ceramic before and after irradiation.</p>
Full article ">Figure 9
<p>O 1s XPS signals fitting of (Eu<sub>x</sub>Y<sub>0.97−x</sub>Zr<sub>0.03</sub>)<sub>2</sub>O<sub>3</sub> (x = 0.001, 0.01, 0.1, and 0.2) ceramic samples before and after irradiation: (<b>a</b>,<b>b</b>) x = 0.001; (<b>c</b>,<b>d</b>) x = 0.01; (<b>e</b>,<b>f</b>) x = 0.1; and (<b>g</b>,<b>h</b>) x = 0.2.</p>
Full article ">Figure 9 Cont.
<p>O 1s XPS signals fitting of (Eu<sub>x</sub>Y<sub>0.97−x</sub>Zr<sub>0.03</sub>)<sub>2</sub>O<sub>3</sub> (x = 0.001, 0.01, 0.1, and 0.2) ceramic samples before and after irradiation: (<b>a</b>,<b>b</b>) x = 0.001; (<b>c</b>,<b>d</b>) x = 0.01; (<b>e</b>,<b>f</b>) x = 0.1; and (<b>g</b>,<b>h</b>) x = 0.2.</p>
Full article ">
26 pages, 10957 KiB  
Article
Micro-Inclusion Engineering via Sc Incompatibility for Luminescence and Photoconversion Control in Ce3+-Doped Tb3Al5−xScxO12 Garnet
by Karol Bartosiewicz, Robert Tomala, Damian Szymański, Benedetta Albini, Justyna Zeler, Masao Yoshino, Takahiko Horiai, Paweł Socha, Shunsuke Kurosawa, Kei Kamada, Pietro Galinetto, Eugeniusz Zych and Akira Yoshikawa
Materials 2024, 17(11), 2762; https://doi.org/10.3390/ma17112762 - 5 Jun 2024
Viewed by 819
Abstract
Aluminum garnets display exceptional adaptability in incorporating mismatching elements, thereby facilitating the synthesis of novel materials with tailored properties. This study explored Ce3+-doped Tb3Al5−xScxO12 crystals (where x ranges from 0.5 to 3.0), revealing a [...] Read more.
Aluminum garnets display exceptional adaptability in incorporating mismatching elements, thereby facilitating the synthesis of novel materials with tailored properties. This study explored Ce3+-doped Tb3Al5−xScxO12 crystals (where x ranges from 0.5 to 3.0), revealing a novel approach to control luminescence and photoconversion through atomic size mismatch engineering. Raman spectroscopy confirmed the coexistence of garnet and perovskite phases, with Sc substitution significantly influencing the garnet lattice and induced A1g mode softening up to Sc concentration x = 2.0. The Sc atoms controlled sub-eutectic inclusion formation, creating efficient light scattering centers and unveiling a compositional threshold for octahedral site saturation. This modulation enabled the control of energy transfer dynamics between Ce3+ and Tb3+ ions, enhancing luminescence and mitigating quenching. The Sc admixing process regulated luminous efficacy (LE), color rendering index (CRI), and correlated color temperature (CCT), with adjustments in CRI from 68 to 84 and CCT from 3545 K to 12,958 K. The Ce3+-doped Tb3Al5−xScxO12 crystal (where x = 2.0) achieved the highest LE of 114.6 lm/W and emitted light at a CCT of 4942 K, similar to daylight white. This approach enables the design and development of functional materials with tailored optical properties applicable to lighting technology, persistent phosphors, scintillators, and storage phosphors. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>As-grown rods and polished radial plates of Tb<sub>2.85</sub>Ce<sub>0.15</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals with increasing Sc<sup>3+</sup> ion concentration.</p>
Full article ">Figure 2
<p>Theoretical [Tb<sub>3</sub>(Al,Sc)<sub>5</sub>O<sub>12</sub>—#PDF 53-0273; TbScO<sub>3</sub>—#PDF27-0599; TbAlO<sub>3</sub>—#PDF 24-127) and experimental PXRD patterns of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5, 1.0, 1.5, 2.0 and 3.0.</p>
Full article ">Figure 3
<p>SEM images and corresponding EDS elemental distribution maps of terbium (Tb, yellow), scandium (Sc, cyan), aluminum (Al, green), and oxygen (O, red) of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5, 1.0, 1.5, 2.0, and 3.0.</p>
Full article ">Figure 4
<p>EDS line scanning profiles of terbium (Tb, yellow), scandium (Sc, cyan), aluminum (Al, green), and oxygen (O, red) recorded along a diameter of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5, 1.0, 1.5, 2.0, and 3.0.</p>
Full article ">Figure 5
<p>(<b>a</b>) Raman spectra acquired at an excitation wavelength of 638 nm probing the secondary phase inclusions present in the Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5 and 3.0 (300 K); (<b>b</b>) Raman spectra acquired at an excitation wavelength of 785 nm, probing the garnet matrix of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5, 1.0, 1.5, 2.0, and 3.0; (<b>c</b>) the <span class="html-italic">A<sub>1g</sub></span> mode peak energy position at 770 cm<sup>−1</sup> for Tb<sub>2.85</sub>Ce<sub>0.15</sub>Al<sub>4.5</sub>Sc<sub>0.5</sub>O<sub>12</sub> as a function of Sc<sup>3+</sup> ion concentration.</p>
Full article ">Figure 6
<p>Evolution of TL glow curves in Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5, 1.0, 1.5, 2.0, and 3.0.</p>
Full article ">Figure 7
<p>Unpolarized optical absorption spectra of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub>, where x = 0.5, 1.0, 1.5, 2.0, and 3.0 (300 K). The absorption spectra exhibit saturation below 300 nm due to the combined effects of the high Tb<sup>3+</sup> concentration (95.5% dodecahedral site occupancy), spin-allowed <span class="html-italic">4f</span> → <span class="html-italic">5d</span> transitions, and local environmental variations within the multiphase structure. The intense background for the Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystal where x = 3.0 arises from increased light scattering caused by reduced crystal transparency.</p>
Full article ">Figure 8
<p>RT excitation spectra of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5, 1.0, 1.5, 2.0, and 3.0 recorded at 550 nm corresponding to the <span class="html-italic">5d<sub>1</sub></span> → <span class="html-italic">4f</span> transition of Ce<sup>3+</sup> ions (300 K).</p>
Full article ">Figure 9
<p>Temperature-dependent photoluminescence (<b>a</b>–<b>e</b>) emission spectra (λ<sub>exc</sub> = 455 nm) and (<b>f</b>) emission spectra integrals between 480 and 750 nm of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, as measured at temperatures ranging from 83 K to 683 K, exhibiting the impact of increasing Sc<sup>3+</sup> ion concentration.</p>
Full article ">Figure 10
<p>(<b>a</b>) Selected decay curves of the emission at 560 nm (λ<sub>exc</sub> = 455 nm) for Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5, 1.5, and 3.0 recorded at 83 K; (<b>b</b>) dependence of photoluminescence decay times on temperature for emission at 560 nm (λ<sub>exc</sub> = 455 nm) in Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> with increasing Sc<sup>3+</sup> ion concentration.</p>
Full article ">Figure 11
<p>The photoconversion spectra of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals delineate the impact of varying Sc<sup>3+</sup> ion concentrations under (<b>a</b>) 445 nm laser diode and (<b>b</b>) 455 nm light-emitting diode excitations; (<b>c</b>) representation of observed emission spectra for LED and LD plotted on the CIE 1931 chromaticity diagram.</p>
Full article ">Figure 12
<p>Radioluminescence spectra of Ce<sup>3+</sup>-doped Tb<sub>3</sub>Al<sub>5−x</sub>Sc<sub>x</sub>O<sub>12</sub> crystals, where x = 0.5, 1.0, 1.5, 2.0, and 3.0.</p>
Full article ">
13 pages, 3456 KiB  
Article
High-Density Glass Scintillators for Proton Radiography—Relative Luminosity, Proton Response, and Spatial Resolution
by Ethan Stolen, Ryan Fullarton, Rain Hein, Robin L. Conner, Luiz G. Jacobsohn, Charles-Antoine Collins-Fekete, Sam Beddar, Ugur Akgun and Daniel Robertson
Sensors 2024, 24(7), 2137; https://doi.org/10.3390/s24072137 - 27 Mar 2024
Cited by 1 | Viewed by 1284
Abstract
Proton radiography is a promising development in proton therapy, and researchers are currently exploring optimal detector materials to construct proton radiography detector arrays. High-density glass scintillators may improve integrating-mode proton radiography detectors by increasing spatial resolution and decreasing detector thickness. We evaluated several [...] Read more.
