Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,557)

Search Parameters:
Keywords = luminescent properties

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
10 pages, 2660 KiB  
Article
Crystal Growth and Energy Transfer Study in Ce3+ and Pr3+ Co-Doped Lu2Si2O7
by Yuka Abe, Takahiko Horiai, Yuui Yokota, Masao Yoshino, Rikito Murakami, Takashi Hanada, Akihiro Yamaji, Hiroki Sato, Yuji Ohashi, Shunsuke Kurosawa, Kei Kamada and Akira Yoshikawa
Crystals 2025, 15(3), 202; https://doi.org/10.3390/cryst15030202 - 20 Feb 2025
Abstract
Ce-doped Lu2Si2O7 has a high density, high luminescence efficiency even at high temperatures, and a high effective atomic number, making it a promising candidate for use as a radiation detector in medical devices and resource exploration equipment. In [...] Read more.
Ce-doped Lu2Si2O7 has a high density, high luminescence efficiency even at high temperatures, and a high effective atomic number, making it a promising candidate for use as a radiation detector in medical devices and resource exploration equipment. In this study, we grow and characterize Pr3+ and Ce3+-doped Lu2Si2O7 single crystals by systematically varying the Ce3+ to Pr3+ ratio to further improve scintillation properties. The optical characterization results show a bidirectional energy transfer: from the Pr3+ 5d levels to the Ce3+ 5d levels and from the Ce3+ 5d levels to the Pr3+ 4f levels. Consistently with this result, the PL decay time of emission from the Pr3+ 5d–4f transition tends to become faster as the Ce3+/Pr3+ ratio increases, due to the energy transfer from the Pr3+ 5d levels to the Ce3+ 5d levels. Additionally, (Ce0.0022 Pr0.0016 Lu0.9962)2Si2O7 exhibits a high light yield comparable to Ce-doped Lu2Si2O7 and a slightly faster decay time than Ce-doped Lu2Si2O7. Full article
(This article belongs to the Special Issue Growth and Properties of Novel Scintillator Crystals)
Show Figures

Figure 1

Figure 1
<p>Appearance of the (Ce<sub>x</sub> Pr<sub>y</sub> Lu<sub>1−x−y</sub>)<sub>2</sub>Si<sub>2</sub>O<sub>7</sub> as grown.</p>
Full article ">Figure 2
<p>The results of the XRD analysis for (Ce<sub>x</sub> Pr<sub>y</sub> Lu<sub>1−x−y</sub>)<sub>2</sub>Si<sub>2</sub>O<sub>7</sub> crystals.</p>
Full article ">Figure 3
<p>(<b>a</b>) Absorption spectra, (<b>b</b>,<b>c</b>) PL emission spectra (λ<sub>ex</sub> = 349 nm, 247 nm), (<b>d</b>) PL excitation spectra (λ<sub>em</sub> = 383 nm) of (Ce<sub>x</sub> Pr<sub>y</sub> Lu<sub>1−x−y</sub>)<sub>2</sub>Si<sub>2</sub>O<sub>7</sub> crystals.</p>
Full article ">Figure 4
<p>PL decay curves of (Ce<sub>x</sub> Pr<sub>y</sub> Lu<sub>1−x−y</sub>)<sub>2</sub>Si<sub>2</sub>O<sub>7</sub> crystals: (<b>a</b>) λ<sub>ex</sub> = 349 nm, λ<sub>em</sub> = 383 nm, (<b>b</b>) λ<sub>ex</sub> = 247 nm, λ<sub>em</sub> = 303 nm.</p>
Full article ">Figure 5
<p>Temperature dependence of the PL decay times (λ<sub>ex</sub> = 349 nm, λ<sub>em</sub> = 383 nm and λ<sub>ex</sub> = 247 nm, λ<sub>em</sub> = 303 nm).</p>
Full article ">Figure 6
<p>Pulse–height spectra of (Ce<sub>x</sub> Pr<sub>y</sub> Lu<sub>1−x−y</sub>)<sub>2</sub>Si<sub>2</sub>O<sub>7</sub> samples.</p>
Full article ">Figure 7
<p>Decay curve of (Ce<sub>x</sub> Pr<sub>y</sub> Lu<sub>1−x−y</sub>)<sub>2</sub>Si<sub>2</sub>O<sub>7</sub> samples under excitation with gamma rays from a <sup>137</sup>Cs source.</p>
Full article ">
19 pages, 3142 KiB  
Article
Antitumor Efficacy of Interleukin 12-Transfected Mesenchymal Stem Cells in B16-F10 Mouse Melanoma Tumor Model
by Urška Kamenšek, Tim Božič, Maja Čemažar and Urban Švajger
Pharmaceutics 2025, 17(3), 278; https://doi.org/10.3390/pharmaceutics17030278 - 20 Feb 2025
Abstract
Background/Objectives: Mesenchymal stromal cells (MSCs) hold the potential for tumor-targeted gene delivery due to their ex vivo manipulability, low immunogenicity, scalability, and inherent tumor-homing properties. Despite the widespread use of viral vectors for MSC genetic modification, safety concerns have prompted interest in [...] Read more.
Background/Objectives: Mesenchymal stromal cells (MSCs) hold the potential for tumor-targeted gene delivery due to their ex vivo manipulability, low immunogenicity, scalability, and inherent tumor-homing properties. Despite the widespread use of viral vectors for MSC genetic modification, safety concerns have prompted interest in non-viral alternatives, such as gene electrotransfer (GET). This study aimed to optimize GET parameters for MSCs transfection, assess MSCs biodistribution after in vivo administration, and evaluate the therapeutic potential of interleukin-12 (IL-12)-modified MSCs in a mouse melanoma model. Methods: Human MSCs were isolated from umbilical cords under ethically approved protocols. GET protocols were optimized using a fluorescent reporter gene to evaluate transfection efficiency and cell viability. MSC biodistribution was examined following intravenous and intratumoral injections in murine tumor models using luminescent reporter gene. The therapeutic efficacy of IL-12-modified MSCs was assessed in a syngeneic mouse melanoma model. Results: Optimized GET protocols achieved a transfection efficiency of 80% and a cell viability of 90%. Biodistribution studies demonstrated effective tumor retention of MSCs following intratumoral injections, whereas intravenous administration resulted in predominant cell localization in the lungs. IL-12-modified MSCs injected intratumorally significantly inhibited tumor growth, delaying tumor progression by five days compared to controls. Conclusions: Optimized GET conditions enabled high-efficiency, high-viability MSCs transfection, facilitating their use as effective vehicles for localized cytokine delivery. While the innate tumor tropism of MSCs was not conclusively demonstrated, the study highlights the potential of GET as a reliable non-viral gene delivery platform and underscores the therapeutic promise of IL-12-modified MSCs in tumor-targeted gene therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Immunophenotype and morphology of umbilical cord-derived MSCs. (<b>a</b>) Flow cytometric analysis data of three independent experiments (three different umbilical cord donors), represented as mean ± SD of percentage of positive cells. (<b>b</b>) Morphology of MSCs toward the end of passage 1 of cell culture, displaying typical fibroblast-like morphology. Magnification: ×100.</p>
Full article ">Figure 2
<p>Validation of MSC electroporation protocol. (<b>a</b>) Initial testing using four different GET pulses. (<b>b</b>) Modified protocol under room temperature conditions using selected GET pulses. (<b>c</b>) Fluorescence microscopy image of transfected cells. GET 1 = 8 × 1300 V/cm (260 V/2 mm), 100 µs, 1 Hz; GET 2 = 8 × 1300 V/cm (260 V/2 mm), 100 µs, 4 Hz; GET 3 = 8 × 1300 V/cm (260 V/2 mm), 100 µs, 5000 Hz; GET 4 = 8 × 600 V/cm (120 V/2 mm), 5000 µs, 1 Hz; LF = Lipofection (Lipofectamine 2000); GET, gene electrotransfer; GFP, green fluorescent protein. Results are presented as means of three independent experiments with three replicates ± SEM. Statistical significance: * <span class="html-italic">p</span> ≤ 0.05 vs. lipofection group. Scale bar: 100 µm.</p>
Full article ">Figure 3
<p>In vivo tracking of MSCs. Distribution of Luc-transfected MSCs after intravenous injection in (<b>a</b>) immunologically cold tumor model B16F10, (<b>b</b>) induced hot tumor model, i.e., irradiated B16-F10 tumors) and (<b>c</b>) immunologically hot tumor models CT26 and MC-38. (<b>d</b>) Retention of Luc-transfected MSCs after intratumoral injection in B16F10 tumor model. Location of tumors is marked with circles. In the dual tumor model (<b>b</b>), irradiated tumors are marked with red circles and unirradiated tumors with blue. Images of all included mice are presented; 3–4 mice were included for the intravenous injections, and 10 mice in two independent experiments for intratumoral injections. Luminescent signal is presented as radiance (photons/s/cm<sup>2</sup>/sr) in a logarithmic rainbow scale. To facilitate better comparison, the same radiance range was applied to all images.</p>
Full article ">Figure 4
<p>Therapeutic efficacy of IL-12-transfected MSCs: (<b>a</b>) Tumor growth curves for individual mice. (<b>b</b>) Average tumor growth curves drawn as mean tumor volume with standard error of mean (<b>c</b>) Tumor growth delay (difference between tumor doubling times in control and therapeutic groups) drawn as violin plot with mean. (<b>d</b>) Average mice body weight drawn as mean body weight with standard error of the mean. CTRL = untreated tumor-bearing mice, Ctrl MSC = mice treated with intratumoral injection of non-modified MSCs; IL-12 MSC = mice treated with intratumoral injection of IL-12-transfected MSCs: Luc MSC = mice treated with intratumoral injection of Luc-transfected MSCs. Two independent experiments with 5 animals per group. Statistical significance: * <span class="html-italic">p</span> ≤ 0.05 vs. all groups.</p>
Full article ">Scheme 1
<p>Study design for determination of (<b>a</b>) in vitro transfection efficiency (scale bar: 100 µm), (<b>b</b>) in vivo distribution, and (<b>c</b>) in vivo therapeutic efficiency.</p>
Full article ">
14 pages, 4858 KiB  
Article
Synthesis and Characterization of Smartphone-Readable Luminescent Lanthanum Borates Doped and Co-Doped with Eu and Dy
by Katya Hristova, Irena P. Kostova, Tinko A. Eftimov, Georgi Patronov and Slava Tsoneva
Photonics 2025, 12(2), 171; https://doi.org/10.3390/photonics12020171 - 19 Feb 2025
Abstract
Despite notable advancements in the development of borate materials, improving their luminescent efficiency remains an important focus in materials research. The synthesis of lanthanum borates (LaBO3), doped and co-doped with europium (Eu3⁺) and dysprosium (Dy3⁺), by the [...] Read more.
Despite notable advancements in the development of borate materials, improving their luminescent efficiency remains an important focus in materials research. The synthesis of lanthanum borates (LaBO3), doped and co-doped with europium (Eu3⁺) and dysprosium (Dy3⁺), by the solid-state method, has demonstrated significant potential to address this challenge due to their unique optical properties. These materials facilitate efficient energy transfer from UV-excited host crystals to trivalent rare-earth activators, resulting in stable and high-intensity luminescence. To better understand their structural and vibrational characteristics, Fourier transform infrared (FTIR) spectroscopy and Raman spectroscopy were employed to identify functional groups and molecular vibrations in the synthesized materials. Additionally, X-ray diffraction (XRD) analysis was conducted to determine the crystalline structure and phase composition of the samples. All observed transitions of Eu3⁺ and Dy3⁺ in the excitation and emission spectra were systematically analyzed and identified, providing a comprehensive understanding of their behavior. Although smartphone cameras exhibit non-uniform spectral responses, their integration into this study highlights distinct advantages, including contactless interrogation, effective UV excitation suppression, and real-time spectral analysis. These capabilities enable practical and portable fluorescence sensing solutions for applications in healthcare, environmental monitoring, and food safety. By combining advanced photonic materials with accessible smartphone technology, this work demonstrates a novel approach for developing low-cost, scalable, and innovative sensing platforms that address modern technological demands. Full article
(This article belongs to the Section Optoelectronics and Optical Materials)
Show Figures

