[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (212)

Search Parameters:
Keywords = lead-free perovskite

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 6120 KiB  
Communication
Identifying Crystal Structure of Halides of Strontium and Barium Perovskite Compounds with EXPO2014 Software
by Jorge A. Perez Franco, Antonieta García Murillo, Felipe de J. Carrillo Romo, Issis C. Romero Ibarra and Arturo Cervantes Tobón
Materials 2025, 18(1), 58; https://doi.org/10.3390/ma18010058 - 26 Dec 2024
Viewed by 429
Abstract
The synthesis of ethylamine-based perovskites has emerged to attempt to replace the lead in lead-based perovskites for the alkaline earth elements barium and strontium, introducing chloride halide to prepare the perovskites in solar cell technology. X-ray diffraction studies were conducted, and EXPO2014 software [...] Read more.
The synthesis of ethylamine-based perovskites has emerged to attempt to replace the lead in lead-based perovskites for the alkaline earth elements barium and strontium, introducing chloride halide to prepare the perovskites in solar cell technology. X-ray diffraction studies were conducted, and EXPO2014 software was utilized to resolve the structure. Chemical characterization was performed using Fourier transform infrared spectroscopy, photophysical properties were analyzed through ultraviolet–visible spectroscopy, and photoluminescence properties were determined to confirm the perovskite characteristics. The software employed can determine new crystal structures, as follows: orthorhombic for barium perovskite CH3CH2NH3BaCl3 and tetragonal for strontium perovskite CH3CH2NH3SrCl3. The ultraviolet–visible spectroscopy data demonstrated that a temperature increase (90–110 °C) contributed to reducing the band gap from 3.93 eV to 3.67 eV for barium perovskite and from 4.05 eV to 3.84 eV for strontium perovskite. The results exhibited that new materials can be obtained through gentle chemistry and specialized software like EXPO2014, both of which are capable of conducting reciprocal and direct space analyses for identifying crystal structures using powder X-ray diffraction. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Steps for synthesizing the strontium and barium ethylamine chloride perovskites.</p>
Full article ">Figure 2
<p>Crystal structure visualization by EXPO2014 of barium perovskite (CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>BaCl<sub>3</sub>) at different temperatures: (<b>a</b>) at 90 °C, (<b>b</b>) at 100 °C, and (<b>c</b>) at 110 °C, with an orthorhombic structure. The structural models shown were drawn with VESTA software (Ver. 3.5.8).</p>
Full article ">Figure 3
<p>Crystal structure visualization by EXPO2014 of strontium perovskite (CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>SrCl<sub>3</sub>) at different temperatures: (<b>a</b>) at 90 °C, (<b>b</b>) at 100 °C, and (<b>c</b>) at 110 °C, with a tetragonal structure. The structural models shown were drawn with VESTA software.</p>
Full article ">Figure 4
<p>Experimental X-ray powder diffraction pattern of barium perovskite with an orthorhombic structure at 90, 100, and 110 °C.</p>
Full article ">Figure 5
<p>Experimental X-ray powder diffraction pattern of strontium perovskite with a tetragonal structure at 90, 100, and 110 °C.</p>
Full article ">Figure 6
<p>FTIR spectra for the barium perovskites (CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>BaCl<sub>3</sub>) at different temperatures.</p>
Full article ">Figure 7
<p>FTIR spectra for the strontium perovskites (CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>SrCl<sub>3</sub>) at different temperatures.</p>
Full article ">Figure 8
<p>Absorption spectra of CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>BaCl<sub>3</sub> at 90, 100, and 110 °C.</p>
Full article ">Figure 9
<p>Absorption spectra of CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>SrCl<sub>3</sub> at 90, 100, and 110 °C.</p>
Full article ">Figure 10
<p>PL emission spectra of CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>BaCl<sub>3</sub> at 90, 100, and 110 °C.</p>
Full article ">Figure 11
<p>PL emission spectra of CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>SrCl<sub>3</sub> at 90, 100, and 110 °C.</p>
Full article ">Figure 12
<p>Normalized UV–Visible spectra of CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>BaCl<sub>3</sub> perovskite at 90, 100, and 110 °C. Inset: corresponding Tauc plots.</p>
Full article ">Figure 13
<p>Normalized UV–Visible spectra of CH<sub>3</sub>CH<sub>2</sub>NH<sub>3</sub>SrCl<sub>3</sub> perovskite at 90, 100, and 110 °C. Inset: corresponding Tauc plots.</p>
Full article ">Figure 14
<p>Band gap structure and energy levels of barium perovskite.</p>
Full article ">Figure 15
<p>Band gap structure and energy levels of strontium perovskite.</p>
Full article ">
18 pages, 2363 KiB  
Article
Mixed Pt-Ni Halide Perovskites for Photovoltaic Application
by Huilong Liu, Rubaiya Murshed and Shubhra Bansal
Materials 2024, 17(24), 6196; https://doi.org/10.3390/ma17246196 - 18 Dec 2024
Viewed by 360
Abstract
Cs2PtI6 is a promising photoabsorber with a direct bandgap of 1.4 eV and a high carrier lifetime; however, the cost of Pt inhibits its commercial viability. Here, we performed a cost analysis and experimentally explored the effect of replacing Pt [...] Read more.