Proton radiography is a promising development in proton therapy, and researchers are currently exploring optimal detector materials to construct proton radiography detector arrays. High-density glass scintillators may improve integrating-mode proton radiography detectors by increasing spatial resolution and decreasing detector thickness. We evaluated several new scintillators, activated with europium or terbium, with proton response measurements and Monte Carlo simulations, characterizing relative luminosity, ionization quenching, and proton radiograph spatial resolution. We applied a correction based on Birks’s analytical model for ionization quenching. The data demonstrate increased relative luminosity with increased activation element concentration, and higher relative luminosity for samples activated with europium. An increased glass density enables more compact detector geometries and higher spatial resolution. These findings suggest that a tungsten and gadolinium oxide-based glass activated with 4% europium is an ideal scintillator for testing in a full-size proton radiography detector. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental setup diagram on the proton therapy gantry at Mayo Clinic Arizona (<b>left</b>), and the experimental plate and scintillators as modeled in TOPAS MC (<b>right</b>).</p>
Full article ">Figure 2
<p>Image of the scintillator samples during irradiation, taken by the CMOS camera at a water-equivalent depth of 132 mm, the position of the Bragg peak. The ROIs used to measure average pixel intensity are outlined in blue.</p>
Full article ">Figure 3
<p>The emission spectra of the scintillator samples and BGO powder.</p>
Full article ">Figure 4
<p>Depth–light curve for sample S6, which displayed the highest relative luminosity. The normalized depth–light curve (black) and corrected depth–dose curve (red) is compared to the Monte Carlo calculated depth–dose curve (blue).</p>
Full article ">Figure 5
<p>Depth–dose curves of a 200 MeV proton beam in Sample S1 (tungsten and gadolinium oxide glass), sample S5 (lead borate glass), and EJ-260 (organic plastic scintillator) (<b>left</b>). Lateral profiles at the Bragg peak of a 200 MeV proton beam for the three samples (<b>right</b>).</p>
Full article ">Figure 6
<p>A 200 MeV proton field with dimensions 15 × 15 cm<sup>2</sup> transported through a 4 cm thickness Al slab inside a 10 cm thickness water tank, with the residual proton energy being absorbed in a block of EJ-260 organic plastic scintillator (<b>upper left</b>), lead oxide glass scintillator (<b>upper right</b>), and tungsten and gadolinium oxide glass scintillator (<b>lower left</b>). The resulting modulation transfer function curves are shown (<b>lower right</b>).</p>
Full article ">
11 pages, 3748 KiB  
Article
Photoluminescent and Scintillating Performance of Eu3+-Doped Boroaluminosilicate Glass Scintillators
by Yujia Gong, Lianjie Li, Junyu Chen and Hai Guo
Materials 2023, 16(13), 4711; https://doi.org/10.3390/ma16134711 - 29 Jun 2023
Cited by 4 | Viewed by 1309
Abstract
In comparison with single crystal scintillators, glass scintillators are more promising materials for their benefits of easy preparation, low cost, controllable size, and large-scale manufacture. The emission of Eu3+ ion at 612 nm matches well with the photoelectric detector, making it suitable [...] Read more.
In comparison with single crystal scintillators, glass scintillators are more promising materials for their benefits of easy preparation, low cost, controllable size, and large-scale manufacture. The emission of Eu3+ ion at 612 nm matches well with the photoelectric detector, making it suitable for the activator in glass scintillators. Therefore, the research on Eu3+ doped glass scintillators attract our attention. The photoluminescent and scintillating properties of Eu3+-activated boroaluminosilicate glass scintillators prepared by the conventional melt-quenching method were investigated in this work. The glass samples present good internal quantum yield. Under X-ray radiation, the optimal sample reveals high X-ray excited luminesce (XEL), and its integrated intensity of XEL is 22.7% of that of commercial crystal scintillator Bi4Ge3O12. Furthermore, the optimal specimen possesses a spatial resolution of 14 lp/mm in X-ray imaging. These results suggest that Eu3+-doped boroaluminosilicate glass is expected to be applied in X-ray imaging. Full article
(This article belongs to the Special Issue Glasses and Ceramics for Luminescence Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns of SBAY host and SBAY:<span class="html-italic">x</span>Eu specimens, (<b>b</b>) FT-IR spectra of SBAY host and SBAY:8Eu specimens.</p>
Full article ">Figure 2
<p>(<b>a</b>) Transmittance spectra of SBAY host and SBAY:<span class="html-italic">x</span>Eu specimens, (<b>b</b>) the relationship of <span class="html-italic">α</span><sup>2</sup> with photon energy for SBAY host.</p>
Full article ">Figure 3
<p>(<b>a</b>) PL spectra (λ<sub>ex</sub> = 464 nm), (<b>b</b>) PLE spectra (λ<sub>em</sub> = 612 nm) of SBAY:<span class="html-italic">x</span>Eu specimens.</p>
Full article ">Figure 4
<p>Excitation lines of BaSO<sub>4</sub> reference, and PL (λ<sub>ex</sub> = 464 nm) spectra of SBAY:<span class="html-italic">x</span>Eu. The inset shows the magnification of PL spectra of SBAY:<span class="html-italic">x</span>Eu.</p>
Full article ">Figure 5
<p>Fluorescent lifetime curves of 612 nm emission of SBAY:<span class="html-italic">x</span>Eu specimens.</p>
Full article ">Figure 6
<p>XEL spectra of SBAY:<span class="html-italic">x</span>Eu specimens and BGO crystal with different Eu<sup>3+</sup> content.</p>
Full article ">Figure 7
<p>XEL mechanism diagram for SBAY:<span class="html-italic">x</span>Eu samples, here CB and VB are conduction band and valence band, respectively.</p>
Full article ">Figure 8
<p>(<b>a</b>) The XEL spectra (measured at five-minute interval) of SBAY:8Eu specimen radiated continuously for 60 min by X-ray (6 W), and (<b>b</b>) time-dependent integrated XEL intensities for SBAY:8Eu specimen.</p>
Full article ">Figure 9
<p>(<b>a</b>) XEL spectra of SBAY:8Eu specimen under X-ray radiation with different input power, (<b>b</b>) transmittance spectra of SBAY:8Eu specimen after X-ray irradiation with different input power.</p>
Full article ">Figure 10
<p>Photos of chip (<b>a</b>), metallic spring in capsule (<b>c</b>), and standard X-ray test pattern plate (<b>e</b>) under daylight. X-ray images of chip (<b>b</b>), metallic spring in capsule (<b>d</b>), and standard X-ray test pattern plate (<b>f</b>) based on SBAY:8Eu.</p>
Full article ">
21 pages, 12182 KiB  
Communication
Polyaromatic Hydrocarbon (PAH)-Based Aza-POPOPs: Synthesis, Photophysical Studies, and Nitroanalyte Sensing Abilities
by Mohammed S. Mohammed, Igor S. Kovalev, Natalya V. Slovesnova, Leila K. Sadieva, Vadim A. Platonov, Alexander S. Novikov, Sougata Santra, Julia E. Morozova, Grigory V. Zyryanov, Valery N. Charushin and Brindaban C. Ranu
Int. J. Mol. Sci. 2023, 24(12), 10084; https://doi.org/10.3390/ijms241210084 - 13 Jun 2023
Cited by 1 | Viewed by 1444
Abstract
1,4-Bis(5-phenyl-2-oxazolyl)benzene (POPOP) is a common scintillation fluorescent laser dye. In this manuscript, the synthesis of 2-Ar-5-(4-(4-Ar’-1H-1,2,3-triazol-1-yl)phenyl)-1,3,4-oxadiazoles (Ar, Ar’ = Ph, naphtalenyl-2, pyrenyl-1, triphenilenyl-2), as PAH-based aza-analogues of POPOP, by means of Cu-catalyzed click reaction between 2-(4-azidophenyl)-5-Ar-1,3,4-oxadiazole and terminal ethynyl-substituted PAHs [...] Read more.