Figure 1

Figure 1
<p>Experimental set-up: left is the basic arrangements for the measurement of the 3D excitation–emission spectra of the samples using a standard optical fiber spectrometer; right is a side view of the arrangement from a smartphone camera equipped with a sheet transmission diffraction grating (1000 L/mm) to observe the spectrum.</p>
Full article ">Figure 2
<p>XRD patterns of LaBO<sub>3</sub> doped with Eu and Dy and software-generated crystal structure model of LaBO<sub>3</sub>.</p>
Full article ">Figure 3
<p>FTIR spectra of samples: LaBO<sub>3</sub>:Eu<sup>3+</sup> (S1), LaBO<sub>3</sub>:Dy<sup>3+</sup> (S2), and LaBO<sub>3</sub>:Eu<sup>3+</sup>:Dy<sup>3+</sup> (S3).</p>
Full article ">Figure 4
<p>Raman analysis of sample S3.</p>
Full article ">Figure 5
<p>Raman analysis of samples S1–S3.</p>
Full article ">Figure 6
<p>Graphic representation of the emission intensity at 591 nm, 615 nm, 683 nm, and 708 nm vs. excitation wavelength of 290 nm, 350, 360, 390, 420, 430, 450, and 465 nm for lanthanum borate doped with Eu.</p>
Full article ">Figure 7
<p>Emission spectra of co-doped LaBO<sub>3</sub> at λ<sub>exc</sub> = 290 nm and 350 nm.</p>
Full article ">Figure 8
<p>Topographic and 3D representation of excitation at 290 nm and 396 nm and emission at 589 nm and 615 nm for sample LaBO<sub>3</sub>:Eu<sup>3+</sup>:Dy<sup>3+</sup>, measured by an Ocean Optics spectrometer.</p>
Full article ">Figure 9
<p>Topographic and 3D representation of excitation at 290 nm and 396 nm and emission at 589 nm and 615 nm for sample LaBO<sub>3</sub>:Eu<sup>3+</sup>:Dy<sup>3+</sup>, measured by mobile phone.</p>
Full article ">Figure 10
<p>(<b>a</b>) Comparative graph of samples S1 and S3 for ~615 nm emission at 396 nm excitation, and for sample S2 for 569 nm emission at excitation at 391 nm; (<b>b</b>) comparative graph of samples S1 and S3 at ~591 nm emission at 290 nm and 396 nm excitation, and S2 at 569 nm emission at 393 nm excitation.</p>
Full article ">Scheme 1
<p>Representation of the synthesis of luminescent materials.</p>
Full article ">
13 pages, 2757 KiB  
Article
Crystal Phase and Morphology Control for Enhanced Luminescence in K3GaF6:Er3+
by Yilin Guo, Xin Pan, Yidi Zhang, Ke Su, Rong-Jun Xie, Jiayan Liao, Lefu Mei and Libing Liao
Nanomaterials 2025, 15(4), 318; https://doi.org/10.3390/nano15040318 - 19 Feb 2025
Abstract
Upconversion luminescent materials (UCLMs) have garnered significant attention due to their broad potential applications in fields such as display technology, biological imaging, and optical sensing. However, optimizing crystal phase and morphology remains a challenge. This study systematically investigates the effects of phase transformation [...] Read more.
Upconversion luminescent materials (UCLMs) have garnered significant attention due to their broad potential applications in fields such as display technology, biological imaging, and optical sensing. However, optimizing crystal phase and morphology remains a challenge. This study systematically investigates the effects of phase transformation and morphology control on the upconversion luminescent properties of K3GaF6:Er3+. By comparing different synthesis methods, we found that the hydrothermal method effectively facilitated the transformation of the NaxK3-xGaF6 crystal phase from cubic to monoclinic, with Na+/K+ ions playing a key role in the preparation process. Furthermore, the hydrothermal method significantly optimized the particle morphology, resulting in the formation of uniform octahedral structures. The 657 nm red emission intensity of the monoclinic phase sample doped with Er3+ was enhanced by 30 times compared to that of the cubic phase, clearly demonstrating the crucial role of phase transformation in luminescent performance. This study emphasizes the synergistic optimization of crystal phase and morphology through phase engineering, which substantially improves the upconversion luminescence efficiency of K3GaF6:Er3+, paving the way for further advancements in the design of efficient upconversion materials. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the comprehensive regulation of crystal phase and morphology through co-precipitation and hydrothermal methods. By precisely controlling reaction parameters, including the reaction temperature, time, and composition (<b>left</b>), the synergistic optimization of crystal structure and morphology was achieved, leading to significantly enhanced luminescent performance (<b>right</b>).</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of K<sub>3</sub>GaF<sub>6</sub> samples prepared by the co-precipitation method, at stirring times of 10, 20, and 30 min, and with KHF<sub>2</sub> amounts that are 10, 30, and 50 times the molar amount of Ga(NO<sub>3</sub>)<sub>3</sub>; (<b>b</b>) XRD patterns of K<sub>3</sub>GaF<sub>6</sub> samples prepared by the hydrothermal method, with holding times of 4, 8, and 10 h, and reaction temperatures of 160, 180, and 200 °C; (<b>c</b>) the SEM images and particle size distribution of samples prepared via the hydrothermal method at 160 °C reaction temperatures; (<b>d</b>) the SEM images and particle size distribution of samples prepared via the hydrothermal method at 200 °C reaction temperatures; (<b>e</b>) XRD patterns of Na<sub>x</sub>K<sub>3-x</sub>GaF<sub>6</sub> samples (x = 2.6, 2.5, 2.4, 2.3); (<b>f</b>) XRD patterns of Na<sub>2.5</sub>K<sub>0.5</sub>GaF<sub>6</sub> samples prepared at different reaction temperatures (140, 180, 200 °C).</p>
Full article ">Figure 3
<p>(<b>a</b>) SEM image of the K<sub>3</sub>GaF<sub>6</sub> sample prepared by the co-precipitation method, when the amount of KHF<sub>2</sub> is 30 times that of Ga(NO<sub>3</sub>)<sub>3</sub>; (<b>b</b>) SEM image of the K<sub>3</sub>GaF<sub>6</sub> sample prepared by the co-precipitation method after stirring for 30 min; (<b>c</b>) SEM image of the K<sub>3</sub>GaF<sub>6</sub> sample prepared by the hydrothermal method at a reaction temperature of 200 °C; (<b>d</b>) SEM image of the K<sub>3</sub>GaF<sub>6</sub> sample prepared by the hydrothermal method after 10 h of thermal treatment.</p>
Full article ">Figure 4
<p>(<b>a</b>) Upconversion luminescence spectra of Na<sub>2.5</sub>K<sub>0.5</sub>GaF<sub>6</sub> samples with monoclinic phases at doping concentrations of 0.5%, 1%, 1.5%, 2%, and 2.5% under the excitation of a (<b>a</b>) 980 nm laser; (<b>b</b>) upconversion luminescence spectra of Na<sub>2.5</sub>K<sub>0.5</sub>GaF<sub>6</sub> samples with cubic phases at doping concentrations of 0.5%, 1%, 1.5%, 2%, and 2.5% under the excitation of a (<b>a</b>) 980 nm laser; (<b>c</b>) the bar chart showing the intensity comparison between the monoclinic phase and cubic phase at 657 nm; (<b>d</b>) CIE color coordinates for different crystal phases.</p>
Full article ">
12 pages, 6817 KiB  
Article
Synthesis of Eco-Friendly Narrow-Band CuAlSe2/Ga2S3/ZnS Quantum Dots for Blue Quantum Dot Light-Emitting Diodes
by Shenghua Yuan, Liyuan Liu, Xiaofei Dong, Xianggao Li, Shougen Yin and Jingling Li
Coatings 2025, 15(2), 245; https://doi.org/10.3390/coatings15020245 - 19 Feb 2025
Abstract
Quantum dot light-emitting diodes (QLEDs) based on high-color-purity blue quantum dots (QDs) are crucial for the development of next-generation displays. I-III-VI type QDs have been recognized as eco-friendly luminescent materials for QLED applications due to their tunable band gap and high-stable properties. However, [...] Read more.
Quantum dot light-emitting diodes (QLEDs) based on high-color-purity blue quantum dots (QDs) are crucial for the development of next-generation displays. I-III-VI type QDs have been recognized as eco-friendly luminescent materials for QLED applications due to their tunable band gap and high-stable properties. However, efficient blue-emitting I-III-VI QDs remain rare owing to the high densities of the intrinsic defects and the surface defects. Herein, narrow-band blue-emissive CuAlSe2/Ga2S3/ZnS QDs is synthesized via a facile strategy. The resulting QDs exhibit a sharp blue emission peak at 450 nm with a full width at half maximum (FWHM) of 35 nm, achieved by coating a double-shell structure of Ga2S3 and ZnS, which is associated with the near-complete passivation of Cu-related defects (e.g., Cu vacancies) that enhances the band-edge emission, accompanied by an improvment in photoluminescence quantum yield up to 69%. QLEDs based on CuAlSe2/Ga2S3/ZnS QDs are fabricated, exhibiting an electroluminescence peak at 453 nm with a FWHM of 39 nm, a current efficiency of 3.16 cd A−1, and an external quantum efficiency of 0.35%. This research paves the way for the development of high-efficiency eco-friendly blue QLEDs. Full article
Show Figures