Cs2PtI6 is a promising photoabsorber with a direct bandgap of 1.4 eV and a high carrier lifetime; however, the cost of Pt inhibits its commercial viability. Here, we performed a cost analysis and experimentally explored the effect of replacing Pt with earth-abundant Ni in solution-processed Cs(PtxNi1−x)(I,Cl)3 thin films on the properties and stability of the perovskite material. Films fabricated with CsI and PtI2 precursors result in a perovskite phase with a bandgap of 2.13 eV which transitions into stable Cs2PtI6 with a bandgap of 1.6 eV upon annealing. The complete substitution of PtI2 in films with CsI + NiCl2 precursors results in a wider bandgap of 2.35 eV and SEM shows two phases—a rod-like structure identified as CsNi(I,Cl)3 and residual white particles of CsI, also confirmed by XRD and Raman spectra. Upon extended thermal annealing, the bandgap reduces to 1.65 eV and transforms to CsNiCl3 with a peak shift to higher 2-theta. The partial substitution of PtI2 with NiCl2 in mixed 50-50 Pt-Ni-based films produces a bandgap of 1.9 eV, exhibiting a phase of Cs(Pt,Ni)(I,Cl)3 composition. A similar bandgap of 1.85 eV and the same diffraction pattern with improved crystallinity is observed after 100 h of annealing, confirming the formation of a stable mixed Pt-Ni phase. Full article
(This article belongs to the Special Issue Advanced Energy Materials for Perovskite Solar Cells)
Show Figures

Figure 1

Figure 1
<p>Atmospheric synthesis of PtI<sub>2</sub>, mixed PtI<sub>2</sub>-NiCl<sub>2</sub>, and NiCl<sub>2</sub>-based films in 50:50 DMF: DMSO via solution processing.</p>
Full article ">Figure 2
<p>USD/Watt (solute) of various Pb and Pb-free perovskite compositions calculated with respect to the PCE and thickness reported in the corresponding literature (blue) and the highest PCE of 25.6% and thickness of 2000 nm reported for the Pb-based FAPbI<sub>3</sub> perovskite (red). <a href="#app1-materials-17-06196" class="html-app">Figure S1</a> represents the USD/watt with the discrete effect of optimized PCE and absorber layer thickness.</p>
Full article ">Figure 3
<p>USD/Watt (solute + encapsulant) of various Pb and Pb-free perovskite compounds calculated with respect to the highest PCE of 25.6% and thickness of 2000 nm reported for the Pb-based FAPbI<sub>3</sub> perovskite. E1, E2, E3, and E4 represent different encapsulants: Polyolefin, Teflon, PET, and EVA, respectively. <a href="#app1-materials-17-06196" class="html-app">Figure S2</a> represents the USD/Watt (solute + encapsulant) calculated with respect to the PCE and absorber layer thickness reported in the corresponding literature. <a href="#app1-materials-17-06196" class="html-app">Figure S3</a> represents the USD/watt (solute + encapsulant) with the discrete effect of optimized PCE reported for the Pb-based FAPbI<sub>3</sub> perovskite and the corresponding absorber layer thickness from the literature. <a href="#app1-materials-17-06196" class="html-app">Figure S4</a> represents the USD/watt (solute + encapsulant) with the discrete effect of the optimized absorber layer thickness reported for the Pb-based FAPbI<sub>3</sub> perovskite and PCE from the literature.</p>
Full article ">Figure 4
<p>(<b>a</b>) Absorption spectrums of 2 h annealed (at −15 in Hg and 100 °C) PtI<sub>2</sub>, mixed PtI<sub>2</sub>-NiCl<sub>2</sub>, and NiCl<sub>2</sub>-based films; (<b>b</b>) Tauc plot showing the optical bandgap of the 2 h annealed (at −15 in Hg and 100 °C) PtI<sub>2</sub>, mixed PtI<sub>2</sub>-NiCl<sub>2</sub>, and NiCl<sub>2</sub>-based films; (<b>c</b>) XRD spectra of the 2 h annealed (at −15 in Hg and 100 °C) PtI<sub>2</sub>, mixed PtI<sub>2</sub>-NiCl<sub>2</sub>, and NiCl<sub>2</sub>-based films; SEM images of (<b>d</b>) PtI<sub>2</sub>, (<b>e</b>) mixed PtI<sub>2</sub>-NiCl<sub>2</sub>, and (<b>f</b>) NiCl<sub>2</sub>-based films; Raman spectra of (<b>g</b>) PtI<sub>2</sub>-based and (<b>h</b>) NiCl<sub>2</sub>-based films, respectively; (<b>i</b>) Goldschmidt and (<b>j</b>) Bartel tolerance factors for Cs(Pt,Ni)(Cl,I)<sub>3</sub>.</p>
Full article ">Figure 5
<p>PtI<sub>2</sub>-based films before and after the dark thermal annealing test with t representing the annealing duration: (<b>a</b>) absorption coefficient; (<b>b</b>) Tauc plot; (<b>c</b>) XRD pattern; (<b>d</b>) cross-section SEM images before annealing; (<b>e</b>) cross-section SEM images after annealing; and (<b>f</b>) EDS analysis showing the atomic % of the elemental distribution.</p>
Full article ">Figure 6
<p>Mixed PtI<sub>2</sub>-NiCl<sub>2</sub>-based films before and after the dark thermal annealing test with t representing the annealing duration: (<b>a</b>) absorption spectrum; (<b>b</b>) Tauc plot; (<b>c</b>) XRD pattern; (<b>d</b>) cross-section SEM image before annealing; (<b>e</b>) cross-section SEM image after annealing; and (<b>f</b>) EDS analysis showing the atomic % of the elemental distribution.</p>
Full article ">Figure 7
<p>NiCl<sub>2</sub>-based films before and after the dark thermal annealing test with t representing the annealing duration: (<b>a</b>) absorption spectrum; (<b>b</b>) Tauc plot; (<b>c</b>) XRD pattern; (<b>d</b>) SEM morphology before annealing; (<b>e</b>) SEM morphology after annealing; and (<b>f</b>) EDS analysis showing the atomic % of the elemental distribution.</p>
Full article ">
28 pages, 7320 KiB  
Review
Recent Advances in Lead-Free All-Inorganic Perovskite CsCdCl3 Crystals for Anti-Counterfeiting Applications
by Nankai Wang, Zhaojie Zhu, Jianfu Li, Chaoyang Tu, Weidong Chen and Yan Wang
Crystals 2024, 14(12), 1077; https://doi.org/10.3390/cryst14121077 - 13 Dec 2024
Viewed by 537
Abstract
This study reviews the advanced anti-counterfeiting applications of CsCdCl3, a lead-free all-inorganic perovskite crystal exhibiting dynamic luminescent properties responsive to temperature and UV light. Using synthesis methods such as Bridgman and hydrothermal techniques and incorporating dopants like bromine and tellurium, this [...] Read more.