1,4-Bis(5-phenyl-2-oxazolyl)benzene (POPOP) is a common scintillation fluorescent laser dye. In this manuscript, the synthesis of 2-Ar-5-(4-(4-Ar’-1H-1,2,3-triazol-1-yl)phenyl)-1,3,4-oxadiazoles (Ar, Ar’ = Ph, naphtalenyl-2, pyrenyl-1, triphenilenyl-2), as PAH-based aza-analogues of POPOP, by means of Cu-catalyzed click reaction between 2-(4-azidophenyl)-5-Ar-1,3,4-oxadiazole and terminal ethynyl-substituted PAHs is reported. An investigation of the photophysical properties of the obtained products was carried out, and their sensory response to nitroanalytes was evaluated. In the case of pyrenyl-1-substituted aza-POPOP, dramatic fluorescence quenching by nitroanalytes was observed. Full article
Show Figures

Figure 1

Figure 1
<p>POPOP and its <span class="html-italic">aza</span>-analog core. POPOP = 1,4-Bis(5-phenyl-2-oxazolyl)benzene.</p>
Full article ">Figure 2
<p>Absorption (<b>A</b>) and emission (<b>B</b>) spectra of POPOP and its <span class="html-italic">aza</span>-analogues <b>3</b> in CH<sub>2</sub>Cl<sub>2</sub> (10<sup>−5</sup> M). POPOP = 1,4-Bis(5-phenyl-2-oxazolyl)benzene.</p>
Full article ">Figure 3
<p>Stern–Volmer plot (<b>A</b>) and overlayed graph (<b>B</b>) of the chemosensor <b>3g</b> fluorescence quenching by PETN. PETN = Pentaerythritol tetranitrate.</p>
Full article ">Figure 4
<p>Energy diagram of the PET quenching process. PET = photon-induced electron transfer; PETN = Pentaerythritol tetranitrate.</p>
Full article ">Figure 5
<p>Normalized emission spectra of aza analogues POPOP. POPOP = 1,4-Bis(5-phenyl-2-oxazolyl)benzene.</p>
Full article ">Figure 6
<p>Fluorescence quenching of chemosensor <b>3g</b> by TNT. TNT = 2,4,6-Trinitrotoluene.</p>
Full article ">Figure 7
<p>Fluorescence quenching of chemosensor <b>3g</b> by DNT. DNT = 2,4-Dinitrotoluene.</p>
Full article ">Figure 8
<p>Fluorescence quenching of chemosensor <b>3g</b> by PETN. PETN = Pentaerythritol tetranitrate.</p>
Full article ">Figure 9
<p>Overlayed graph of the chemosensor <b>3g</b> quenching by DNT (UV-Vis). DNT = 2,4-Dinitrotoluene.</p>
Full article ">Figure 10
<p>Overlayed graph of the chemosensor <b>3g</b> quenching by DNT (Emission). DNT = 2,4-Dinitrotoluene.</p>
Full article ">Figure 11
<p>Overlayed graph of the chemosensor <b>3g</b> quenching by TNT (UV-Vis). TNT = 2,4,6-Trinitrotoluene.</p>
Full article ">Figure 12
<p>Overlayed graph of the chemosensor <b>3g</b> quenching by TNT (Emission). TNT = 2,4,6-Trinitrotoluene.</p>
Full article ">Figure 13
<p>Overlayed graph of the chemosensor <b>3g</b> quenching by PETN (UV-Vis). PETN = Pentaerythritol tetranitrate.</p>
Full article ">Figure 14
<p>Overlayed graph of the chemosensor <b>3g</b> quenching by PETN (Emission). PETN = Pentaerythritol tetranitrate.</p>
Full article ">Figure 15
<p>Normalized emission and absorption spectra <b>3a</b>.</p>
Full article ">Figure 16
<p>Normalized emission and absorption spectra <b>3b</b>.</p>
Full article ">Figure 17
<p>Normalized emission and absorption spectra <b>3c</b>.</p>
Full article ">Figure 18
<p>Normalized emission and absorption spectra <b>3d</b>.</p>
Full article ">Figure 19
<p>Normalized emission and absorption spectra <b>3e</b>.</p>
Full article ">Figure 20
<p>Normalized emission and absorption spectra <b>3f</b>.</p>
Full article ">Figure 21
<p>Normalized emission and absorption spectra <b>3g</b>.</p>
Full article ">Figure 22
<p>Excitation spectrum <b>3a</b>.</p>
Full article ">Figure 23
<p>Excitation spectrum <b>3b</b>.</p>
Full article ">Figure 24
<p>Excitation spectrum <b>3c</b>.</p>
Full article ">Figure 25
<p>Excitation spectrum <b>3d</b>.</p>
Full article ">Figure 26
<p>Excitation spectrum <b>3e</b>.</p>
Full article ">Figure 27
<p>Excitation spectrum <b>3f</b>.</p>
Full article ">Figure 28
<p>Excitation spectrum <b>3g</b>.</p>
Full article ">Scheme 1
<p>Synthesis of the azido components <b>2a</b>,<b>b</b>. PTSA = <span class="html-italic">p</span>-Toluenesulfonic acid.</p>
Full article ">Scheme 2
<p>Synthesis of the POPOP analogues <b>3</b>.</p>
Full article ">
14 pages, 6029 KiB  
Article
On the Quenching Mechanism of Ce, Tb Luminescence and Scintillation in Compositionally Disordered (Gd, Y, Yb)3Al2Ga3O12 Garnet Ceramics
by Valery Dubov, Daria Kuznetsova, Irina Kamenskikh, Ilia Komendo, Georgii Malashkevich, Andrei Ramanenka, Vasili Retivov, Yauheni Talochka, Andrei Vasil’ev and Mikhail Korzhik
Photonics 2023, 10(6), 615; https://doi.org/10.3390/photonics10060615 - 26 May 2023
Cited by 3 | Viewed by 1502
Abstract
Two series of (Gd, Y, Yb)3Al2Ga3O12 quintuple compounds with a garnet structure and solely doped with Ce and Tb were prepared in the form of ceramics by sintering in oxygen at 1600 °C for two hours [...] Read more.
Two series of (Gd, Y, Yb)3Al2Ga3O12 quintuple compounds with a garnet structure and solely doped with Ce and Tb were prepared in the form of ceramics by sintering in oxygen at 1600 °C for two hours and studied for the interaction of activator ions with ytterbium ions entering the matrix. It was shown that the photoluminescence and scintillation of Ce3+ ions are completely suppressed, predominantly by tunneling ionization through the charge transfer state (CTS) of the Ce4+-Yb2+ ions. The photoluminescence of Tb3+ ions is quenched in the presence of ytterbium, but not completely due to the poor resonance conditions of Tb3+ intraconfiguration transitions and the CTS of the single Yb3+ and the CTS of Ce4+-Yb2+ ions. The scintillation in the visible range of both Ce3+- and Tb3+-doped samples is intensely quenched as well, which indicates strong competition from Yb3+ ions to activators in interaction with the Gd substrate. Full article
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction patterns of cerium/terbium-activated ceramic samples with different Gd:Yb:Y ratios.</p>
Full article ">Figure 2
<p>SEM images of representative (indicated) ceramic samples: (<b>a</b>)—Gd<sub>2.85</sub>Al<sub>2</sub>Ga<sub>2.97</sub>O<sub>12</sub>Ce<sub>0.015</sub>, (<b>b</b>)—Gd<sub>1.485</sub>Y<sub>0.75</sub>Yb<sub>0.75</sub>Al<sub>2</sub>Ga<sub>2.91</sub>O<sub>12</sub>Ce<sub>0.015</sub>, (<b>c</b>)—Gd<sub>1.485</sub>Y<sub>1.2</sub>Yb<sub>0.3</sub>Al<sub>2</sub>Ga<sub>2.91</sub>O<sub>12</sub>Ce<sub>0.015</sub>.</p>
Full article ">Figure 3
<p>Images of the samples presented in <a href="#photonics-10-00615-t001" class="html-table">Table 1</a> in daylight (<b>a</b>); excitation using UV radiation with λ = 365 nm (<b>b</b>) and λ = 250 nm (<b>c</b>).</p>
Full article ">Figure 4
<p>Luminescence excitation spectra of representative Ce (<b>a</b>)- and Tb (<b>b</b>)-doped samples at the registration of 1023 nm; room temperature measured luminescence spectra in the IR region at 365 nm excitation of the same Ce (<b>c</b>)- and Tb (<b>d</b>)-doped samples.</p>
Full article ">Figure 5
<p>Luminescence spectra of Gd1.5Yb0.3Y1.2-Ce (#6) at excitation 225 nm at different temperatures.</p>
Full article ">Figure 6
<p>Deconvolution of the experimental luminescence band contour (blue line) in the spectral region 450–700 nm into the total band of Ce<sup>3+</sup> ions (red line) and three CTS→<sup>2</sup>F<sub>5/2</sub> bands (black lines) at a temperature of 160 K. The total contour of the band due to transitions with an involvement of Ce<sup>3+</sup> and Yb<sup>3+</sup> ions is superimposed on the experimental curve (orange line).</p>
Full article ">Figure 7
<p>Temperature dependence of the integrated luminescence intensity of the bands in the Gd1.5Yb0.3Y1.2-Ce (#6) sample: (<b>a</b>)—Ce<sup>3+</sup>: 4f<sup>0</sup>5d<sub>1</sub> →<sup>2</sup>F<sub>7/2,5/2</sub> (red), Yb<sup>3+</sup>: CTS→<sup>2</sup>F<sub>5/2</sub> (blue), and CTS→<sup>2</sup>F<sub>7/2</sub> (black); (<b>b</b>)—intracenter Yb<sup>3+</sup> NIR luminescence; (<b>c</b>)—Gd<sup>3+</sup>: <sup>6</sup>P<sub>J</sub>→<sup>8</sup>S<sub>7/2</sub>. Solid lines at the panels (<b>a</b>,<b>c</b>) are the approximating curves.</p>
Full article ">Figure 8
<p>Configuration diagrams for isolated Ce<sup>3+</sup> (<b>a</b>), Yb<sup>3+</sup> (<b>b</b>) ions. Electronic transitions caused by photon absorption and emission are shown by vertical arrows.</p>
Full article ">Figure 9
<p>Configuration diagram for Ce + Yb complex. Electronic transitions caused by photon absorption and emission are shown by vertical arrows. Points A, B, and C correspond to the minima of the ground state, relaxed Ce<sup>3+*</sup>, and relaxed position of Ce<sup>4+</sup> + Yb<sup>2+</sup> of the dynamic complex, respectively. Red and green lines indicate intersection of configuration surfaces where electron passes from cerium to ytterbium (red) and returns to cerium (green). Blue arrows indicate relaxation of the system in Ce<sup>3+</sup> + Yb<sup>3+</sup> state, whereas cyan arrows show the relaxation pathway in dynamic Ce<sup>4+</sup> + Yb<sup>2+</sup> state.</p>
Full article ">
14 pages, 2968 KiB  
Article
Luminescence Characteristics of the MOCVD GaN Structures with Chemically Etched Surfaces
by Tomas Ceponis, Jevgenij Pavlov, Arunas Kadys, Augustas Vaitkevicius and Eugenijus Gaubas
Materials 2023, 16(9), 3424; https://doi.org/10.3390/ma16093424 - 27 Apr 2023
Viewed by 1370
Abstract
Gallium nitride is a wide-direct-bandgap semiconductor suitable for the creation of modern optoelectronic devices and radiation tolerant detectors. However, formation of dislocations is inevitable in MOCVD GaN materials. Dislocations serve as accumulators of point defects within space charge regions covering cores of dislocations. [...] Read more.