Figure 1

Figure 1
<p>HRTEM images of (<b>a</b>) CuAlSe<sub>2</sub>, (<b>b</b>) CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>, and (<b>c</b>) CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs; (<b>d–f</b>) the corresponding size distribution histograms.</p>
Full article ">Figure 2
<p>XRD patterns of CuAlSe<sub>2</sub>, CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>, and CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs.</p>
Full article ">Figure 3
<p>(<b>a</b>) UV-Vis absorption spectra of CuAlSe<sub>2</sub>, CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>, and CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs; (<b>b</b>) the corresponding (α<span class="html-italic">һν</span>)<sup>2</sup> (where α represent absorption coefficient) as a function of photon energy (<span class="html-italic">һν</span>).</p>
Full article ">Figure 4
<p>(<b>a</b>) Normalized PL spectra of CuAlSe<sub>2</sub>, CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>, and CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs monitored at an excitation wavelength of 370 nm; (<b>b</b>) the corresponding statistical distributions of PL quantum yield (PL QY) for the 8 QDs.</p>
Full article ">Figure 5
<p>(<b>a–c</b>) PL spectra of CuAlSe<sub>2</sub> and CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub> resolved into three subspectra, along with CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs resolved into two subspectra; (<b>d</b>) the corresponding decomposed emission peaks via Gaussian fitting.</p>
Full article ">Figure 6
<p>(<b>a</b>) PL decay traces of CuAlSe<sub>2</sub>, CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>, and CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs monitored at the emission wavelengths of 458, 455, and 450 nm, respectively; (<b>b</b>) TG-DSC curves of CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic structure of device C with CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs as EML; (<b>b</b>) EL spectrum of CuAlSe<sub>2</sub>/Ga<sub>2</sub>S<sub>3</sub>/ZnS QDs-based blue QLEDs and the corresponding PL spectrum; (<b>c</b>) <span class="html-italic">J</span>-<span class="html-italic">L</span>-<span class="html-italic">V</span> curves of QLEDs; (<b>d</b>) Current efficiency and external quantum efficiency <span class="html-italic">versus</span> current density characteristics of the device.</p>
Full article ">
11 pages, 3669 KiB  
Article
The Crystal Structure and Luminescence Behavior of Self-Activated Halotungstates Ba3WO5Cl2 for W-LEDs Applications
by Liuyang Zhang, Shijin Zhou, Jiani Meng, Yuxin Zhang, Jiarui Zhang, Qinlan Ma, Lin Qin and Man Luo
Nanomaterials 2025, 15(4), 311; https://doi.org/10.3390/nano15040311 - 18 Feb 2025
Abstract
The self-activated halotungstate Ba3WO5Cl2 was successfully synthesized using a high-temperature solid-state method. X-ray diffraction analysis (XRD) confirmed the formation of a single-phase compound with a monoclinic crystal structure, ensuring the material’s purity and structural integrity. The luminescence properties [...] Read more.
The self-activated halotungstate Ba3WO5Cl2 was successfully synthesized using a high-temperature solid-state method. X-ray diffraction analysis (XRD) confirmed the formation of a single-phase compound with a monoclinic crystal structure, ensuring the material’s purity and structural integrity. The luminescence properties of Ba3WO5Cl2 were thoroughly investigated using both optical and laser-excitation spectroscopy. The photoluminescent excitation (PLE) and emission (PL) spectra, together with the corresponding decay curves, were recorded across a broad temperature range, from 10 K to 480 K. The charge transfer band (CTB) of the [WO5Cl] octahedron was clearly identified in both the PL and the PLE spectra under ultraviolet light excitation, indicating efficient energy transfer within the material’s structure. A strong blue emission could be detected around 450 nm, which is characteristic of the material’s luminescent properties. However, this emission exhibited thermal quenching as the temperature increased, a common phenomenon where the luminescence intensity diminishes due to thermal effects. To better understand the thermal quenching behavior, variations in luminescence intensity and decay time were analyzed using a straightforward thermal quenching model. This comprehensive study of Ba3WO5Cl2 luminescent properties not only deepens the understanding of its photophysical behavior but also contributes to the development of novel materials with tailored optical properties for specific technological applications. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental (black points) and calculated XRD patterns (red solid line) and their differences (blue solid line) for the Rietveld fit of the Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub> phosphor. The short vertical lines (pink lines) show the position of Bragg reflections of the calculated patterns.</p>
Full article ">Figure 2
<p>(<b>a</b>) The configuration of the atoms in crystal structure of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>; (<b>b</b>,<b>c</b>) SEM images of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>.</p>
Full article ">Figure 3
<p>(<b>a</b>) The calculated energy band structure of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>; (<b>b</b>) DOSs of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub> (Fermi level at 0 energy).</p>
Full article ">Figure 4
<p>(<b>a</b>) The UV-Vis absorption spectrum and PL and PLE spectra of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>; the insert shows the diagram of the device; (<b>b</b>) internal quantum efficiency of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>; inset: zoomed-in fluorescence spectrum around the 460 nm peak.</p>
Full article ">Figure 5
<p>(<b>a</b>) Normalized temperature-dependent emission spectra of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>; (<b>b</b>) calculated emission intensities of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>; (<b>c</b>) normalized temperature-dependent decay curves of Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>; (<b>d</b>) relationship between average lifetime and temperature for Ba<sub>3</sub>WO<sub>5</sub>Cl<sub>2</sub>.</p>
Full article ">
10 pages, 1261 KiB  
Article
Optical Absorption and Luminescence Spectra of Terbium Gallium Garnet TbGaG and Terbium Aluminum Garnet TbAlG
by Nosirjon S. Bozorov, Ismailjan M. Kokanbayev, Akmaljon M. Madaliev, Mavzurjon X. Kuchkarov, Muxtarjan Meliboev, Kobiljon K. Kurbonaliev, Ravshan R. Sultonov, Khayrullo F. Makhmudov, Feruza O. Dadaboyeva, Nargiza Z. Mamadalieva and Shakhlo R. Kukanbaeva
Inorganics 2025, 13(2), 61; https://doi.org/10.3390/inorganics13020061 - 17 Feb 2025
Abstract
In this paper, we investigate the optical absorption and luminescence spectra of rare-earth garnets activated by the terbium (Tb3+) ion, as well as their magneto-optical properties. Crystals of terbium gallium garnet (TbGaG) and terbium aluminum garnet (TbAlG) are considered. The focus [...] Read more.
In this paper, we investigate the optical absorption and luminescence spectra of rare-earth garnets activated by the terbium (Tb3+) ion, as well as their magneto-optical properties. Crystals of terbium gallium garnet (TbGaG) and terbium aluminum garnet (TbAlG) are considered. The focus is on the physical and optical properties and structural features of the energy levels of rare-earth ions in the crystal field of garnets. This work highlights the importance of studying intraconfigurational 4f-4f and interconfigurational 4f-5d transitions, as well as the influence of the crystal field on the magnetic and optical properties of materials. Integrated methods are used, including absorption spectroscopy, luminescence and magneto-optical studies, which allows us to obtain detailed information on the excited states of rare-earth ions. The experimental results show the presence of significant Zeeman shifts, as well as anisotropy of the absorption and luminescence spectra, depending on the orientation of the crystal lattice and the external magnetic field. This work contributes to our understanding of the mechanisms of light absorption and emission in rare-earth garnets, which may facilitate the development of new optoelectronic devices based on them. Full article
(This article belongs to the Special Issue Synthesis and Application of Luminescent Materials, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>The spectrum of the absorption band <sup>7</sup>F<sub>6</sub>→<sup>5</sup>D<sub>4</sub> (<b>left</b>) and the spectrum of the luminescence band <sup>5</sup>D<sub>4</sub>→<sup>7</sup>F<sub>6</sub> (<b>right</b>) in TbGaG, recorded at T = 78 K.</p>
Full article ">Figure 2
<p>Absorption bands <sup>7</sup>F<sub>6</sub>→<sup>5</sup>D<sub>4</sub> at T = 1.8 K. TbGaG. Characteristic features of absorption and luminescence bands are indicated by numbered vertical arrows.</p>
Full article ">Figure 3
<p>A Spectrum of the absorption band <sup>7</sup>F<sub>6</sub>→<sup>5</sup>D<sub>4</sub> in TbAlG, recorded in the right <span class="html-italic">σ</span><sub>+</sub> (solid lines) and left <span class="html-italic">σ</span><sub>−</sub> (dashed lines) polarizations at T = 90K in an external magnetic field H = 7 kOe parallel to the crystallographic axis [001]. In the inset: the field dependence of the Zeeman splitting of the absorption line 1 at T = 90 K.</p>
Full article ">Figure 4
<p>Luminescence spectrum of paramagnetic garnet Tb<sub>3</sub>Ga<sub>5</sub>O<sub>12</sub>, recorded at T = 78 K.</p>
Full article ">Figure 5
<p>Schematic diagram of a modified single-beam spectrophotometer with continuous signal recording: LS—light source; M—monochromator; Z<sub>1</sub>—spherical; Z<sub>2</sub>—flat mirrors of the mirror illuminator; P—polarizer; L<sub>1</sub>—collecting lens; S—sample; PMT—photomultiplier; BS—PMT average current stabilization unit; HVS—high-voltage source (high-voltage rectifier); DCA—direct current amplifier; DV—digital voltmeter; CR—chart recorder; SSM—spectrum scan motor.</p>
Full article ">
19 pages, 3794 KiB  
Article
Generalized Solvent Effect on the Fluorescence Performance of Spiropyran for Advanced Quick Response Code Dynamic Anti-Counterfeiting Sensing
by Junji Xuan, Lingjie Chen and Jintao Tian
Int. J. Mol. Sci. 2025, 26(4), 1531; https://doi.org/10.3390/ijms26041531 - 12 Feb 2025
Abstract