This study reviews the advanced anti-counterfeiting applications of CsCdCl3, a lead-free all-inorganic perovskite crystal exhibiting dynamic luminescent properties responsive to temperature and UV light. Using synthesis methods such as Bridgman and hydrothermal techniques and incorporating dopants like bromine and tellurium, this research achieves improved luminescent stability, spectral diversity, and afterglow characteristics in CsCdCl3. The crystal demonstrates extended afterglow, photochromic shifts, and temperature-sensitive luminescence, enabling applications in 4D encoding for secure data encryption and in cold-chain temperature monitoring for pharmaceuticals. Despite these promising attributes, the challenges related to photostability, batch consistency, and environmental resilience persist, necessitating further exploration into the optimized synthesis and doping strategies to enhance material stability. These findings underscore the potential of CsCdCl3 for high-security information storage, pharmaceutical anti-counterfeiting, and real-time environmental sensing, positioning it as a valuable material for the next generation of secure, intelligent packaging solutions. Full article
(This article belongs to the Special Issue Recent Development and Research Trend of Laser Crystals)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Crystal structure of CsCdCl<sub>3</sub> in the [001] direction. The blue unit is [Cd<sub>2</sub>Cl<sub>9</sub>]<sup>5−</sup> dimers and the yellow unit is [CdCl<sub>6</sub>]<sup>4−</sup> octahedron. (<b>b</b>) Bonding diagram of an octahedron to its surrounding units. (<b>c</b>) Bonding diagram of a dimer to its surrounding units [<a href="#B28-crystals-14-01077" class="html-bibr">28</a>].</p>
Full article ">Figure 2
<p>Several crystals grown in our lab: (<b>a</b>) CsCdCl<sub>3</sub>: Tb crystal. (<b>b</b>) CsCdCl<sub>3</sub>: Cu crystal. (<b>c</b>) CsCdCl<sub>3</sub>: Eu crystal. (<b>d</b>) CsCdCl<sub>3</sub>: Pr crystal.</p>
Full article ">Figure 3
<p>(<b>a</b>) Crystal structure of CsCdCl<sub>3</sub>. (<b>b</b>) XRD patterns showing CsCdCl<sub>3</sub> doped with various Br<sup>−</sup> concentrations. (<b>c</b>) Enlarged view of XRD patterns for CsCdCl<sub>3</sub> with different levels of Br<sup>−</sup> doping. (<b>d</b>) Fluorescence emission spectra for CsCdCl<sub>3</sub> crystals with increasing Br<sup>−</sup> doping concentrations. (<b>e</b>) Fluorescence decay curve at 480 nm under 254 nm UV excitation, analyzed for CsCdCl<sub>3</sub> with differing Br<sup>−</sup> amounts. (<b>f</b>) Fluorescence decay curve at 580 nm under 254 nm UV excitation for Br<sup>−</sup>-doped CsCdCl<sub>3</sub>. (<b>g</b>) In situ fluorescence spectra for CsCdCl<sub>3</sub>:12.5% Br recorded under 254 nm UV irradiation. (<b>h</b>) Time-resolved fluorescence intensity variations at 480 nm and 580 nm for CsCdCl<sub>3</sub>:12.5% Br, with images showing fluorescence changes over time [<a href="#B16-crystals-14-01077" class="html-bibr">16</a>].</p>
Full article ">Figure 4
<p>Trap defect engineering and persistent luminescence in CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup>/0.1% Sb<sup>3+</sup> crystals. (<b>a</b>) Depiction of the hexagonal CsCdCl<sub>3</sub> crystal structure and its modification via Mn<sup>2+</sup> and Sb<sup>3+</sup> doping. (<b>b</b>) XRD patterns for pure CsCdCl<sub>3</sub>, CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup>, and CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup>/0.1% Sb<sup>3+</sup>. (<b>c</b>) Trap distribution profiles of CsCdCl<sub>3</sub>, CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup>, and CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup>/0.1% Sb<sup>3+</sup>, shown through thermoluminescence glow curves measured at a heating rate of 60 K/min after X-ray irradiation. (<b>d</b>) Comparative radioluminescence afterglow decay curves for CsCdCl<sub>3</sub>, CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup>, and CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup>/0.1% Sb<sup>3+</sup> crystals at room temperature, following 3 min of X-ray excitation at 50 kV. (<b>e</b>) Sequential photographs of CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup> and CsCdCl<sub>3</sub>:5% Mn<sup>2+</sup>/0.1% Sb<sup>3+</sup> showing afterglow emission at various intervals after cessation of X-ray exposure at room temperature [<a href="#B17-crystals-14-01077" class="html-bibr">17</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) PLE and PL spectra of CsCdCl<sub>3</sub> samples. (<b>b</b>) PL decay curves monitored at a wavelength of 580 nm for CsCdCl<sub>3</sub> (λex = 254 nm). (<b>c</b>) PL spectra of CsCdCl<sub>3</sub>:x Te<sup>4+</sup> (x = 0, 1, 5, 10, 15, and 20%) excited by 430 nm. The inset shows the PL intensity as a function of Te<sup>4+</sup> concentration. (<b>d</b>) PLE and PL spectra of CsCdCl<sub>3</sub>:10% Te<sup>4+</sup> samples. (<b>e</b>) PL decay curves monitored at a wavelength of 580 nm for CsCdCl<sub>3</sub>:10% Te<sup>4+</sup> (λex = 254 nm/λex = 430 nm) samples. (<b>f</b>) Luminescence mechanism of CsCdCl<sub>3</sub>:10% Te<sup>4+</sup> [<a href="#B19-crystals-14-01077" class="html-bibr">19</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Photographs of CsCdCl<sub>3</sub> and