Gallium nitride is a wide-direct-bandgap semiconductor suitable for the creation of modern optoelectronic devices and radiation tolerant detectors. However, formation of dislocations is inevitable in MOCVD GaN materials. Dislocations serve as accumulators of point defects within space charge regions covering cores of dislocations. Space charge regions also may act as local volumes of enhanced non-radiative recombination, deteriorating the photoluminescence efficiency. Surface etching has appeared to be an efficient means to increase the photoluminescence yield from MOCVD GaN materials. This work aimed to improve the scintillation characteristics of MOCVD GaN by a wet etching method. An additional blue photo-luminescence (B-PL) band peaking at 2.7–2.9 eV and related to dislocations was discovered. This B-PL band intensity appeared to be dependent on wet etching exposure. The intensity of the B-PL was considerably enhanced when recorded at rather low temperatures. This finding resembles PL thermal quenching of B-PL centers. The mechanisms of scintillation intensity and spectrum variations were examined by coordinating the complementary photo-ionization and PL spectroscopy techniques. Analysis of dislocation etch pits was additionally performed by scanning techniques, such as confocal and atomic force microscopy. It was proved that this blue luminescence band, which peaked at 2.7–2.9 eV, is related to point defects those decorate dislocation cores. It was shown that the intensity of this blue PL band was increased due to enhancement of light extraction efficiency, dependent on the surface area of either single etch-pit or total etched crystal surface. Full article
(This article belongs to the Section Materials Physics)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The photo-activation centers (E1–E6) resolved from the photo-ionization spectral steps, measured on MOCVD GaN (scattered circles-data) and simulated using Kopylov–Pikhtin [<a href="#B32-materials-16-03424" class="html-bibr">32</a>] approach (Equation (1))-dashed lines. These traps were tentatively identified using the activation energy values taken from the literature referenced. The peak values of spectral bands were adjusted by varying the trap concentration parameters. Arrows denote the photon energy for fixed wavelength excitation of photoluminescence employed in these experiments. The gray rectangles indicate a relative efficiency of PL excitation by intersection of photon energy with spectral shapes of the identified photo-ionization centers. (<b>b</b>) PL spectrum (black circle symbols) recorded on MOCVD GaN material at room temperature using the multi-functional WITEC system Alpha 300 in confocal microscopy mode. This measured spectrum was fitted (gray solid line) using the van Roosbroeck–Shockley model (Equation (2)). This simulated gray solid line represents a resultant PL spectrum of simultaneous action of photoactive centers revealed by pulsed photo-ionization spectroscopy (PPIS) techniques. The contribution of each center (dashed lines) was simulated using parameters extracted from PPIS analysis. Fitting of the experimental PL spectrum (scattered circles) by a simulated resultant PL spectrum (solid gray curve) was implemented by slightly adjusting the contribution of various centers (dashed lines). The latter procedure was performed by varying the peak amplitudes (related to concentrations of definite centers).</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) Evolution of optical microscopy (OM) images dependent on etching exposure. (<b>d</b>–<b>f</b>) PL spectra measured at different sample temperatures and dependent on etching exposure duration (<span class="html-italic">t</span>) when an H<sub>3</sub>PO<sub>4</sub> temperature of 150 °C was kept over etching procedures. A sketch (curve 6) of the PL spectrum typical for the MOCVD grown GaN is taken from Ref. [<a href="#B43-materials-16-03424" class="html-bibr">43</a>] and plotted in Figure (<b>d</b>) as a dashed curve for comparison of inherent spectral structures.</p>
Full article ">Figure 3
<p>Correlation among confocal microscopy (of PL integrated within 450 to 720 nm spectral range) and optical microscopy images (<b>a</b>,<b>b</b>), as well as atomic force microscopy (AFM) profiles (<b>c</b>,<b>d</b>) and PL spectral variations (<b>e</b>,<b>f</b>). The numbered localized etch-pits (<b>a</b>) and the intersection of etched areas (<b>b</b>) obtained with enhanced spatial resolution are illustrated in top figures (<b>a</b>,<b>b</b>). The related optical images (bottom figures (<b>a</b>,<b>b</b>)) obtained in reflected light on the examined sample areas. The atomic force microscopy profiles scanned close to a single etch-pit (<b>c</b>) highlighted after short etching exposure; and rather extended areas (<b>d</b>) of the intersecting etched surfaces, formed under long etching procedures. The shape of the dislocation-ascribed single etch-pit can be employed for identification of dislocation type (as the edge dislocation ascribed etch pit illustrated in (<b>c</b>)). The deep etch-pits ascribed to either screw- or mixed type dislocations can be assumed by analyzing intersections of wide area etched hexagon valleys (<b>d</b>). The blue PL band spectral components appear only (<b>e</b>,<b>f</b>) when the confocal microscopy probe is localized either close to the core (location 2 in (<b>a</b>,<b>c</b>,<b>e</b>)—assuming dimensions of space charge region <span class="html-italic">R</span><sub>0</sub>) of the single dislocation or steep planes of hexagonal valleys ((<b>b</b>,<b>d</b>,<b>e</b>), respectively). Numbers (1–8) denote locations nearby etch-pits where the PL spectra are recorded.</p>
Full article ">Figure 4
<p>(<b>a</b>) Evolution of the PL spectra with H<sub>3</sub>PO<sub>4</sub> etching exposure. (<b>b</b>) The B-PL band intensity (peaking at 2.7) evolution with etching exposure recorded (symbols) in MOCVD GaN using pulsed UV excitation and fitted (dashed curve) using Equation (3).</p>
Full article ">Figure 5
<p>The horizontal etch rate as a function of H<sub>3</sub>PO<sub>4</sub> acid temperature.</p>
Full article ">
12 pages, 3363 KiB  
Communication
The Saturation of the Response to an Electron Beam of Ce- and Tb-Doped GYAGG Phosphors for Indirect β-Voltaics
by Petr Karpyuk, Mikhail Korzhik, Andei Fedorov, Irina Kamenskikh, Ilya Komendo, Daria Kuznetsova, Elena Leksina, Vialy Mechinsky, Vladimir Pustovarov, Valentina Smyslova, Vasilii M. Retivov, Yauheni Talochka, Dmitry Tavrunov and Andrei Vasil’ev
Appl. Sci. 2023, 13(5), 3323; https://doi.org/10.3390/app13053323 - 6 Mar 2023
Cited by 3 | Viewed by 1675
Abstract
GYAGG:Tb (Ce) scintillators have been confirmed to be promising sources of light emission when excited by an intense 150 keV electron beam. The saturation of the scintillation yield under such excitation conditions has been studied. To explain the results obtained, a model that [...] Read more.
GYAGG:Tb (Ce) scintillators have been confirmed to be promising sources of light emission when excited by an intense 150 keV electron beam. The saturation of the scintillation yield under such excitation conditions has been studied. To explain the results obtained, a model that considers the Auger quenching mechanism was used. The Ce-doped material did not show saturation, whereas a moderate 30% drop of the yield was measured in the Tb-doped sample at the highest excitation beam intensity ~1 A/cm2. This put forward a way to exploit the Tb-doped scintillator for indirect β-voltaic batteries. Full article
(This article belongs to the Topic Innovative Materials for Energy Conversion and Storage)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction pattern of GYAGG:Tb ceramics samples with different terbium concentration.</p>
Full article ">Figure 2
<p>Room temperature PL spectra at excitation in Tb<sup>3+</sup> interconfiguration transition (λex = 273 nm) of GYAGG:Tb samples versus the activator concentration (<b>a</b>); scintillation spectrum of the GYAGG:Tb, Tb = 0.15 at excitation by the electron beam at the maximum intensity (<b>b</b>).</p>
Full article ">Figure 3
<p>Room temperature PL kinetics <sup>5</sup>D<sub>4</sub> → <sup>7</sup>F<sub>J</sub> (λreg = 546 nm) in GYAGG:Tb ceramic samples versus Tb content.</p>
Full article ">Figure 4
<p>Room temperature PL spectra at excitation in Tb<sup>3+</sup> interconfiguration transition (λex = 273 nm) of GYAGG:Tb-2 and GYAGG:Tb-3 samples. Inset includes the Gd/Y ratio, assuming a homogeneous distribution of the Tb ions between positions occupied by Gd and Y ions. Therefore, the Gd and Y indexes are each increased by 0.1.</p>
Full article ">Figure 5
<p>Room temperature PL kinetics <sup>5</sup>D<sub>4</sub> → <sup>7</sup>F<sub>J</sub> (λreg = 546 nm) in GYAGG:Tb-2 и GYAGG:Tb-3, samples.</p>
Full article ">Figure 6
<p>Room temperature scintillation kinetics of GYAGG:Tb, Tb = 0.15, in the spectral range of the transitions: (<b>a</b>) <sup>5</sup>D<sub>3</sub> → <sup>7</sup>F<sub>J</sub> (380 nm); (<b>b</b>) <sup>5</sup>D<sub>4</sub> → <sup>7</sup>F<sub>J</sub>; (546 nm).</p>
Full article ">Figure 7
<p>Normalized integrated intensity of the scintillation in the spectral range of the transitions: (<b>a</b>) <sup>5</sup>D<sub>4</sub> → <sup>7</sup>F<sub>J</sub>; (<b>b</b>) <sup>5</sup>D<sub>3</sub> → <sup>7</sup>F<sub>J</sub> of GYAGG:Tb, Tb = 0.15; (<b>c</b>) GYAGG:Ce.</p>
Full article ">Figure 8
<p>Scintillation yield versus beam power: (<b>a</b>) GYAGG:Tb; (<b>b</b>) GYAGG:Ce.</p>
Full article ">
12 pages, 11515 KiB  
Article
Organic Scintillator-Fibre Sensors for Proton Therapy Dosimetry: SCSF-3HF and EJ-260
by Crystal Penner, Samuel Usherovich, Jana Niedermeier, Camille Belanger-Champagne, Michael Trinczek, Elisabeth Paulssen and Cornelia Hoehr
Electronics 2023, 12(1), 11; https://doi.org/10.3390/electronics12010011 - 20 Dec 2022
Cited by 2 | Viewed by 1770
Abstract
In proton therapy, the dose from secondary neutrons to the patient can contribute to side effects and the creation of secondary cancer. A simple and fast detection system to distinguish between dose from protons and neutrons both in pretreatment verification as well as [...] Read more.