Spiropyran has an attractive and mysterious fluorescence switch and dual-color conversion characteristics, as it exhibits both aggregation-caused quenching (ACQ) in solvents and fluorescence enhancement in polymer matrices. The explanation for this phenomenon has always been of great controversy. Hence, the solvent effect on [...] Read more.
Spiropyran has an attractive and mysterious fluorescence switch and dual-color conversion characteristics, as it exhibits both aggregation-caused quenching (ACQ) in solvents and fluorescence enhancement in polymer matrices. The explanation for this phenomenon has always been of great controversy. Hence, the solvent effect on the emission of spiropyran (SP) was investigated in 16 solvents. By means of molecular orbital theory and the Jablonski diagram, several special parameters (e.g., Hansen solubility parameters and viscosity) were selected for this analysis, with excellent goodness of fit. Subsequently, the main factors that affected the blue shift, red shift, and luminescence efficiency of the emission of the ring-opened form merocyanine (MC) were found to be the hydrogen bonding and polarity, aggregation effect, and viscosity, respectively. A newly modified Jablonski diagram was proposed to clarify the emission behaviors of spiropyran influenced by solvent polarity and isomerization. Meanwhile, the solvent effect could also be extended to a solid polymer matrix (six kinds of polyethylene glycol (PEG) with different molecular weights), which is proposed to be defined as the generalized solvent effect. Accordingly, we have demonstrated that the unique fluorescence properties of spiropyran are dominated by the generalized solvent effect. The security information storage capacity of the simulated quick response (QR) code sensor combined with SP for anti-counterfeiting was significantly improved to six dimensions in taking advantage of the former theoretical analysis. Full article
(This article belongs to the Special Issue Recent Advances in Luminescence: From Mechanisms to Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The spectral properties of <b>SP</b> (7.5 × 10<sup>−5</sup> mol/L) in methanol: (<b>a</b>) absorption spectra (10 min Vis irradiation for <b>SP</b>, and 10 min UV irradiation for <b>MC</b>, the inserts were actual photographs of the experimental samples); (<b>b</b>) emission spectra (ibid); (<b>c</b>) coloration and discoloration processes (the inserts, from left to right, display the actual experimental sample photographs of the coloring process at 0, 2, 4, and 6 min, and the discoloring process at 0.5, 1, 2, 4, 6, 8, 10, and 12 min, respectively); (<b>d</b>) recyclability, the sample was repeatedly treated with 10 min Vis or UV.</p>
Full article ">Figure 2
<p>Changes in the emission peak intensity and λ<sub>max</sub> along with the <b>SP</b> concentrations in methanol after 10 min irradiation using a 365 nm UV lamp. The inserted photographs refer to their corresponding Tyndall effect. The original spectral data can be found in <a href="#app1-ijms-26-01531" class="html-app">Figure S3</a>.</p>
Full article ">Figure 3
<p>The normalized absorption (<b>a</b>) and fluorescence emission spectra (<b>b</b>) of <b>SP</b> (7.5 × 10<sup>−5</sup> mol/L) and their corresponding photographs in 16 solvents after 10 min UV irradiation. The same number refers to the same solvent. MIN and MAX in (<b>b</b>) refer to the spectra of glycerol and hexane with the minimum and maximum λ<sub>max</sub> values of all solvents, respectively. The original spectral data of (<b>a</b>,<b>b</b>) can be found in <a href="#app1-ijms-26-01531" class="html-app">Figure S4a and S4b</a>, respectively.</p>
Full article ">Figure 4
<p>The relationships between λ<sub>max</sub> of the emission band of <b>MC</b> in 16 solvents and different parameters: (<b>a</b>,<b>b</b>) polar force <span class="html-italic">δ</span><sub>P</sub>; (<b>c</b>) hydrogen bonding <span class="html-italic">δ</span><sub>H</sub>; (<b>d</b>) Hansen solubility parameter <span class="html-italic">δ</span><sub>T</sub>; (<b>e</b>,<b>f</b>) a revised parameter <span class="html-italic">δ</span><sub>R</sub> derived from <span class="html-italic">δ</span><sub>T</sub>, the fittings through <span class="html-italic">δ</span><sub>T</sub> were obtained in using statistical method (<b>e</b>) and standard deviation method (<b>f</b>), respectively. The original spectral data used for analyses and fittings can be found in <a href="#app1-ijms-26-01531" class="html-app">Figure S5</a>. Parameters and λ<sub>max</sub> values in different solvents used above and in <a href="#app1-ijms-26-01531" class="html-app">Figure S5 are listed in Table S1</a>.</p>
Full article ">Figure 5
<p>The relationships between <span class="html-italic">K</span> of <b>MC</b> in 16 solvents and different parameters: (<b>a</b>) Hansen solubility parameter <span class="html-italic">δ</span><sub>T</sub>; (<b>b</b>) solvent viscosity; (<b>c</b>) square root of the solvent viscosity. <span class="html-italic">A</span>, <span class="html-italic">F</span>, <span class="html-italic">K</span>, <span class="html-italic">δ</span><sub>T</sub>, and viscosity values in different solvents used above are listed in <a href="#app1-ijms-26-01531" class="html-app">Table S2</a>.</p>
Full article ">Figure 6
<p>The relationships (<b>a</b>,<b>c</b>) between λ<sub>max</sub> of the emission band of <b>MC</b> (c<sub>SP</sub> = 10<sup>−7</sup> mol/g) in PEG ((<b>a</b>) solidified state, (<b>c</b>) seven d after curing) with different molecular weights (Mn = 2000, 4000, 6000, 8000, 10,000, 20,000) and square root of OHV. (<b>b</b>,<b>d</b>) represent the relationship between λ<sub>max</sub> from (<b>a</b>,<b>c</b>) and all its corresponding fluorescence intensities, respectively. The original spectral data of (<b>a</b>,<b>b</b>) and (<b>c</b>,<b>d</b>) can be found in <a href="#app1-ijms-26-01531" class="html-app">Figures S8 and S9</a>, respectively. OHV values of PEG with different molecular weights used above are listed in <a href="#app1-ijms-26-01531" class="html-app">Table S3</a>.</p>
Full article ">Figure 7
<p>Photobleaching of <b>MC</b> (c<sub>SP</sub> = 10<sup>−7</sup> mol/g) in PEG with different molecular weights (Mn = 2000, 4000, 6000, 8000, 10,000, 20,000). The curves were obtained from <a href="#app1-ijms-26-01531" class="html-app">Figure S8a′–f′</a>.</p>
Full article ">Figure 8
<p>QR code-sensing application. All QR code sensors are simulated based on the color changes in the 2 × 2 polymer square grid under different stimulations. Without UV irradiation (middle), the non-fluorescent 2 × 2 polymer square grid can produce weak reversible color changes with dark background and light foreground in visible light (middle left, state 1) or dark (middle right, state 2). With UV irradiation (left and right, state 3), the state of the square grid could be reversed with presentation of the dark foreground and light background both in its emission (left) and coloration (right). Meanwhile, the process is also reversible. Three states exist, along with one contrast change and four kinds of color changes.</p>
Full article ">Scheme 1
<p>Photoisomerization process between the two isomers of <b>SP</b>. The synthesis and characterization of <b>SP</b> are shown in <a href="#app1-ijms-26-01531" class="html-app">Scheme S1, Figures S1 and S2</a>.</p>
Full article ">Scheme 2
<p>The influence of solvent polarity on the orbital energy of π–π* and n–π* electron transition corresponding to the characteristic molecular structure of <b>SP</b>, <b>MC</b><sub>α</sub>, and <b>MC</b><sub>β</sub>.</p>
Full article ">Scheme 3
<p>A simplified Jablonski diagram with consideration of the solvent polarity and isomerization of <b>SP</b>. The <b>SP</b><sub>1</sub>*/<b>SP</b><sub>2</sub>* and <b>MC</b><sub>1</sub>*/<b>MC</b><sub>2</sub>* refer to the excited state of the round states <b>SP</b><sub>1</sub>/<b>SP</b><sub>2</sub> and <b>MC</b><sub>1</sub>/<b>MC</b><sub>2</sub> stimulated by UV, respectively. Different solid and dashed arrows denote different processes: (1) absorption (e.g., 365 nm UV); (2) photoisomerization, <b>SP</b>* → <b>MC</b>; (3) vibrational relaxation; (4) fluorescence (radiative transition); (5) external conversion (non-radiative de-excitation); (6) solvent-induced isomerization (non-radiative de-excitation), <b>MC</b>* → <b>SP</b>.</p>
Full article ">
18 pages, 3092 KiB  
Review
Pigments and Near-Infrared Phosphors Based on Mn5+
by Sanja Kuzman, Tatjana Dramićanin, Anatoli I. Popov, Mikhail G. Brik and Miroslav D. Dramićanin
Nanomaterials 2025, 15(4), 275; https://doi.org/10.3390/nano15040275 - 11 Feb 2025
Abstract
The optical properties of Mn5+ ions, which are responsible for the intense green–turquoise–blue coloration of Mn5+-based pigments and the near-infrared emission of phosphors, are the focus of this article. Mn5+ ions enter crystalline matrices in four-fold coordinated positions and [...] Read more.
The optical properties of Mn5+ ions, which are responsible for the intense green–turquoise–blue coloration of Mn5+-based pigments and the near-infrared emission of phosphors, are the focus of this article. Mn5+ ions enter crystalline matrices in four-fold coordinated positions and can maintain their 5+ valence state when crystalline hosts meet the conditions described in this work. Mn5+ ions have [Ar]3d2 electronic configuration and always experience a strong crystal field due to a high electric charge; therefore, their lower electronic states have the 3A2 < 1E < 1A1 < 3T2 < 3T1 progression in energy. We present the properties of several Mn5+-based pigments and discuss the electronic transitions responsible for their coloration. Specifically, we show that the color is determined by the spin-allowed 3A23T1(3F) absorption, which extends across the orange–red–deep red spectral region and is strongly influenced by crystal field strength. The narrow-band emission Mn5+-activated near-infrared phosphors arise from the spin-forbidden 1E → 3A2 transition, whose energy is independent of the crystal field strength and determined by the nephelauxetic effect. We demonstrate the linear relationship between 1E state energy and the nephelauxetic parameter β1 using Racah parameter literature data for Mn5+ phosphors. Lastly, we address the recent applications of these Mn5+ phosphors in luminescence thermometry. Full article
(This article belongs to the Section Nanophotonics Materials and Devices)
Show Figures