CsCdCl<sub>3</sub>:10% Te<sup>4+</sup> crystals under UV light at different temperatures. (<b>b</b>) Pseudocolor temperature-dependent PL maps for CsCdCl<sub>3</sub> under 254 nm excitation. (<b>c</b>) Pseudocolor temperature-dependent PL maps for CsCdCl<sub>3</sub>:10% Te<sup>4+</sup> under 254 nm excitation. (<b>d</b>) Normalized PL spectra of CsCdCl<sub>3</sub> at three representative temperatures. (<b>e</b>) Normalized PL spectra of CsCdCl<sub>3</sub>:10% Te<sup>4+</sup> at three selected temperatures. (<b>f</b>) PL intensity of CsCdCl<sub>3</sub>:10% Te<sup>4+</sup> at different temperatures. (<b>g</b>) Fitted fwhm curve of CsCdCl<sub>3</sub>:10% Te<sup>4+</sup> as a function of temperature [<a href="#B19-crystals-14-01077" class="html-bibr">19</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic of the custom-built optical setup. (<b>b</b>) CIE chromaticity coordinates and corresponding images of the CsCdCl<sub>3</sub>:0.5% Sb<sup>3+</sup> SC sample under 350 and 400 nm fs laser excitation. (<b>c</b>) PL spectra recorded with 350 and 400 nm fs laser excitation. (<b>d</b>) PL spectra obtained at various pump densities under 350 and 400 nm fs laser excitation. (<b>e</b>) Emission intensity as a function of pump density under 350 nm fs laser excitation. (<b>f</b>) Emission intensity as a function of pump density under 400 nm fs laser excitation. (<b>g</b>) PersL spectra recorded after pre-irradiation with 350 and 400 nm fs lasers for 1 s. (<b>h</b>) PersL decay curves with corresponding photographs recorded post-pre-irradiation with 350 and 400 nm fs laser excitation. (fs laser settings: 1 kHz, 150 fs; spot size: 200 µm) [<a href="#B15-crystals-14-01077" class="html-bibr">15</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) Photographs of labels prepared byCsCdCl3:x%Br and CsCdCl3:x%Sn sample powders under 254 nm irradiation, along with corresponding afterglow emission after ceasing excitation at (<b>b</b>) 1 s, (<b>c</b>) 3 s and (<b>d</b>) 120 s. The symbol × in the figure represents an incorrect result, while the symbol √ represents the correct result after decryption [<a href="#B7-crystals-14-01077" class="html-bibr">7</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) QR code anti-counterfeiting design with dynamic fluorescence. (<b>b</b>) Dynamic anti-counterfeiting pattern produced by CsCdCl<sub>2.92</sub>Br<sub>0.08</sub> encapsulated in PDMS. (<b>c</b>) Anti-counterfeiting pattern design based on the long afterglow properties of CsCdCl<sub>2.92</sub>Br<sub>0.08</sub>. (<b>d</b>) Design of dynamic 4D code [<a href="#B16-crystals-14-01077" class="html-bibr">16</a>].</p>
Full article ">Figure 10
<p>Photographs of phosphors CsCdCl₃:0.6%Sb<sup>3+</sup>, 1%Mn<sup>2+</sup>, arranged in a “QR code” pattern, show cyan and orange colors under 254 nm and 365 nm UV irradiation, respectively [<a href="#B41-crystals-14-01077" class="html-bibr">41</a>].</p>
Full article ">
34 pages, 6710 KiB  
Review
Prospects of Halide Perovskites for Solar-to-Hydrogen Production
by Huilong Liu, Tulja Bhavani Korukonda and Shubhra Bansal
Nanomaterials 2024, 14(23), 1914; https://doi.org/10.3390/nano14231914 - 28 Nov 2024
Viewed by 896
Abstract
Solar-driven hydrogen generation is one of the promising technologies developed to address the world’s growing energy demand in an sustainable way. While, for hydrogen generation (otherwise water splitting), photocatalytic, photoelectrochemical, and PV-integrated water splitting systems employing conventional semiconductor oxides materials and their electrodes [...] Read more.
Solar-driven hydrogen generation is one of the promising technologies developed to address the world’s growing energy demand in an sustainable way. While, for hydrogen generation (otherwise water splitting), photocatalytic, photoelectrochemical, and PV-integrated water splitting systems employing conventional semiconductor oxides materials and their electrodes have been under investigation for over a decade, lead (Pb)- halide perovskites (HPs) made their debut in 2016. Since then, the exceptional characteristics of these materials, such as their tunable optoelectronic properties, ease of processing, high absorption coefficients, and long diffusion lengths, have positioned them as a highly promising material for solar-driven water splitting. Like in solar photovoltaics, a solar-driven water splitting field is also dominated by Pb-HPs with ongoing efforts to improve material stability and hydrogen evolution/generation rate (HER). Despite this, with the unveiling potential of various Pb-free HP compositions in photovoltaics and optoelectronics researchers were inspired to explore the potential of these materials in water splitting. In this current review, we outlined the fundamentals of water splitting, provided a summary of Pb HPs in this field, and the associated issues are presented. Subsequently, Pb-free HP compositions and strategies employed for improving the photocatalytic and/or electrochemical activity of the material are discussed in detail. Finally, this review presents existing issues and the future potential of lead-free HPs, which show potential for enhancing productivity of solar-to-hydrogen conversion technologies. Full article
(This article belongs to the Special Issue Advances in Nanomaterials for Photocatalysis)