In proton therapy, the dose from secondary neutrons to the patient can contribute to side effects and the creation of secondary cancer. A simple and fast detection system to distinguish between dose from protons and neutrons both in pretreatment verification as well as potentially in vivo monitoring is needed to minimize dose from secondary neutrons. Two 3 mm long, 1 mm diameter organic scintillators were tested for candidacy to be used in a proton–neutron discrimination detector. The SCSF-3HF (1500) scintillating fibre (Kuraray Co. Chiyoda-ku, Tokyo, Japan) and EJ-260 plastic scintillator (Eljen Technology, Sweetwater, TX, USA) were irradiated at the TRIUMF Neutron Facility and the Proton Therapy Research Centre. In the proton beam, we compared the raw Bragg peak and spread-out Bragg peak response to the industry standard Markus chamber detector. Both scintillator sensors exhibited quenching at high LET in the Bragg peak, presenting a peak-to-entrance ratio of 2.59 for the EJ-260 and 2.63 for the SCSF-3HF fibre, compared to 3.70 for the Markus chamber. The SCSF-3HF sensor demonstrated 1.3 times the sensitivity to protons and 3 times the sensitivity to neutrons as compared to the EJ-260 sensor. Combined with our equations relating neutron and proton contributions to dose during proton irradiations, and the application of Birks’ quenching correction, these fibres provide valid candidates for inexpensive and replicable proton-neutron discrimination detectors. Full article
(This article belongs to the Special Issue Applications of Optical Fiber Sensors)
Show Figures

Figure 1

Figure 1
<p>Assembly of an SCSF-3HF fibre-scintillator detector while in the coupling jig.</p>
Full article ">Figure 2
<p>PTRC extracted 74 MeV beamline showing structures traversed by the beam before ending in the water box. The modulator wheel is absent for RBP measurements and present for the SOBP.</p>
Full article ">Figure 3
<p>Dose and dose rate responses: (<bold>a</bold>) dose response comparison between the EJ-260 and SCSF-3HF fibres at three dose levels: 10,000 MC (1.4 Gy), 100,000 MC (14 Gy), and 500,000 MC (70 Gy). Error bars, some too small to be seen against the data point, indicate standard deviation in the mean of three averaged irradiations at each dose; (<bold>b</bold>) dose rate response comparison between EJ-260 and SCSF-3HF at three different beam currents: 2, 6, and 10 nA for an exposure of 50,000 MC (7 Gy) each.</p>
Full article ">Figure 4
<p>RBP comparison between EJ-260, SCSF-3HF, and Markus chamber dosimeters.</p>
Full article ">Figure 5
<p>SOBP comparison between EJ-260, SCSF-3HF, and Markus chamber dosimeters. The Markus chamber and SCSF-3HF data were shifted to align the peak positions with EJ-260 due to different sizes of the detectors. This shift is not an unexpected consequence of different initial measurement positions of the active area in the water phantom.</p>
Full article ">Figure 6
<p>(<bold>a</bold>) Scintillator sensor output vs. LET with Birks’ first-order fit. (<bold>b</bold>) QCF vs. LET with linear fits. Curves are normalized to 1 at <inline-formula><mml:math id="mm14"><mml:semantics><mml:mrow><mml:mn>15</mml:mn><mml:mo> </mml:mo><mml:mi>MeV</mml:mi><mml:mo>/</mml:mo><mml:mi>cm</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 7
<p>Proton (purple) and neutron (orange) sensitivity comparison between the EJ-260 and SCSF-3HF sensors. Error bars indicate standard deviation of the three trials for each experiment.</p>
Full article ">Figure 8
<p>Spectra for the two scintillators normalized with their maxima at 1.</p>
Full article ">
7 pages, 686 KiB  
Article
The Yield of Cherenkov and Scintillation Radiation Generated by the 2.7 MeV Electron Beam in Plate PMMA Samples
by Boris Alekseev, Viktor Tarasenko, Evgeniy Baksht, Alexaner Potylitsyn, Alexander Burachenko, Michail Shevelev, Sergey Uglov and Artem Vukolov
Micro 2022, 2(4), 663-669; https://doi.org/10.3390/micro2040044 - 25 Nov 2022
Cited by 2 | Viewed by 1921
Abstract
In this paper, we have investigated characteristics of ultraviolet and visible radiation generated by the 2.7 MeV electrons. It is shown that the Cherenkov radiation (ChR) intensity predominates over scintillations including wavelength shifting and cathodoluminescence quenching in pure poly(methylmethacrylate) (PMMA) for such electron [...] Read more.
In this paper, we have investigated characteristics of ultraviolet and visible radiation generated by the 2.7 MeV electrons. It is shown that the Cherenkov radiation (ChR) intensity predominates over scintillations including wavelength shifting and cathodoluminescence quenching in pure poly(methylmethacrylate) (PMMA) for such electron energy. To separate ChR and scintillations, we measured emission spectra and orientation dependence of the PMMA samples and compared with GEANT4 model taking into account only ChR mechanism. Full article
(This article belongs to the Section Analysis Methods and Instruments)
Show Figures

Figure 1

Figure 1
<p>Scheme of the microtron experiments for temporal (<bold>a</bold>) and spectral (<bold>b</bold>) studies. 1—collimating diaphragm; 2—PMMA sample; 3—SiPM; 4—spectrometer.</p>
Full article ">Figure 2
<p>Simulation of the electron beam penetration into the 10 mm PMMA sample (<bold>a</bold>) and the relative energy dependence of an electron beam passing through targets various thicknesses <italic>L</italic> (<bold>b</bold>).</p>
Full article ">Figure 3
<p>Transmmittance of PMMA samples and optical emission spectra for 2.7 MeV electron beam penetrating into the 10 mm PMMA sample. The fiber was placed under angle <inline-formula><mml:math id="mm32"><mml:semantics><mml:mrow><mml:mi>ψ</mml:mi><mml:mo>=</mml:mo><mml:msup><mml:mn>47</mml:mn><mml:mo>∘</mml:mo></mml:msup></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 4
<p>Beam current <inline-formula><mml:math id="mm33"><mml:semantics><mml:msub><mml:mi>I</mml:mi><mml:mi>b</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> registered using Faraday cap and the voltage at the silicon SiPM <inline-formula><mml:math id="mm34"><mml:semantics><mml:msub><mml:mi>U</mml:mi><mml:mrow><mml:mi>S</mml:mi><mml:mi>i</mml:mi><mml:mi>P</mml:mi><mml:mi>M</mml:mi></mml:mrow></mml:msub></mml:semantics></mml:math></inline-formula> registering emission in the visible and UV range.</p>
Full article ">Figure 5
<p>(<bold>a</bold>) Visible and UV spectral energy density distribution for 1 mm PMMA sample, which are registered by a spectrometer with the fiber locating under angle <inline-formula><mml:math id="mm35"><mml:semantics><mml:mrow><mml:mi>ψ</mml:mi><mml:mo>=</mml:mo><mml:msup><mml:mn>47</mml:mn><mml:mo>∘</mml:mo></mml:msup></mml:mrow></mml:semantics></mml:math></inline-formula>. (<bold>b</bold>) Orientational dependence of emission energy in the 350–700 nm range from the PMMA sample angle <inline-formula><mml:math id="mm36"><mml:semantics><mml:mi>ψ</mml:mi></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 6
<p>The darkening PMMA sample after electron beam exposition (<bold>a</bold>), photos of PMMA optical radiation for PMMA samples with thicknesses <inline-formula><mml:math id="mm37"><mml:semantics><mml:mrow><mml:mi>L</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> mm (<bold>b</bold>) and <inline-formula><mml:math id="mm38"><mml:semantics><mml:mrow><mml:mi>L</mml:mi><mml:mo>=</mml:mo><mml:mn>10</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> mm (<bold>c</bold>). The dimensions of all photos are <inline-formula><mml:math id="mm39"><mml:semantics><mml:mrow><mml:mn>40</mml:mn><mml:mo>×</mml:mo><mml:mn>40</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> mm.</p>
Full article ">
17 pages, 7965 KiB  
Article
Wide Concentration Range of Tb3+ Doping Influence on Scintillation Properties of (Ce, Tb, Gd)3Ga2Al3O12 Crystals Grown by the Optical Floating Zone Method
by Tong Wu, Ling Wang, Yun Shi, Xintang Huang, Qian Zhang, Yifei Xiong, Hui Wang, Jinghong Fang, Jinqi Ni, Huan He, Chaoyue Wang, Zhenzhen Zhou, Qian Liu, Qin Li, Jianding Yu, Oleg Shichalin and Evgeniy Papynov
Materials 2022, 15(6), 2044; https://doi.org/10.3390/ma15062044 - 10 Mar 2022
Cited by 5 | Viewed by 2320
Abstract
To obtain a deeper understand of the energy transfer mechanism between Ce3+ and Tb3+ ions in the aluminum garnet hosts, (Ce, Tb, Gd)3Ga2Al3O12 (GGAG:Ce, Tb) single crystals grown by the optical floating zone (OFZ) [...] Read more.