Figure 1

Figure 1
<p>Tanabe–Sugano diagram for the Mn<sup>5+</sup> ions in the tetrahedral coordination.</p>
Full article ">Figure 2
<p>Kubelka–Munk transformation of the diffuse reflectance spectrum of the Ca<sub>6</sub>Ba(PO<sub>4</sub>)<sub>4</sub>O:Mn<sup>5+</sup> (<b>left</b>) and Mn<sup>5+</sup> energy levels and electronic transitions responsible for color and photoluminescence emission of Ca<sub>6</sub>Ba(PO<sub>4</sub>)<sub>4</sub>O:Mn<sup>5+</sup> (<b>right</b>).</p>
Full article ">Figure 3
<p>(<b>a</b>) The correlation between the energy of the <sup>1</sup>E state and (8<span class="html-italic">B</span> + 2<span class="html-italic">C</span>); the full line shows the linear correlation with the slope of 0.9817 ≈ 1, and (<b>b</b>) the dependence of the <sup>1</sup>E state energy on the nephelauxetic parameter <span class="html-italic">β</span><sub>1</sub> for Mn<sup>5+</sup> in different hosts listed in <a href="#nanomaterials-15-00275-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Diffuse reflectance spectra of Ba<sub>2</sub>In<sub>2−x</sub>Mn<sub>x</sub>O<sub>5+x</sub> samples and their corresponding powder color variations: (<b>a</b>) Ba<sub>2</sub>In<sub>2−x</sub>Mn<sub>x</sub>O<sub>5+x</sub> (x = 0, 0.1, 0.2, 0.3), where the spectrum for x = 0 (Ba<sub>2</sub>In<sub>2</sub>O<sub>5</sub>, white color) is shown for comparison; and (<b>b</b>) Ba<sub>2</sub>In<sub>2−x</sub>Mn<sub>x</sub>O<sub>5+x</sub> (x = 0.4, 0.5, 0.6, 0.7). Reprinted with permission from Ref. [<a href="#B60-nanomaterials-15-00275" class="html-bibr">60</a>]. 2013, American Chemical Society.</p>
Full article ">Figure 5
<p>(<b>a</b>) L*, a*, b* color parameters of Ba<sub>5</sub>Mn<sub>3-x</sub>V<sub>x</sub>O<sub>12</sub>Cl and Ba<sub>5</sub>Mn<sub>3−x</sub>P<sub>x</sub>O<sub>12</sub>Cl (x = 0; 0.5; 1; 1.5; 2.0; 2.5; 2.8; 2.9; 3.0) samples as a function of the Mn content (x) and color changes with Mn<sup>5+</sup> doping. (<b>b</b>) Diffuse-reflectance spectra of the Ba<sub>5</sub>Mn<sub>3−x</sub>V<sub>x</sub>O<sub>12</sub>Cl series; (<b>c</b>) UV-vis and NIR reflectance (%) of Ba<sub>5</sub>Mn<sub>3−x</sub>V<sub>x</sub>O<sub>12</sub>Cl (x = 0, 2.5, 3) and Ba<sub>5</sub>Mn<sub>3−x</sub>P<sub>x</sub>O<sub>12</sub>Cl (x = 0, 2.5, 3) samples as a function of wavelength (nm). Adapted with permission from Ref. [<a href="#B69-nanomaterials-15-00275" class="html-bibr">69</a>]. 2016, Elsevier.</p>
Full article ">Figure 6
<p>Photographs of the Ca<sub>6</sub>Ba(P<sub>1−x</sub>Mn<sub>x</sub>)<sub>4</sub>O<sub>17</sub> (0 ≤ x ≤ 0.13) pigments. Tables present color coordination values in the <span class="html-italic">L*</span>, <span class="html-italic">a*</span>, <span class="html-italic">b*</span>, <span class="html-italic">C, H</span>⁰, and Δ<span class="html-italic">E</span> system of Ca<sub>6</sub>Ba(P<sub>0.99</sub>Mn<sub>0.01</sub>)<sub>4</sub>O<sub>17</sub> pigments after the thermal stability test and acid and alkali resistance test. Adapted with permission from Ref. [<a href="#B71-nanomaterials-15-00275" class="html-bibr">71</a>]. 2017, Elsevier.</p>
Full article ">Figure 7
<p>(<b>a</b>) Photoluminescence emission spectra of Ca<sub>6</sub>Ba(PO<sub>4</sub>)<sub>4</sub>O:Mn<sup>5+</sup> powder measured at different temperatures; (<b>b</b>) luminescence intensity ratio (LIR) as a function of temperature (experimental data—diamond markers). The insert shows the LIR distribution histogram measured at 303.15 K (30 °C)—filled diamond marker; (<b>c</b>) calculated absolute and relative sensitivities (marked values at 303.15 K (30 °C)). Reprinted from Ref. [<a href="#B42-nanomaterials-15-00275" class="html-bibr">42</a>].</p>
Full article ">
12 pages, 3699 KiB  
Article
Preparation of Glass-Ceramic Materials by Controlled Crystallization of Eu2O3-Doped WO3-B2O3-La2O3 Glasses and Their Luminescent Properties
by Aneliya Yordanova, Margarita Milanova, Lyubomir Aleksandrov, Reni Iordanova, Peter Tzvetkov, Pavel Markov and Petia Petrova
Molecules 2025, 30(4), 832; https://doi.org/10.3390/molecules30040832 - 11 Feb 2025
Abstract
In this paper, the crystallization behavior of 52WO3:22B2O3:26La2O3:0.5Eu2O3 glass has been investigated in detail by XRD and TEM analysis. The luminescent properties of the resulting glass-ceramics were also investigated. By [...] Read more.
In this paper, the crystallization behavior of 52WO3:22B2O3:26La2O3:0.5Eu2O3 glass has been investigated in detail by XRD and TEM analysis. The luminescent properties of the resulting glass-ceramics were also investigated. By XRD and TEM analysis, crystallization of β-La2W2O9 and La2WO6 crystalline phases has been proved. Photoluminescent spectra showed increased emission in the resulting glass-ceramic samples compared to the parent glass sample due to higher asymmetry of Eu3+ ions in the obtained crystalline phases, where the active Eu3+ ions are incorporated. Also, in the glass-ceramics, the crystalline particles are embedded in the amorphous matrix and more of them are separated from each other which improves the light scattering intensity from the free interfaces of the nanocrystallites, resulting in the enhancement of the PL intensity. It was established that the optimum emission intensity is registered for glass-ceramic samples obtained after an 18 h heat treatment of the parent glass. After 21 h of glass crystallization, the amount of crystallite particles is high enough, and they are in close proximity to each other, and hence, the average distance between europium ions decreases, resulting in quenching of Eu3+ and a decrease in the emission intensity. Additionally, at 21 h of glass crystallization, formation of new crystalline phase—La2WO6 is established. A redistribution of Eu3+ ions in the different crystalline compounds is most likely taking place, which is also not favorable for the emission intensity. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Figure 1