Show Figures

Figure 1

Figure 1
<p>HP-based water splitting process and systems. (<b>a</b>) The process of photocatalytic water splitting employing two cocatalysts, A and B-cocatalyst. A is responsible for the H<sub>2</sub>O oxidation into O<sub>2</sub> and cocatalyst B is responsible for the hydrogen evolution from the reduction of the H<sup>+</sup> in the oxidation process. (<b>b</b>) Photocatalysis system, showing the dispersed photocatalyst in an aqueous medium. (<b>c</b>) Photo-electrocatalytic system, showing the perovskite-based photoelectrode, counter and reference electrodes immersed in the aqueous medium. (<b>d</b>) Photovoltaic-electrocatalytic system, showing the (perovskite-based) solar cell electrically connected to aqueous medium through the electrodes immersed in the medium.</p>
Full article ">Figure 2
<p>Structure of different dimensional perovskites: (<b>a</b>) 3-D; (<b>b</b>) 2-D; (<b>c</b>) 1-D; and (<b>d</b>) 0-D.</p>
Full article ">Figure 3
<p>Photocatalytic Water Splitting using Pb-HPs: (<b>a</b>) Band alignment of MAPbI<sub>3</sub> with respect to NHE. Reproduced with permission: Copyright 2016, Springer Nature Limited [<a href="#B6-nanomaterials-14-01914" class="html-bibr">6</a>]. (<b>b</b>) (<b>i</b>). Bandgap funnel structure of MAPbBr<sub>3−<span class="html-italic">x</span></sub>I<span class="html-italic"><sub>x</sub></span>, with the bandgap narrowing to the surface, with Pt co-catalyst loaded on the surface of the perovskite. (<b>ii</b>). Cross-sectional SEM image MAPbBr<sub>3−<span class="html-italic">x</span></sub>I<span class="html-italic"><sub>x</sub></span> particle with the corresponding EDX mapping along the yellow line profile with I-element distribution profile. Reproduced with permission: Copyright 2018, American Chemical Society [<a href="#B21-nanomaterials-14-01914" class="html-bibr">21</a>]. (<b>c</b>) (<b>i</b>). Band diagram of the CsPbBr<sub>3</sub>, CsPbBr<sub>3−<span class="html-italic">x</span></sub>I<span class="html-italic"><sub>x</sub></span> and CsPbI<sub>3</sub>. (<b>ii</b>). Photographic images of CsPbBr<sub>3−<span class="html-italic">x</span></sub>I<span class="html-italic"><sub>x</sub></span> powders with different halide exchange reaction times. Reproduced with permission: Copyright 2019, Elsevier [<a href="#B22-nanomaterials-14-01914" class="html-bibr">22</a>]. (<b>d</b>) SEM image of MAPbI<sub>3</sub>/rGO (inset-Schematic illustration MAPbI<sub>3</sub>/rGO composite photocatalyst. Reproduced with permission: 2018 WILEY-VCH [<a href="#B23-nanomaterials-14-01914" class="html-bibr">23</a>].</p>
Full article ">Figure 4
<p>Toxicity of Pb-HP devices: (<b>a</b>) Illustration of possible leaching of lead into ground water. (<b>b</b>) The photographs of mint plants grown on control soil (<b>left</b>) and 250 mg kg<sup>−1</sup> Pb<sup>2+</sup> perovskite-contaminated soil (<b>right</b>). Reproduced with permission: Copyright 2020, Springer Nature [<a href="#B34-nanomaterials-14-01914" class="html-bibr">34</a>]. (<b>c</b>) Assessment of the lead contamination on the human Lead Weekly intake level, considering the data for the world population and the total PSC lead necessary for electricity generation in 2050 vs. the adult Lead Weekly intake limit in 2010 and 3000 to 5000 years ago. Reproduced with permission: Copyright 2022, Elsevier Ltd. [<a href="#B35-nanomaterials-14-01914" class="html-bibr">35</a>].</p>
Full article ">Figure 5
<p>Stability of CH<sub>3</sub>NH<sub>3</sub>PbI<sub>3</sub> Perovskite: (<b>a</b>) Decomposition of path of the perovskite in presence of water molecules A water molecule H<sub>2</sub>O is required to initiate the process with the decomposition, leading to the formation of hydrated MAPI and eventually to Pb and PbI<sub>2</sub> with the gases HI and CH<sub>3</sub>NH<sub>2</sub> leaving the substrate. Reproduced with permission: Copyright 2014, American Chemical Society [<a href="#B42-nanomaterials-14-01914" class="html-bibr">42</a>]. (<b>b</b>) absorption spectra for the perovskite films on glass before (black) and after (red) aging in different conditions. The inset shows the photographs of films before (<b>left</b>) and after (<b>right</b>) aging. Reproduced with permission: Copyright 2016, The Royal Society of Chemistry [<a href="#B46-nanomaterials-14-01914" class="html-bibr">46</a>]. (<b>c</b>) X-ray diffractograms of the perovskite films degraded in different atmospheres for 24 h. (The inset shows the photographs of films). Reproduced with permission: Copyright 2016, The Royal Society of Chemistry [<a href="#B47-nanomaterials-14-01914" class="html-bibr">47</a>]. (<b>d</b>) SEM pictures of perovskite on ITO: pristine layers, initial stages of degradation, and later stages of degradation. The scale bars represent 200 nm followed by the illustration of the progression of oxygen induced degradation from the grain boundaries. Reproduced with permission: Copyright 2017, WILEY-VCH [<a href="#B48-nanomaterials-14-01914" class="html-bibr">48</a>].</p>
Full article ">Figure 6
<p>Bandgaps of various lead-free HP compositions discussed in this section.</p>
Full article ">Figure 7