To obtain a deeper understand of the energy transfer mechanism between Ce3+ and Tb3+ ions in the aluminum garnet hosts, (Ce, Tb, Gd)3Ga2Al3O12 (GGAG:Ce, Tb) single crystals grown by the optical floating zone (OFZ) method were investigated systematically in a wide range of Tb3+ doping concentration (1–66 at.%). Among those, crystal with 7 at.% Tb reached a single garnet phase while the crystals with other Tb3+ concentrations are mixed phases of garnet and perovskite. Obvious Ce and Ga loss can be observed by an energy dispersive X-ray spectroscope (EDS) technology. The absorption bands belonging to both Ce3+ and Tb3+ ions can be observed in all crystals. Photoluminescence (PL) spectra show the presence of an efficient energy transfer from the Tb3+ to Ce3+ and the gradually quenching effect with increasing of Tb3+ concentration. GGAG: 1% Ce3+, 7% Tb3+ crystal was found to possess the highest PL intensity under excitation of 450 nm. The maximum light yield (LY) reaches 18,941 pho/MeV. The improved luminescent and scintillation characteristics indicate that the cation engineering of Tb3+ can optimize the photoconversion performance of GGAG:Ce. Full article
Show Figures

Figure 1

Figure 1
<p>Photographs of the 1 at.% Ce doped (Tb<sub>x</sub>Gd<sub>1-x</sub>)<sub>3</sub>Ga<sub>2</sub>Al<sub>3</sub>O<sub>12</sub> (GGAG:Ce, xTb) (x = 0–66 at.%) crystals grown by the optical floating zone method.</p>
Full article ">Figure 2
<p>Powder XRD pattern of (<b>a</b>) the powders sintered at 1350 °C for 8 h in air, (<b>b</b>) GGAG:Ce, xTb crystals, (<b>c</b>) expanded view of diffraction peaks between 32° and 38°; (<b>d</b>) calculated percentage of perovskite phase; (<b>e</b>) Laue photograph of GGAG:Ce, 7% Tb<sup>3+</sup> crystal.</p>
Full article ">Figure 2 Cont.
<p>Powder XRD pattern of (<b>a</b>) the powders sintered at 1350 °C for 8 h in air, (<b>b</b>) GGAG:Ce, xTb crystals, (<b>c</b>) expanded view of diffraction peaks between 32° and 38°; (<b>d</b>) calculated percentage of perovskite phase; (<b>e</b>) Laue photograph of GGAG:Ce, 7% Tb<sup>3+</sup> crystal.</p>
Full article ">Figure 2 Cont.
<p>Powder XRD pattern of (<b>a</b>) the powders sintered at 1350 °C for 8 h in air, (<b>b</b>) GGAG:Ce, xTb crystals, (<b>c</b>) expanded view of diffraction peaks between 32° and 38°; (<b>d</b>) calculated percentage of perovskite phase; (<b>e</b>) Laue photograph of GGAG:Ce, 7% Tb<sup>3+</sup> crystal.</p>
Full article ">Figure 3
<p>XPS spectra of the as grown (<b>a</b>) GGAG:Ce, 7% and (<b>b</b>) GGAG:Ce , 50% Tb<sup>3+</sup> crystals using commercial (<b>c</b>) Tb<sub>4</sub>O<sub>7</sub> powder as reference sample.</p>
Full article ">Figure 4
<p>SEM-EDS mapping images for (<b>a</b>) Ce element (<b>b</b>) Tb element of GGAG:Ce, 7% Tb<sup>3+</sup> crystal and (<b>c</b>) Ce element (<b>d</b>) Tb element of GGAG:Ce, 50% Tb<sup>3+</sup> crystal.</p>
Full article ">Figure 5
<p>Absorptance spectra of the as grown GGAG:Ce, xTb (x = 0–66 at.%) crystals (after double face polished to1.0 mm thickness).</p>
Full article ">Figure 6
<p>Photo-luminescence spectra of the as grown GGAG:Ce, xTb (x = 0–66 at.%) crystals; (<b>a</b>) PLE λ<sub>em</sub> = 530 nm; (<b>b</b>) PL λ<sub>ex</sub> = 270 nm; (<b>c</b>) PL λ<sub>ex</sub> = 320 nm; (<b>d</b>) PL λ<sub>ex</sub> = 450 nm; (<b>e</b>) Normalized PL spectra, λ<sub>ex</sub> = 450 nm.</p>
Full article ">Figure 6 Cont.
<p>Photo-luminescence spectra of the as grown GGAG:Ce, xTb (x = 0–66 at.%) crystals; (<b>a</b>) PLE λ<sub>em</sub> = 530 nm; (<b>b</b>) PL λ<sub>ex</sub> = 270 nm; (<b>c</b>) PL λ<sub>ex</sub> = 320 nm; (<b>d</b>) PL λ<sub>ex</sub> = 450 nm; (<b>e</b>) Normalized PL spectra, λ<sub>ex</sub> = 450 nm.</p>
Full article ">Figure 7
<p>Photo-luminescence integral intensity of the as grown GGAG:Ce, xTb (x = 0–66 at.%) crystals, λ<sub>ex</sub> = 450 nm.</p>
Full article ">Figure 8
<p>The PL QY of the as-grown GGAG:Ce, xTb (x = 0–66 at.%) crystals.</p>
Full article ">Figure 9
<p>Radio-luminescence spectra of the as grown GGAG:Ce, xTb (x = 0–66 at.%) crystals, X-ray tube: 70 kV, 1.5 mA.</p>
Full article ">Figure 10
<p>Pulse height spectra of the as grown GGAG:Ce, xTb (x = 0–66 at.%) crystals. The shaping time is 0.75 μs, under <sup>137</sup>Cs (662 keV) gamma ray.</p>
Full article ">Figure 11
<p>Calculated relative light yield of as grown GGAG:Ce, xTb (x = 0–66 at.%) crystals with respect to the reference GGAG:Ce crystal (LY 58,000 pho/MeV).</p>
Full article ">
14 pages, 2853 KiB  
Article
Fabrication of Lead Free Borate Glasses Modified by Bismuth Oxide for Gamma Ray Protection Applications
by Yas Al-Hadeethi, M. I. Sayyed, Abeer Z. Barasheed, Moustafa Ahmed and Mohamed Elsafi
Materials 2022, 15(3), 789; https://doi.org/10.3390/ma15030789 - 21 Jan 2022
Cited by 42 | Viewed by 2620
Abstract
In the present work, bismuth borate glass samples with the composition of (99-x) B2O3 + 1Cr2O3 + (x) Bi2O3 (x = 0, 5, 10, 15, 20, and 25 wt [...] Read more.
In the present work, bismuth borate glass samples with the composition of (99-x) B2O3 + 1Cr2O3 + (x) Bi2O3 (x = 0, 5, 10, 15, 20, and 25 wt %) were prepared using the melt quenching technique. The mass attenuation coefficient (MAC) of the prepared glass samples was measured through a narrow beam technique using a NaI(Tl) scintillation detector. Four point sources were used (241Am, 133Ba, 152Eu, and 137Cs) to measure the MAC for the prepared glasses. The experimental data were compared with the theoretical results obtained from the XCOM, and it was shown that for all samples at all tested energies, the relative deviation between the samples is less than 3%. This finding signifies that the experimental data can adequately be used to evaluate the shielding ability of the glasses. The MAC of the sample with x = 25 wt % was compared with different lead borate glasses and the results indicated that the present sample has high attenuation which is very close to commercial lead borate glasses. We determined the transmission factor (TF), and found that it is small at low energies and increases as the energy increases. The addition of Bi2O3 leads to reduction in the TF values, which improves the shielding performance of the glass system. The half value layer (HVL) of the BCrBi-10 sample was 0.400 cm at 0.595 MeV, 1.619 cm at 0.2447 MeV, and 4.946 cm at 1.4080 MeV. Meanwhile, the HVL of the BCrBi-20 sample is equal to 0.171 and 4.334 cm at 0.0595 and 1.4080 MeV, respectively. The HVL data emphasize that higher energy photons tend to penetrate through the glasses with greater ease than lower energy photons. Furthermore, the fast neutron removable cross section (FNRC) was determined for the present samples and compared with lead borate glass and concrete, and the results showed a remarkable superiority of the bismuth borate glass samples. Full article
(This article belongs to the Special Issue Radiation Shielding Materials)
Show Figures

Figure 1

Figure 1
<p>The steps of sample preparation in the present study.</p>
Full article ">Figure 2
<p>The energy dispersive X-ray analysis (EDX) for BCrBi-15 glass sample.</p>
Full article ">Figure 3
<p>The schematic diagram of the experimental setup of the narrow beam method.</p>
Full article ">Figure 4
<p>The incident and transmitted spectra for Cs-137 at different thicknesses of the BCrBi-20 glass sample.</p>
Full article ">Figure 5
<p>The MAC as a function of energy for the BCrBi glass samples.</p>
Full article ">Figure 6
<p>The transmission factor for the BCrBi glass samples.</p>
Full article ">Figure 7
<p>The radiation protection efficiency as a function of glass thickness at 0.0596 MeV.</p>
Full article ">Figure 8
<p>The radiation protection efficiency as a function of glass thickness at 0.6617 MeV.</p>
Full article ">Figure 9
<p>The half value layer of the BCrBiX glasses.</p>
Full article ">Figure 10
<p>The effective atomic number (<span class="html-italic">Z<sub>eff</sub></span>) of the BCrBiX glasses.</p>
Full article ">Figure 11
<p>The MAC of BCrBi-25 glass sample compared with other commercial glasses.</p>
Full article ">Figure 12
<p>The effective removal for fast neutrons of the present glass samples compared with lead borate glass and concrete.</p>
Full article ">
32 pages, 1234 KiB  
Review
Novel Cs2HfCl6 Crystal Scintillator: Recent Progress and Perspectives
by Serge Nagorny
Physics 2021, 3(2), 320-351; https://doi.org/10.3390/physics3020023 - 13 May 2021
Cited by 24 | Viewed by 4115
Abstract
Recent progress in Cs2HfCl6 (CHC) crystal production achieved within the last five years is presented. Various aspects have been analyzed, including the chemical purity of raw materials, purification methods, optimization of the growth and thermal conditions, crystal characterization, defect structure, [...] Read more.