Figure 1
<p>DTA curves of 52WO<sub>3</sub>:22B<sub>2</sub>O<sub>3</sub>:26La<sub>2</sub>O<sub>3</sub>:0.5Eu<sub>2</sub>O<sub>3</sub> glass.</p>
Full article ">Figure 2
<p>XRD patterns of 52WO<sub>3</sub>:22B<sub>2</sub>O<sub>3</sub>:26La<sub>2</sub>O<sub>3</sub>:0.5Eu<sub>2</sub>O<sub>3</sub> glass ceramics crystallized at 680 °C with different durations. Indexed peaks represent the La<sub>2</sub>W<sub>2</sub>O<sub>9</sub> cubic phase. For comparison, vertical red bars correspond to Bragg peaks of β-La<sub>2</sub>W<sub>1.7</sub>Mo<sub>0.3</sub>O<sub>9</sub>. Asterisk (*) represents the most intensive (312) peak of orthorhombic La<sub>2</sub>WO<sub>6</sub> (PDF # 00-057-1075).</p>
Full article ">Figure 3
<p>Excitation spectra of 0.5% Eu<sup>3+</sup>-doped 52WO<sub>3</sub>:22B<sub>2</sub>O<sub>3</sub>:26La<sub>2</sub>O<sub>3</sub> glass and glass-ceramic samples.</p>
Full article ">Figure 4
<p>Emission spectra of 0.5% Eu<sup>3+</sup>-doped 52WO<sub>3</sub>:22B<sub>2</sub>O<sub>3</sub>:26La<sub>2</sub>O<sub>3</sub> glass and glass-ceramic samples.</p>
Full article ">Figure 5
<p>CIE chromaticity diagram of Eu<sup>3+</sup>-doped 52WO<sub>3</sub>:22B<sub>2</sub>O<sub>3</sub>:26La<sub>2</sub>O<sub>3</sub>:0.5Eu<sub>2</sub>O<sub>3</sub> glass and corresponding glass-ceramics.</p>
Full article ">Figure 6
<p>Bright field micrograph (<b>a</b>) and HRTEM (<b>b</b>) of nanosized particles of La<sub>2</sub>W<sub>2</sub>O<sub>9</sub>.</p>
Full article ">Figure 7
<p>Bright field micrograph (<b>a</b>) and HRTEM (<b>b</b>) of La<sub>2</sub>W<sub>2</sub>O<sub>9</sub> and La<sub>2</sub>WO<sub>6</sub>.</p>
Full article ">Figure 8
<p>Bright field micrograph (<b>a</b>) and particle size distribution (<b>b</b>) of GC-18 h.</p>
Full article ">
13 pages, 3806 KiB  
Article
Influence of the Annealing Temperature on the Properties of {ZnO/CdO}30 Superlattices Deposited on c-Plane Al2O3 Substrate by MBE
by Anastasiia Lysak, Aleksandra Wierzbicka, Piotr Dłużewski, Marcin Stachowicz, Jacek Sajkowski and Ewa Przezdziecka
Crystals 2025, 15(2), 174; https://doi.org/10.3390/cryst15020174 - 10 Feb 2025
Abstract
{CdO/ZnO}m superlattices (SLs) have been grown on c-plane sapphire substrates by plasma-assisted molecular beam epitaxy (PA-MBE). The observation of satellite peaks in the XRD studies of the as-grown and annealed samples confirms the presence of a periodic superlattice structure. The properties [...] Read more.
{CdO/ZnO}m superlattices (SLs) have been grown on c-plane sapphire substrates by plasma-assisted molecular beam epitaxy (PA-MBE). The observation of satellite peaks in the XRD studies of the as-grown and annealed samples confirms the presence of a periodic superlattice structure. The properties of as-grown and annealed SLs deposited on c-oriented sapphire were investigated by transmission electron microscopy, X-ray diffraction and temperature dependent PL studies. The deformation of the SLs structure was observed after rapid thermal annealing. As the thermal annealing temperature increases, the diffusion of Cd ions from the quantum well layers into the ZnO barrier increases. The formation of CdZnO layers causes changes in the luminescence spectrum in the form of peak shifts, broadening and changes in the spacing of the satellite peaks visible in X-ray analysis. Full article
(This article belongs to the Special Issue Materials and Devices Grown via Molecular Beam Epitaxy)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Cross-sectional HAADF/STEM image of <span class="html-italic">as-grown</span> {ZnO/CdO}<sub>30</sub> SL. (<b>b</b>) Fourier transform of <a href="#crystals-15-00174-f001" class="html-fig">Figure 1</a>a, where white arrows indicate the positions of spatial frequencies corresponding to the SL periodicity. (<b>c</b>) HAADF/STEM cross section at higher magnification. The enlarged area of SL shows the thickness of the CdO and ZnO layers.</p>
Full article ">Figure 2
<p>X-ray diffraction θ/2θ scans of <span class="html-italic">as-grown</span> and annealed {ZnO/CdO}<sub>30</sub> SLs deposited on <span class="html-italic">c</span>-plane Al<sub>2</sub>O<sub>3</sub> substrate (black vertical dotted lines correspond to the peak positions of wurtzite ZnO (JCPDS Card 00-005-0664), whereas * indicates the peaks originated from the <span class="html-italic">c</span>-plane Al<sub>2</sub>O<sub>3</sub> substrate (JCPDS Card 00-050-0792).</p>
Full article ">Figure 3
<p>High resolution XRD 2θ/ϖ scans of the 00.2 {ZnO/CdO} SL peaks. The XRD experimental data are shown as solid lines.</p>
Full article ">Figure 4
<p>(<b>a</b>) Normalized PL spectra of <span class="html-italic">as-grown</span> and annealed {ZnO/CdO}<sub>30</sub> SLs measured at ~10 K. Temperature-dependent PL spectra of SLs annealed at different temperatures: (<b>b</b>) <span class="html-italic">as-grown</span> structure; annealed structure at: (<b>c</b>) 500 °C, (<b>d</b>) 600 °C, (<b>e</b>) 700 °C and (<b>f</b>) 900 °C.</p>
Full article ">Figure 5
<p>Temperature-dependent energy position of the PL peaks observed in {ZnO/CdO}<sub>30</sub> SLs annealed at different temperatures: (<b>a</b>) <span class="html-italic">as-grown</span>; (<b>b</b>) 500 °C; (<b>c</b>) 600 °C; and (<b>d</b>) 700 °C.</p>
Full article ">
19 pages, 5119 KiB  
Review
Carbon Quantum Dots: Synthesis, Characteristics, and Quenching as Biocompatible Fluorescent Probes
by Arif Kamal, Seongin Hong and Heongkyu Ju
Biosensors 2025, 15(2), 99; https://doi.org/10.3390/bios15020099 - 10 Feb 2025
Abstract
Carbon quantum dots (CQDs), a new class of carbon-based nanomaterials, have emerged as nano-scaled probes with photoluminescence that have an eco-friendly and bio-compatible nature. Their cost-efficient synthesis and high photoluminescence quantum yields make them indispensable due to their application in opto-electronic devices, including [...] Read more.
Carbon quantum dots (CQDs), a new class of carbon-based nanomaterials, have emerged as nano-scaled probes with photoluminescence that have an eco-friendly and bio-compatible nature. Their cost-efficient synthesis and high photoluminescence quantum yields make them indispensable due to their application in opto-electronic devices, including biosensors, bioimaging, environmental monitoring, and light sources. This review provides intrinsic properties of CQDs such as their excitation-dependent emission, biocompatibility, and quenching properties. Diverse strategies for their easy synthesis are divided into bottom-up and top-down approaches and detailed herein. In particular, we highlight their luminescence properties, including quenching mechanisms that could even be utilized for the precise and rapid detection of biomolecules. We also discuss methodologies for the mitigation of fluorescence quenching, which is pivotal for the application of CQDs in biosensors and bioimaging. Full article
Show Figures