<p>Compositional Engineering: (<b>a</b>) Schematic of Interfacial interaction of MA<sub>3</sub>Bi<sub>2</sub>Cl<sub>9−<span class="html-italic">x</span></sub>I<span class="html-italic"><sub>x</sub></span> with the bandgap funnel structure. Reproduced with permission: Copyright 2022, Wiley-VCH [<a href="#B88-nanomaterials-14-01914" class="html-bibr">88</a>]. (<b>b</b>) Absorption spectra of the as-prepared Cs<sub>3</sub>Bi<sub>2<span class="html-italic">x</span></sub>Sb<sub>2−2<span class="html-italic">x</span></sub>I<sub>9</sub> (<span class="html-italic">x</span> = 0.9, 0.7, 0.5, 0.3, 0.1). Reproduced with permission: Copyright 2020, Wiley-VCH [<a href="#B89-nanomaterials-14-01914" class="html-bibr">89</a>]. (<b>c</b>) (<b>i</b>). Crystal structure and transformation relationship of Cs<sub>2</sub>Pt<span class="html-italic"><sub>x</sub></span>Sn<sub>1−<span class="html-italic">x</span></sub>Cl<sub>6</sub>. (<b>ii</b>). Charge-carrier dynamics model of Cs<sub>2</sub>Pt<sub>0.05</sub>Sn<sub>0.95</sub>Cl<sub>6</sub> and Cs<sub>2</sub>Pt<sub>0.75</sub>Sn<sub>0.25</sub>Cl<sub>6</sub>. Reproduced with permission: Copyright 2021, Wiley-VCH [<a href="#B90-nanomaterials-14-01914" class="html-bibr">90</a>].</p>
Full article ">Figure 8
<p>Semiconductor/Pb-free HP heterojunctions: (<b>a</b>) Photocurrent response of BiVO<sub>4</sub> and BiVO<sub>4</sub>/Cs<sub>2</sub>PtI<sub>6</sub> heterojunction (Inset-Band alignment of the photoanode as a function of the wavelength). Reproduced with permission: Copyright 2021, American Chemical Society [<a href="#B98-nanomaterials-14-01914" class="html-bibr">98</a>]. (<b>b</b>) Schematic representation of the formation of MA<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/DMA<sub>3</sub>BiI<sub>6</sub> heterojunctions. (I) MA<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> in the pre-reaction solution prior to the solvothermal process. (II) The DMA<sup>+</sup> ions generated through DMF hydrolysis reacted with the unreacted bismuth ions on the MA<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> perovskites. (III) In-situ formation of MA<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/DMA<sub>3</sub>BiI<sub>6</sub> perovskite heterojunctions in BBP-5. Reproduced with permission: Copyright 2020, Wiley-VCH [<a href="#B99-nanomaterials-14-01914" class="html-bibr">99</a>]. (<b>c</b>) Schematic representation of synthesis of Cs<sub>2</sub>AgBiBr<sub>6</sub>/N-C Photocatalyst. (<b>i</b>) Conventional one-pot synthesis is employed for the synthesis of the Cs<sub>2</sub>AgBiBr<sub>6</sub>/N-C). (<b>ii</b>) The band alignment of the various heterojunction photocatalysts.Reproduced with permission: Copyright 2021, American Chemical Society [<a href="#B100-nanomaterials-14-01914" class="html-bibr">100</a>]. (<b>d</b>) (<b>i</b>). Schematic representation of Cs<sub>3</sub>Rh<sub>2</sub>I<sub>9</sub> bulk crystal. (<b>ii</b>). Electrochemical reduction of Cs<sub>3</sub>Rh<sub>2</sub>I<sub>9</sub> clusters on NC (Cs<sub>3</sub>Rh<sub>2</sub>I<sub>9</sub>/NC) to form Cs<sub>3</sub>Rh<sub>2</sub>I<sub>9</sub>/NC-R. (<b>iii</b>). Electrochemical reduction of bulk Cs<sub>3</sub>Rh<sub>2</sub>I<sub>9</sub> to form Cs<sub>3</sub>Rh<sub>2</sub>I<sub>9</sub>-R with large particle size. Reproduced with permission: Copyright 2023, Springer Nature [<a href="#B101-nanomaterials-14-01914" class="html-bibr">101</a>].</p>
Full article ">Figure 9
<p>g-C<sub>3</sub>N<sub>4</sub>/Pb-free HP Heterojunctions: (<b>a</b>) Band alignment of the PEA<sub>2</sub>SnBr<sub>4</sub> and g-C<sub>3</sub>H<sub>4</sub> relative to NHE potential. Reproduced with permission: Copyright 2020, Royal Society of Chemistry [<a href="#B80-nanomaterials-14-01914" class="html-bibr">80</a>]. (<b>b</b>) Band alignment of the Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> and g-C<sub>3</sub>H<sub>4</sub> relative to NHE potential. Reproduced with permission: Copyright 2020, Elsevier B.V. [<a href="#B102-nanomaterials-14-01914" class="html-bibr">102</a>]. (<b>c</b>) Schematic representation of H<sub>2</sub> evolution mechanism by Cs<sub>2</sub>AgBiBr<sub>6</sub>-rGO under visible light irradiation (inset: SEM image of the photocatalyst). Reproduced with permission: Copyright 2019, Elsevier B.V. [<a href="#B103-nanomaterials-14-01914" class="html-bibr">103</a>]. (<b>d</b>) Band alignment of g-C<sub>3</sub>N<sub>4</sub> and DMASnBr<sub>3</sub> aligned with respect to water, H<sup>+</sup>/H<sub>2</sub> and TEOA/TEOA<sup>+</sup> redox level. Reproduced with permission: Copyright 2020, Wiley-VCH [<a href="#B79-nanomaterials-14-01914" class="html-bibr">79</a>]. (<b>e</b>) Band alignment of the Cs<sub>2</sub>AgBiBr<sub>6</sub> and g-C<sub>3</sub>H<sub>4</sub> relative to NHE potential. Reproduced with permission: Copyright 2020, Springer Nature [<a href="#B105-nanomaterials-14-01914" class="html-bibr">105</a>].</p>
Full article ">
25 pages, 6320 KiB  
Article
Tunable Optical Properties and Relaxor Behavior in Ni/Ba Co-Doped NaNbO3 Ceramics: Pathways Toward Multifunctional Applications
by Tawfik Chaabeni, Zohra Benzarti, Najmeddine Abdelmoula and Slim Zghal
Ceramics 2024, 7(4), 1670-1694; https://doi.org/10.3390/ceramics7040107 - 8 Nov 2024
Viewed by 824
Abstract
In this study, Ni/Ba co-doped NaNbO3 ceramics (NBNNOx) were synthesized using a solid-state method to explore the effects of Ni2+ and Ba2+ ion substitution on the structural, optical, and dielectric properties of NaNbO3. X-ray diffraction (XRD) [...] Read more.