Recent progress in Cs2HfCl6 (CHC) crystal production achieved within the last five years is presented. Various aspects have been analyzed, including the chemical purity of raw materials, purification methods, optimization of the growth and thermal conditions, crystal characterization, defect structure, and internal radioactive background. Large volume, crack-free, and high quality CHC crystals with an ultimate scintillating performance were produced as a result of such extensive research and development (R & D) program. For example, the CHC crystal sample with dimensions ∅23 × 30 mm3 demonstrates energy resolution of 3.2% FWHM at 662 keV, the relative light output at the level of 30,000 ph/MeV and excellent linearity down to 20 keV. Additionally, this material exhibits excellent pulse shape discrimination ability and low internal background of less than 1 Bq/kg. Furthermore, attempts to produce a high quality CHC crystal resulted in research on this material optimization by constitution of either alkali ions (Cs to Tl), or main element (Hf to Zr), or halogen ions (Cl to Br, I, or their mixture in different ratio), as well as doping with various active ions (Te4+, Ce3+, Eu3+, etc.). This leads to a range of new established scintillating materials, such as Tl2HfCl6, Tl2ZrCl6, Cs2HfCl4Br2, Cs2HfCl3Br3, Cs2ZrCl6, and Cs2HfI6. To exploit the whole potential of these compounds, detailed studies of the material’s fundamental properties, and understanding of the variety of the luminescence mechanisms are required. This will help to understand the origin of the high light yield and possible paths to further extend it. Perspectives of CHC crystals and related materials as detectors for rare nuclear processes are also discussed. Full article
(This article belongs to the Special Issue Radiation Spectroscopy with Solid Scintillators for Rare Events)
Show Figures

Figure 1

Figure 1
<p>Pulse-height spectrum collected with a ∅23 × 30 mm<sup>3</sup> CHC crystal sample resulting in the energy resolution of 3.5% (FWHM) at 662 keV of <sup>137</sup>Cs gamma source (see [<a href="#B13-physics-03-00023" class="html-bibr">13</a>] for more detail).</p>
Full article ">Figure 2
<p>Mean time versus Energy scatter plot for the data collected in low-background measurements with the 7 g CHC scintillating crystal over 2848 h [<a href="#B7-physics-03-00023" class="html-bibr">7</a>]. The mean time values corresponding to for alphas or betas/gammas events, along with intervals containing 99% events of each type are shown.</p>
Full article ">Figure 3
<p>Dependence of the quenching factor (QF) on the energy of alpha particles measured by internal alpha decays of <sup>224</sup>Ra, <sup>220</sup>Rn, <sup>216</sup>Po nuclides in the 7 g CHC crystal (blue squares) [<a href="#B7-physics-03-00023" class="html-bibr">7</a>]. The QF model obtained as a global fit of these data points following the phenomenological model of [<a href="#B37-physics-03-00023" class="html-bibr">37</a>] is also presented (red dotted line). The QF measured with external collimated <sup>241</sup>Am alpha source from [<a href="#B15-physics-03-00023" class="html-bibr">15</a>] is also shown (cyan circle).</p>
Full article ">Figure 4
<p>Left: Schematic cross-section view of the low-background experimental set-up with HPGe detector (not in scale) used for measurements of the radioactive contamination of raw materials, materials after purification and the final CHC crystals. Right: Adaptation of the same experimental set-up with HPGe detector for <sup>174</sup>Hf rare alpha decay measurements in a coincidence with scintillation channel (PMT). See more description in the text (<a href="#sec9-physics-03-00023" class="html-sec">Section 9</a> and <a href="#sec10-physics-03-00023" class="html-sec">Section 10</a>).</p>
Full article ">Figure 5
<p>Energy spectrum of the “pure” alpha selected by pulse shape analysis events in the energy range from 1.5 to 8.0 MeV from the data collected in the low-background measurements with the 7 g CHC scintillating crystal over 2848 h [<a href="#B7-physics-03-00023" class="html-bibr">7</a>]. The energy scale is calibrated taking into account quenching factor values for internal alpha particles. Above 4 MeV, the global fit of the data by the background model built based on alpha decays of radionuclides from <sup>232</sup>Th and <sup>238</sup>U decay chains is shown by blue solid line (individual components of the background model are also shown). Below 4 MeV, the fit of the data with model built from <sup>174</sup>Hf alpha decay (red line) and <sup>147</sup>Sm alpha decay plus degraded alphas and an exponential function (to describe residual of beta/gamma events) is shown (green dashed line). The yellow area represents the background model with respect to the signal from <sup>174</sup>Hf alpha decay.</p>
Full article ">
23 pages, 2553 KiB  
Article
Luminescence Response and Quenching Models for Heavy Ions of 0.5 keV to 1 GeV/n in Liquid Argon and Xenon
by Akira Hitachi
Instruments 2021, 5(1), 5; https://doi.org/10.3390/instruments5010005 - 11 Jan 2021
Cited by 2 | Viewed by 3246
Abstract
Biexcitonic collision kinetics with prescribed diffusion in the ion track core have been applied for scintillation response due to heavy ions in liquid argon. The quenching factors q = EL/E, where E is the ion energy and EL [...] Read more.
Biexcitonic collision kinetics with prescribed diffusion in the ion track core have been applied for scintillation response due to heavy ions in liquid argon. The quenching factors q = EL/E, where E is the ion energy and EL is the energy expended for luminescence, for 33.5 MeV/n 18O and 31.9 MeV/n 36Ar ions in liquid Ar at zero field are found to be 0.73 and 0.46, compared with measured values of 0.59 and 0.46, respectively. The quenching model is also applied for 80–200 keV Pb recoils in α-decay, background candidates in direct dark matter searches, in liquid argon. Values obtained are ~0.09. A particular feature of Birks’ law has been found and exploited in evaluating the electronic quenching factor qel in liquid Xe. The total quenching factors qT for 0.5–20 keV Xe recoils needed for weakly interacting massive particle (WIMP) searches are estimated to be ~0.12–0.14, and those for Pb recoils of 103 and 169 keV are 0.08 and 0.09, respectively. In the calculation, the nuclear quenching factor qnc = Eη/E, where Eη is the energy available for the electronic excitation, is obtained by Lindhard theory and a semi-empirical theory by Ling and Knipp. The electronic linear energy transfer plays a key role. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Quenching factor <span class="html-italic">q</span> for various ions as a function of linear energy transfer (LET) in liquid Ar. Measurements are as follows: 5.305 MeV <sup>210</sup>Po α-particles (▪), relativistic heavy ions (RHIs) of ~1 GeV/n (●) [<a href="#B12-instruments-05-00005" class="html-bibr">12</a>], 33.5 MeV/n <sup>18</sup>O and 31.9 MeV/n <sup>36</sup>Ar ions (closed diamonds) [<a href="#B14-instruments-05-00005" class="html-bibr">14</a>] and <sup>252</sup>Cf fission fragments (closed diamonds) [<a href="#B12-instruments-05-00005" class="html-bibr">12</a>]. Heavy fragments (HFs) and light fragments (LFs) are shown separately. Crosses (×,+) show the results calculated numerically for the hard-sphere cross-section divided by 4, σ<sub>HS</sub>/4. The previous result obtained for LFs with <span class="html-italic">T</span><sub>c</sub>/<span class="html-italic">T</span> = 1 is also shown (-). Dashed curves show Birks’ law applied to the numerical calculation for core quenching (Equation (27)) as shown in Figure 3b. RHIs and nonrelativistic ions (<sup>18</sup>O, <sup>36</sup>Ar and FFs) are treated separately. Open diamonds show <span class="html-italic">q</span> obtained by the sum of scintillation and charge signals [<a href="#B14-instruments-05-00005" class="html-bibr">14</a>] (see <a href="#sec4dot3-instruments-05-00005" class="html-sec">Section 4.3</a>). (<b>b</b>) Initial radial distribution of excited species, essentially the same as the prompt dose profile, in the cylindrical track core produced by various ions in liquid Ar (see <a href="#sec2dot1-instruments-05-00005" class="html-sec">Section 2.1</a>). The distribution for 30 keV Ar recoils (RC Ar) is also shown (dot-dashed curve). A Gaussian distribution is assumed.</p>
Full article ">Figure 2
<p>The stopping powers and the electronic LET for Pb ions: (<b>a</b>) in argon; (<b>b</b>) in xenon. <span class="html-italic">LET</span><sub>el</sub> was obtained by using a semi-empirical power-law approximation for <span class="html-italic">q</span><sub>nc</sub>.</p>
Full article ">Figure 3