Figure 1

Figure 1
<p>Synthesis and potential applications of carbon quantum dots (CQDs).</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic representation of a CQD core (grey) and its small domains made of fully (<b>b</b>) sp3- and (<b>c</b>) sp2-hybridized carbon atoms. Blue and red spheres are C atoms, and white spheres are H atoms. The angle between the C–C bonds is 109.5° in (<b>b</b>) and 120° in (<b>c</b>). Reprinted with permission from reference [<a href="#B2-biosensors-15-00099" class="html-bibr">2</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 3
<p>CQD synthesis methods: top-down (<b>left</b>) and bottom-up (<b>right</b>) approaches.</p>
Full article ">Figure 4
<p>Laser ablation of carbon targets for CQD synthesis. Reprinted with permission from reference [<a href="#B32-biosensors-15-00099" class="html-bibr">32</a>]. Copyright 2011 Royal Society of Chemistry.</p>
Full article ">Figure 5
<p>Hydrothermal synthesis of CQDs from alkali lignin obtained from spruce tree for formaldehyde detection. Reprinted with permission from reference [<a href="#B48-biosensors-15-00099" class="html-bibr">48</a>]. Copyright 2024 Royal Society of Chemistry.</p>
Full article ">Figure 6
<p>Surface-functionalized N-doped CQDs synthesized by one-step microwave method. Reprinted with permission from reference [<a href="#B7-biosensors-15-00099" class="html-bibr">7</a>]. Copyright 2009 Royal Society of Chemistry.</p>
Full article ">Figure 7
<p>Sono-chemical synthesis of CQDs for luminescence of various colors. Reprinted with permission from reference [<a href="#B61-biosensors-15-00099" class="html-bibr">61</a>]. Copyright 2022 Nanomaterials.</p>
Full article ">Figure 8
<p>(<b>a</b>) Steady-state absorption, (<b>b</b>,<b>c</b>) emission spectra of CQDs in ethanol for various excitation wavelengths. Reprinted with permission from reference [<a href="#B71-biosensors-15-00099" class="html-bibr">71</a>]. Copyright 2016 American Chemical Society.</p>
Full article ">Figure 9
<p>CQDs PL spectra at 405 nm (<b>a</b>), 488 nm (<b>b</b>), 638 nm (<b>c</b>); FWHM distruibution (<b>d</b>) and proposed CQD structure (<b>e</b>). Reprinted with permission from Ref. [<a href="#B73-biosensors-15-00099" class="html-bibr">73</a>]. Copyright 2017 SMALL.</p>
Full article ">Figure 10
<p>Statistics of the effects of PbI<sub>2</sub> on (<b>A</b>) zebrafish embryos (at 7 days post-fertilization) and (<b>C</b>) Japanese medaka embryos at 16 days as well as their hatching rates ((<b>B</b>) and (<b>D</b>), respectively). PbI<sub>2</sub> exposure results in developmental failure and an atypical appearance of embryos. The differences between controlled and exposed groups (<span class="html-italic">p</span> &lt; 0.05) are indicated with asterisks. Reprinted with permission from reference [<a href="#B92-biosensors-15-00099" class="html-bibr">92</a>]. Copyright 2021 Elsevier.</p>
Full article ">Figure 11
<p>(<b>A</b>) Bright-field image of zebra fish incubated with CQDs, (<b>B</b>) fluorescence image and (<b>C</b>) merged image of a zebrafish incubated with CQDs at λ<sub>max</sub> = 488 nm [<a href="#B93-biosensors-15-00099" class="html-bibr">93</a>]; <span class="html-italic">Int. J. Nanomed.</span> <b>2020</b>, <span class="html-italic">15</span>, 6519–6529. Originally published by and used with permission from Dove Medical Press Ltd.</p>
Full article ">Figure 12
<p>Stern–Volmer plots for the fluorescence quenching of CQDs by 2,4-dinitrotoluene at 425 nm (○) and 407 nm excitation (●) (upper panel) and by DEA at 400 nm excitation in methanol (○) and chloroform (□) and at 407 nm excitation in methanol (●) (lower panel). The low-concentration portion of the same plot for DEA-induced quenching in methanol (-·-) is shown for comparison. Solid and dashed lines are the best fit of the respective data. Reprinted with permission of minor edition from reference [<a href="#B27-biosensors-15-00099" class="html-bibr">27</a>]. Copyright 2009 Royal Society of Chemistry.</p>
Full article ">
13 pages, 7368 KiB  
Article
Effects of High Temperature and High Pressure on the Photoluminescence of CdTe Quantum Dots: Implication for the High-Temperature Resistance Application of Nano-Stress Sensing Materials
by Jundiao Wang, Ke Bao, Yue Liu, Feihong Mao and Peirong Ren
Materials 2025, 18(4), 746; https://doi.org/10.3390/ma18040746 - 8 Feb 2025
Abstract
Nano-sized quantum dots (QDs) have the potential for the application of stress sensing materials based on their pressure-sensitive photoluminescence (PL) properties, while the influence of a more realistic loading environment on the PL characteristics of QDs under a high-temperature environment remains to be [...] Read more.
Nano-sized quantum dots (QDs) have the potential for the application of stress sensing materials based on their pressure-sensitive photoluminescence (PL) properties, while the influence of a more realistic loading environment on the PL characteristics of QDs under a high-temperature environment remains to be further studied. Herein, we studied the PL response of CdTe QDs under repetitive loading–unloading conditions under high-temperature coupling to explore the stability of its high temperature stress sensing potential. The results show that the CdTe QDs with size of 3.2 nm can detect pressure in the range of 0–5.4 GPa, and the pressure sensitivity coefficient of PL emission peak energy (EPL) is about 0.054 eV/GPa. Moreover, the relationship between EPL and pressure of CdTe QDs is not sensitive to high temperature and repeated loading, which meets the stability requirements of the sensing function required for stress sensing materials under high temperature. However, the disappearance of PL intensity caused by spontaneous growth as well as the ligand instability of QDs induced by high temperature/high pressure affects the availability of EPL, which has a great influence on the application of CdTe QDs as high-temperature-resistant nano-stress sensing materials. The research provides the mechanical luminescence response mechanism of CdTe QDs under high-temperature/high-pressure coupling conditions, which provides experimental support for the design of high-temperature/high-pressure-resistant QD structures. Full article
(This article belongs to the Special Issue Fatigue, Damage and Fracture of Alloys)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Transmission electron microscope and high-resolution transmission electron microscope images of CdTe QDs. The scale bars in the transmission electron microscope and high-resolution transmission electron microscope correspond to 20 nm and 2 nm, respectively [<a href="#B23-materials-18-00746" class="html-bibr">23</a>]. (<b>b</b>) Schematic of the composition of the experimental detection platform for the PL response of QDs under high -temperature/high-pressure.</p>
Full article ">Figure 2
<p>(<b>a</b>) Response of PL spectra of CdTe QDs with pressure. (<b>b</b>) Comparison of PL spectra of CdTe QDs before and after loading. (<b>c</b>) Response of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>P</mi> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math> of CdTe QDs with pressure during loading–unloading.</p>
Full article ">Figure 3
<p>(<b>a</b>) Response of PL intensity of CdTe QDs with loading pressure during three continuous loading–unloading processes. The positions marked by the arrows represent luminescence quenching. (<b>b</b>) The evolution of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>P</mi> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math> of CdTe QDs with pressure under three loading–unloading cycles (dots) and the corresponding fitting results (lines). The solid and hollow symbols represent loading and unloading results, respectively. The solid and dashed lines represent the fitting results of loading and unloading, respectively. Colors of red, blue and green represent the results of the first, second and third loading–unloading, respectively. The directions of the solid and dashed arrows represent the directions of loading and unloading, respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) Response of the PL spectra of CdTe QDs with increasing temperature. (<b>b</b>) Comparison of the PL spectra of CdTe QDs before and after the heating–cooling process.</p>
Full article ">Figure 5
<p>(<b>a</b>) Response of the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>P</mi> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math> of CdTe QDs with temperature during the heating–cooling process. (<b>b</b>) Comparison of the PL spectra of CdTe QDs before and after the heating–cooling process.</p>
Full article ">Figure 6
<p>(<b>a</b>) Response of PL intensity of CdTe QDs with pressure at different temperatures. The positions marked by arrows represent luminescence quenching. (<b>b</b>) Response of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>P</mi> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math> of CdTe QDs with pressure at different temperatures. (<b>c</b>) Response of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>P</mi> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math> of CdTe QDs with pressure under repetitive loading–unloading and the fitting results (<span class="html-italic">T</span> = 313 K).</p>
Full article ">
10 pages, 2021 KiB  
Communication
Stable Fluorenyl Radicals Showing Tunable Doublet Emission
by Xudong Hou, Tingting Xu, Jun Zhu, Shaofei Wu and Jishan Wu
Chemistry 2025, 7(1), 21; https://doi.org/10.3390/chemistry7010021 - 7 Feb 2025
Abstract
Neutral organic radicals with intrinsic spin-allowed doublet emission have emerged as a promising class of luminescent materials, garnering significant research interest. However, the development of stable luminescent radicals with tunable emission remains challenging. Herein, we present the synthesis of a series of 9-(2,4,6-trichlorophenyl)-substituted [...] Read more.
Neutral organic radicals with intrinsic spin-allowed doublet emission have emerged as a promising class of luminescent materials, garnering significant research interest. However, the development of stable luminescent radicals with tunable emission remains challenging. Herein, we present the synthesis of a series of 9-(2,4,6-trichlorophenyl)-substituted fluorenyl radicals functionalized with various substituents at the 3,6-positions. These radicals exhibit enhanced stability through efficient spin delocalization and kinetic protection. Notably, they display red-shifted photoluminescence compared to traditional polychlorotriphenylmethyl radicals, with maximum emission wavelengths ranging from 679 nm to 744 nm. The mechanisms underlying the doublet emission, as well as their electrochemical properties, have been thoroughly investigated. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The structures of some reported stable radicals with room-temperature doublet emission, and the new luminescent fluorenyl radicals in this report.</p>
Full article ">Figure 2
<p>(<b>a</b>) UV−vis absorption and normalized fluorescence (FL) spectra of <b>FL-R1</b>, <b>FL-R2</b>, <b>FL-R3</b>, and <b>FL-R4</b> measured in DCM. The excitation wavelengths for FL are 380, 384, 389, and 371 nm, respectively; (<b>b</b>) cyclic voltammograms of <b>FL-R1</b>, <b>FL-R2</b>, <b>FL-R3</b>, and <b>FL-R4</b> measured in DCM (with 0.1 M <span class="html-italic">n</span>Bu<sub>4</sub>N•PF<sub>6</sub> as supporting electrolyte, scan rate: 100 mV/s); (<b>c</b>) ESR spectrum of <b>FL-R1</b> recorded in DCM at room temperature; (<b>d</b>) calculated (UB3LYP/6-31G(d,p)) spin-density distribution map for the <b>FL-R1</b>.</p>
Full article ">Figure 3
<p>(<b>a</b>) The calculated (UB3LYP/6-31G(d,p)) frontier molecular orbital profiles and energy diagram of <b>FL-R1</b>; (<b>b</b>) a summary of the calculated UB3LYP/6-31G(d,p)) energy diagrams of <b>FL-R1</b>, <b>FL-R2</b>, <b>FL-R3</b>, and <b>FL-R4</b>.</p>
Full article ">Figure 4
<p>The time-dependent decay of the absorbance at the absorption maximums of <b>FL-R1</b>, <b>FL-R2</b>, <b>FL-R3</b>, and <b>FL-R4</b> in DCM (1 × 10<sup>−4</sup> M) under ambient air and light conditions.</p>
Full article ">Scheme 1
<p>The synthetic route of the new fluorenyl radicals.</p>
Full article ">
14 pages, 3607 KiB  
Article
Self-Enhanced Near-Infrared Copper Nanoscale Electrochemiluminescence Probe for the Sensitive Detection of Ciprofloxacin in Foods
by Jie Wu, Yuanjie Qin, Xiaoxin Mei, Lin Cai, Wen Hao and Guozhen Fang
Foods 2025, 14(3), 538; https://doi.org/10.3390/foods14030538 - 6 Feb 2025
Abstract
Ciprofloxacin (CIP), a widely used broad-spectrum antibiotic, poses a serious threat to human health and environmental safety due to its residues. The complementary monomers molecularly imprinted electrochemiluminescence sensor (MIECLS) based on a polyvinylpyrrolidone-functionalized copper nanowires (CuNWs@PVP) luminescent probe was constructed for the ultra-sensitive [...] Read more.
Ciprofloxacin (CIP), a widely used broad-spectrum antibiotic, poses a serious threat to human health and environmental safety due to its residues. The complementary monomers molecularly imprinted electrochemiluminescence sensor (MIECLS) based on a polyvinylpyrrolidone-functionalized copper nanowires (CuNWs@PVP) luminescent probe was constructed for the ultra-sensitive detection of CIP. CuNWs with low cost and high conductivity exhibited near-infrared electrochemiluminescence (NIR ECL) properties, yet their self-aggregation and oxidation led to a weakened emission phenomenon. PVP with solvent affinity and large skeleton was in situ attached to CuNWs surface to avoid CuNWs sedimentation and aggregation, and self-enhanced ECL signals were achieved. The bifunctional monomers molecularly imprinted polymer (MIP) possessed complementary active centers that increased their affinity with CIP, enhancing the accurate and sensitive detection of the target substances. The linear range of CIP using MIECLS was 5.00 × 10−9–5.00 × 10−5 mol L−1 with a low limit of detection (LOD) of 2.59 × 10−9 mol L−1, while the recovery rates of CIP in the spiking recovery experiment were 84.39% to 92.48%. The combination of bifunctional monomer MIP and NIR copper-based nano-luminescent probe provides a new method for the detection of CIP in food. Full article
(This article belongs to the Special Issue Food Contaminants: Detection, Toxicity and Safety Risk Assessment)
Show Figures