In this study, Ni/Ba co-doped NaNbO3 ceramics (NBNNOx) were synthesized using a solid-state method to explore the effects of Ni2+ and Ba2+ ion substitution on the structural, optical, and dielectric properties of NaNbO3. X-ray diffraction (XRD) confirmed that the ceramics retained an orthorhombic structure, with crystallinity improving as the doping content (x) increased. Significant lattice distortions induced by the Ni/Ba co-doping were observed, which were essential for preserving the perovskite structure. Raman spectroscopy revealed local structural distortions, influencing optical properties and promoting relaxor behavior. Diffuse reflectance measurements revealed a significant decrease in band gap energy from 3.34 eV for undoped NaNbO3 to 1.08 eV at x = 0.15, highlighting the impact of co-doping on band gap tunability. Dielectric measurements indicated relaxor-like behavior at room temperature for x = 0.15, characterized by frequency-dependent anomalies in permittivity and dielectric loss, likely due to ionic disorder and structural distortions. These findings demonstrate the potential of Ni/Ba co-doped NaNbO3 ceramics for lead-free perovskite solar cells and other functional devices, where tunable optical and dielectric properties are highly desirable. Full article
(This article belongs to the Special Issue Advances in Electronic Ceramics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) X-ray diffraction patterns of sintered NBNNO<sub>x</sub> ceramics recorded at room temperature; (<b>b</b>) Enlarged view of the XRD patterns in the rectangular window between 31° and 33°, focusing on the principal diffraction peaks.</p>
Full article ">Figure 2
<p>The Rietveld pattern obtained from X-ray powder diffraction of the NBNNO<sub>x</sub> ceramics. (<b>a</b>) NNO; (<b>b</b>) NBNNO<sub>0.05</sub>; (<b>c</b>) NBNNO<sub>0.10</sub>; (<b>d</b>) NBNNO<sub>0.15</sub>. The points represent the experimental profile, and the red solid line corresponds to the calculated profile. Positions of the Bragg reflections are indicated by green vertical bars. The blue curve at the bottom of the figure represents the discrepancy between the experimental and calculated profiles. The crystal structures were plotted inside using the “Diamond” program.</p>
Full article ">Figure 3
<p>(<b>a</b>) Variation in the unit cell volume of the NBNNO<sub>x</sub> orthorhombic phases; (<b>b</b>) Lattice parameters a and b shown on the left-hand side ordinate, and c on the right-hand side ordinate as a function of the doping content (x).</p>
Full article ">Figure 4
<p>Arrangements of the BO<sub>6</sub> octahedra along the c-axis of the NBNNO<sub>x</sub> structures. <span class="html-italic">θ</span><sub>1</sub> and <span class="html-italic">θ</span><sub>2</sub> represent the periodically modulated tilt angles of the structural chains. The oxygen ions are represented by red spheres, and the Nb/Ni ions are represented by yellow spheres.</p>
Full article ">Figure 5
<p>(<b>a</b>) Variation in the tilt angles within the structural octahedral chain; (<b>b</b>) Variation in the axial, apical and average &lt;(Nb/Ni)–O&gt; bond lengths as a function of doping content (x).</p>
Full article ">Figure 6
<p>Raman spectra recorded at room temperature for NBNNO<sub>x</sub> ceramics with different Ni/Ba contents. (<b>a</b>) NNO; (<b>b</b>) NBNNO<sub>0.05</sub>; (<b>c</b>) NBNNO<sub>0.10</sub>; (<b>d</b>) NBNNO<sub>0.15</sub>. Various vibration modes are identified from the deconvoluted Raman spectra, represented in different colors.</p>
Full article ">Figure 7
<p>Schematic representations of internal vibrational modes of BO<sub>6</sub> octahedra, showing that the ϑ<sub>1</sub>, ϑ<sub>2</sub>, and ϑ<sub>5</sub> modes are symmetric; the ϑ<sub>3</sub>, ϑ<sub>4</sub>, and ϑ<sub>6</sub> modes are non-symmetric; <sup>1</sup>A<sub>1g</sub> (ϑ<sub>1</sub>), <sup>1</sup>E<sub>g</sub> (ϑ<sub>2</sub>), and <sup>1</sup>F<sub>1u</sub> (ϑ<sub>3</sub>) are stretching modes; and <sup>1</sup>F<sub>1u</sub> (ϑ<sub>4</sub>), <sup>1</sup>F<sub>2g</sub> (ϑ<sub>5</sub>), and <sup>1</sup>F<sub>2u</sub> (ϑ<sub>6</sub>) are bending modes. The blue arrows indicate the direction of atom vibrations. The oxygen ions are represented by red spheres, and the Nb/Ni ions are represented by yellow spheres.</p>
Full article ">Figure 8
<p>Variation in the Raman peak position of the principal internal vibration modes of the BO<sub>6</sub> octahedra, ϑ<sub>1</sub> (bending mode) and ϑ<sub>5</sub> (stretching mode), as a function of the doping content.</p>
Full article ">Figure 9
<p>Variation in the microstructural parameters: crystallite size (<span class="html-italic">D</span>) and microstrain (<span class="html-italic">ε</span>) as a function of the doping content (x), determined through X-ray diffraction peak broadening and derived from the Halder–Wagner approach.</p>
Full article ">Figure 10
<p>SEM micrographs using secondary electron imaging of (<b>a</b>) parent phase NNO; (<b>b</b>) doped NBNNO<sub>0.15</sub> ceramic.</p>
Full article ">Figure 11
<p>Diffuse reflectance spectra of NBNNO<sub>x</sub> ceramics measured at room temperature in the 350–900 nm range.</p>
Full article ">Figure 12
<p>Kubelka–Munk transformed reflectance spectra of NBNNO<sub>x</sub> ceramics at room temperature (solid lines), showing the determination of the band gap energy (E<sub>g</sub>) through the extrapolation of the linear portion of the spectra (dotted lines).</p>
Full article ">Figure 13
<p>Dependence of the band gap and Urbach energy on doping content. The Urbach energy (E<sub>u</sub>) is extracted from the slope of the linear fit in the plot of ln(<span class="html-italic">F(R)</span>) versus photon energy (hν).</p>
Full article ">Figure 14
<p>Temperature dependence of the real part of the dielectric permittivity (ε<sub>r</sub>′) and dielectric loss (δ) of the NBNNO<sub>x</sub> ceramics: (<b>a</b>) NNO; (<b>b</b>) NBNNO<sub>0.05</sub>; (<b>c</b>) NBNNO<sub>0.10</sub>; (<b>d</b>) NBNNO<sub>0.15</sub>. The dielectric loss (δ) curves are presented in the inset of each figure.</p>
Full article ">Figure 15