<p>Calculated quenching in liquid argon. (<b>a</b>) Evolution of quenching in the track core produced by α-particles, Au, <sup>18</sup>O, <sup>36</sup>Ar and Ba ions (as heavy fission fragments). The ratio of the number of self-trapped excitons with and without quenching as a function of time is plotted. The initial radii are <span class="html-italic">a</span><sub>0</sub> = 0.39 and 5.8 nm for α-particles (dotted curve) and Au ions (dashed curve), respectively. <span class="html-italic">a</span><sub>0</sub> values for <sup>18</sup>O, <sup>36</sup>Ar and Ba ions (solid curves) are ~1.5 nm. (<b>b</b>) The inverse of <span class="html-italic">q</span><sub>c</sub> calculated for <sup>18</sup>O, <sup>36</sup>Ar, Mo (LF) and Ba (HF) ions plotted as a function of core LET. Birks’ law was used to fit values for <sup>18</sup>O and <sup>36</sup>Ar ions. Birks’ law applied for RHIs is also shown (<span class="html-italic">a</span><sub>0</sub> = 5.4–6.2 nm).</p>
Full article ">Figure 4
<p>Quenching factors <span class="html-italic">q<sub>T</sub></span> and <span class="html-italic">q</span><sub>nc</sub> for heavy recoils in α-decay. (<b>a</b>) Quenching factors in liquid Ar. Calculations are for <sup>208</sup>Pb ions. Symbols are <span class="html-italic">W</span><sub>α</sub>/<span class="html-italic">W</span><sub>NR</sub> measurements in gas for <sup>206</sup>Pb, <sup>208</sup>Tl and <sup>208</sup>Pb ions by Madsen (•) [<a href="#B27-instruments-05-00005" class="html-bibr">27</a>]. Points for 103 keV <sup>206</sup>Pb recoil ions by Cano (closed diamond) [<a href="#B29-instruments-05-00005" class="html-bibr">29</a>] and Jesse and Sadauskis (triangle) [<a href="#B28-instruments-05-00005" class="html-bibr">28</a>] are shifted for clarity. Numerical results (dashed curve) and the power-law approximation (dot-dashed curve) for <span class="html-italic">q</span><sub>nc</sub> by Lindhard et al. are shown [<a href="#B23-instruments-05-00005" class="html-bibr">23</a>]. The dotted curve shows <span class="html-italic">q</span><sub>nc</sub> estimated by Ling and Knipp [<a href="#B30-instruments-05-00005" class="html-bibr">30</a>]. Measured <span class="html-italic">q</span><sub>T</sub> values by Xu et al. [<a href="#B35-instruments-05-00005" class="html-bibr">35</a>] are shown with open diamonds (corrected for <span class="html-italic">L</span><sub>γ</sub><sub>0</sub>) together with present calculation (solid curve). (<b>b</b>) Quenching factors in liquid Xe. <span class="html-italic">W</span><sub>α</sub>/<span class="html-italic">W</span><sub>NR</sub> measurements for <sup>206</sup>Pb in gas by Cano (closed diamond) [<a href="#B29-instruments-05-00005" class="html-bibr">29</a>]. Calculated values for <span class="html-italic">q</span><sub>nc</sub> are the power-law approximation (dot-dashed curve) by Lindhard et al. and the present approximation (dotted curve). The present calculation for <span class="html-italic">q</span><sub>T</sub> is shown by the solid line.</p>
Full article ">Figure 5
<p>Quenching factors in liquid Xe: (<b>a</b>) The electronic quenching factor <span class="html-italic">q</span><sub>el</sub>, estimated for Xe and Pb recoils as a function of <span class="html-italic">LET</span><sub>el</sub>. A closed circle shows <span class="html-italic">q</span><sub>c</sub> = 0.70 for 5.49 MeV α-particle. Curves show Birks’ law. Horizontal scales show the energy of Xe and Pb recoils in keV; the diamond show a value corrected for core expansion: (<b>b</b>) Quenching factors calculated for Xe recoils in liquid Xe as a function of the energy. The solid curve shows Lindhard <span class="html-italic">q</span><sub>nc</sub>, the dotted curve is an extrapolation. The dot-dashed curve shows the total quenching factor <span class="html-italic">q<sub>T</sub></span> = <span class="html-italic">q</span><sub>nc</sub>·<span class="html-italic">q</span><sub>el</sub>. The dashed curves show uncertainties due to the estimated <span class="html-italic">T<sub>c</sub></span>/<span class="html-italic">T</span> ratio, 0.76±0.04, in α track. The <span class="html-italic">q<sub>T</sub></span> curve may approach to Lindhard <span class="html-italic">q</span><sub>nc</sub> curve faster than the dot-dashed curve as the energy decreases below about 2 keV (see <a href="#sec4dot1-instruments-05-00005" class="html-sec">Section 4.1</a> and <a href="#sec4dot2-instruments-05-00005" class="html-sec">Section 4.2</a>). The result obtained previously by the α-core approximation is shown with a dot-dot dashed curve [<a href="#B25-instruments-05-00005" class="html-bibr">25</a>].</p>
Full article ">Figure A1
<p>Schematic drawing of quenching. Level 1 is the excited level involved in quenching, level 2 gives fluorescence and level 3 is the quencher. (<b>a</b>) Pseudo-first-order reaction. (<b>b</b>) Second-order reaction under diffusion.</p>
Full article ">
11 pages, 3556 KiB  
Article
Synthesis of Green-Emitting Gd2O2S:Pr3+ Phosphor Nanoparticles and Fabrication of Translucent Gd2O2S:Pr3+ Scintillation Ceramics
by Zhigang Sun, Bin Lu, Guiping Ren and Hongbing Chen
Nanomaterials 2020, 10(9), 1639; https://doi.org/10.3390/nano10091639 - 20 Aug 2020
Cited by 30 | Viewed by 3586
Abstract
A translucent Gd2O2S:Pr ceramic scintillator with an in-line transmittance of ~31% at 512 nm was successfully fabricated by argon-controlled sintering. The starting precipitation precursor was obtained by a chemical precipitation route at 80 °C using ammonia solution as the [...] Read more.
A translucent Gd2O2S:Pr ceramic scintillator with an in-line transmittance of ~31% at 512 nm was successfully fabricated by argon-controlled sintering. The starting precipitation precursor was obtained by a chemical precipitation route at 80 °C using ammonia solution as the precipitate, followed by reduction at 1000 °C under flowing hydrogen to produce a sphere-like Gd2O2S:Pr powder with an average particle size of ~95 nm. The Gd2O2S:Pr phosphor particle exhibits the characteristic green emission from 3P0,13H4 transitions of Pr3+ at 512 nm upon UV excitation into a broad excitation band at 285–335 nm arising from 4f2→4f5d transition of Pr3+. Increasing Pr3+ concentrations induce two redshifts for the two band centers of 4f2→4f5d transition and lattice absorption on photoluminescence excitation spectra. The optimum concentration of Pr3+ is 0.5 at.%, and the luminescence quenching type is dominated by exchange interaction. The X-ray excited luminescence spectrum of the Gd2O2S:Pr ceramic is similar to the photoluminescence behavior of its particle. The phosphor powder and the ceramic scintillator have similar lifetimes of 2.93–2.99 μs, while the bulk material has rather higher external quantum efficiency (~37.8%) than the powder form (~27.2%). Full article
(This article belongs to the Special Issue Advanced Nanomaterials for Radiation Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of the Gd<sub>2</sub>O<sub>2</sub>S (GOS):Pr precursor powder and reduction products (<b>A</b>), and schematic crystal structure of GOS (<b>B</b>).</p>
Full article ">Figure 2
<p>FE-SEM micrographs showing morphologies of the precipitation precursor (<b>A</b>) and its reduction product at 1000 °C (<b>B</b>).</p>
Full article ">Figure 3
<p>Appearance (<b>A</b>) and in-line transmittance (<b>B</b>) of the GOS:Pr ceramic scintillator, and SEM micrographs of the surface (<b>C</b>) and fracture surface (<b>D</b>) of the sintered ceramic. The lower part in panel (<b>A</b>) is the appearance of GOS:Pr ceramic under irradiation from a 254 nm UV lamp. The inset in panel (<b>B</b>) is the enlargement of its in-line transmittance curve from 200 to 700 nm.</p>
Full article ">Figure 4
<p>Photoluminescence excitation (PLE) (<b>A</b>) and photoluminescence (PL) (<b>B</b>) spectra of (Gd<sub>1−<span class="html-italic">x</span></sub>Pr<span class="html-italic"><sub>x</sub></span>)<sub>2</sub>O<sub>2</sub>S (<span class="html-italic">x</span> = 0.001–0.0075) phosphor powders, PLE intensities of the 306 nm excitations normalized to 1 for the lowest value (<b>C</b>), and PL intensities of the 512 nm emissions normalized to 1 for the lowest value (<b>D</b>).</p>
Full article ">Figure 5
<p>The relationship between <span class="html-italic">log</span>(<span class="html-italic">I</span>/<span class="html-italic">c</span>) and <span class="html-italic">log</span>(<span class="html-italic">c</span>) for the GOS:Pr phosphors.</p>
Full article ">Figure 6
<p>X-ray excited luminescence (XEL) spectrum of the translucent GOS:Pr ceramic scintillator (<b>A</b>) and its 1931 CIE chromaticity diagram (<b>B</b>).</p>
Full article ">Figure 7
<p>Fluorescence decay behaviors of the GOS:Pr phosphor powder and scintillation ceramic for the 512 nm Pr<sup>3+</sup> emission under 306 nm excitation (<b>A</b>) and their quantum efficiency spectra obtained under 306 nm excitation (<b>B</b>).</p>
Full article ">
Back to TopTop