Figure 1

Figure 1
<p>The SEM images of CuNWs (<b>A</b>) and CuNWs@PVP (<b>B</b>); the TEM image of CuNWs@PVP (<b>C</b>); the SEM element mapping of CuNWs@PVP ((<b>D</b>–<b>F</b>), red signifies Cu, purple signifies C, and green signifies N); the SEM images before (<b>G</b>) and after (<b>H</b>) elution of MIP/CuNWs@PVP/GCE.</p>
Full article ">Figure 2
<p>ECL spectrum with voltage variation (<b>A</b>); ECL intensity under different wavelength filters (<b>B</b>); the Fourier transform infrared spectrum of CuNWs@PVP (<b>C</b>); MIP electropolymerization for 10 cycles (<b>D</b>).</p>
Full article ">Figure 3
<p>The ECL spectra (<b>A</b>), EIS curves (<b>D</b>), and CV curves (<b>G</b>) of different modified electrodes; the ECL spectra (<b>B</b>), EIS curves (<b>E</b>), and CV curves (<b>H</b>) of MIP/CuNWs@PVP/GCE; the ECL spectra (<b>C</b>), EIS curves (<b>F</b>), and CV curves (<b>I</b>) of NIP/CuNWs@PVP/GCE.</p>
Full article ">Figure 4
<p>Optimization of ratio between PVP and CuNWs (<b>A</b>); the concentration optimization for CuNWs@PVP (<b>B</b>); optimization of the ratio of template molecules (CIP) to bifunctional monomers (o-PD: Py) (<b>C</b>); optimization of the number of cycles of MIP electropolymerization (<b>D</b>); optimization of eluent types (<b>E</b>); optimization of elution time and readsorption time (<b>F</b>).</p>
Full article ">Figure 5
<p>ECL responses of MIP/CuNWs@PVP/GCE under different CIP concentrations (<b>A</b>); standard curve for detecting CIP (<b>B</b>); stability test of CuNWs@PVP/GCE (a), and elution (b) and readsorption (c) of MIP/CuNWs@PVP/GCE (<b>C</b>); molecular formulas of CIP and its interfering substances (<b>D</b>); selective test (OFX: ofloxacin, LMF: lomefloxacin, ENX: enoxacin, and NOR: norfloxacin) of MIECLS (<b>E</b>); reproducibility test of MIECLS (<b>F</b>).</p>
Full article ">Scheme 1
<p>The design principle of MIP/CuNWs@PVP/GCE and the specific detection mechanism for CIP.</p>
Full article ">
Back to TopTop