<p>Temperature dependence of the inverse dielectric constant for the doped NBNNO<sub>x</sub> ceramics at 1 kHz: (<b>a</b>) NBNNO<sub>0.05</sub>; (<b>b</b>) NBNNO<sub>0.10</sub>; (<b>c</b>) NBNNO<sub>0.15</sub>. Solid lines represent the experimental data, while dotted lines indicate the extrapolation of the linear fit to the Curie–Weiss law. <span class="html-italic">T<sub>m</sub></span> marks the temperature of the dielectric constant maximum, and <span class="html-italic">T<sub>dev</sub></span> denotes the temperature above which the dielectric constant follows the Curie–Weiss law. The deviation from the Curie–Weiss law is illustrated for the first dielectric transition.</p>
Full article ">Figure 16
<p>Plot of ln(<span class="html-italic">1/</span><math display="inline"><semantics> <mi>ε</mi> </semantics></math><span class="html-italic"><sub>r</sub>′</span> − <span class="html-italic">1/</span><math display="inline"><semantics> <mi>ε</mi> </semantics></math><span class="html-italic"><sub>r</sub>′<sub>m</sub></span>) versus ln(<span class="html-italic">T</span> − <span class="html-italic">T<sub>m</sub></span>) for the doped NBNNO<sub>x</sub> ceramics at 1 kHz: (<b>a</b>) NBNNO<sub>0.05</sub>; (<b>b</b>) NBNNO<sub>0.10</sub>; (<b>c</b>) NBNNO<sub>0.15</sub>. The filled squares represent the experimental data, while the dotted lines indicate the extrapolation of the linear fit to the modified Curie–Weiss law. The slope of the fitting curves reflects the degree of diffuseness, γ.</p>
Full article ">
13 pages, 3649 KiB  
Article
Enhancing the Photocatalytic Activity of Lead-Free Halide Perovskite Cs3Bi2I9 by Compositing with Ti3C2 MXene
by Tao Tang, Xiaoyu Dou, Haoran Zhang, Hexu Wang, Ming Li, Guanghui Hu, Jianfeng Wen and Li Jiang
Molecules 2024, 29(21), 5096; https://doi.org/10.3390/molecules29215096 - 28 Oct 2024
Viewed by 700
Abstract
In recent years, halide perovskite materials have become widely used in solar cells, photovoltaics, and LEDs, as well as photocatalysis. Lead-free perovskite Cs3Bi2I9 has been demonstrated as an effective photocatalyst; however, the fast recombination of the photogenerated carriers [...] Read more.
In recent years, halide perovskite materials have become widely used in solar cells, photovoltaics, and LEDs, as well as photocatalysis. Lead-free perovskite Cs3Bi2I9 has been demonstrated as an effective photocatalyst; however, the fast recombination of the photogenerated carriers hinders further improvements of its photocatalytic activity. In this work, Ti3C2 was composited with Cs3Bi2I9 to promote the transfer and separation of photogenerated carriers, and thus the pollutant degradation efficiency was effectively improved. The visible-light photocatalytic reduction of Cs3Bi2I9/Ti3C2 on rhodamine B (RhB), methylene blue (MB), and malachite green (MG) was as high as 97.3%, 96%, and 98.8%, respectively, improvements of almost 31.2%, 37.8%, and 37.2% compared to that of sole Cs3Bi2I9. Our study provides a simple way to enhance the photocatalytic activity of lead-free halide perovskites. Full article
(This article belongs to the Special Issue Advances in Composite Photocatalysts)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) SEM image of Ti<sub>3</sub>C<sub>2</sub>. (<b>b</b>) TEM and (<b>c</b>) high-resolution TEM images of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>. (<b>d</b>) TEM and (<b>e</b>) high-resolution TEM images of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub> composite. (<b>f</b>) Lattice streak diagram of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub> composite.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD and (<b>b</b>) XPS spectra of Ti<sub>3</sub>C<sub>2</sub>, Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>, and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>. (<b>c</b>) C 1s and (<b>e</b>) Ti 2p spectra of Ti<sub>3</sub>C<sub>2</sub> and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>. (<b>d</b>) Bi 4f and (<b>f</b>) I 3d spectra of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub> composites, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) Water contact angles of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>. (<b>b</b>) Time-dependent PL spectra of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub> aqueous solutions. The excitation wavelength is 390 nm. (<b>c</b>) UV-vis absorption spectra of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>. The inset is the Tauc plots of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>. (<b>d</b>) Photocurrent responses and (<b>e</b>) electrochemical impedances of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Effects of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>, Ti<sub>3</sub>C<sub>2</sub>, Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>-1, Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>-2, and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>-3 on the photocatalytic degradation of RhB under visible light. (<b>b</b>) The corresponding degradation kinetic behavior. (<b>c</b>) Photocatalytic cycle tests of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>. (<b>d</b>) Effects of scavengers on the photocatalytic activity of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Effects of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub> and Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub> on photocatalytic degradation of MB and MG under visible light. (<b>b</b>) Corresponding degradation kinetic behaviors.</p>
Full article ">Figure 6
<p>(<b>a</b>) The UPS measurements of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>, and inset are the onset and cutoff energies. (<b>b</b>) Schematic of photocatalytic mechanism of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub> under visible light.</p>
Full article ">Figure 7
<p>Schematic diagram of the synthesis process of Cs<sub>3</sub>Bi<sub>2</sub>I<sub>9</sub>/Ti<sub>3</sub>C<sub>2</sub> composite.</p>
Full article ">
11 pages, 3350 KiB  
Article
CsPbBr3 and Cs2AgBiBr6 Composite Thick Films with Potential Photodetector Applications
by Merida Sotelo-Lerma, Leunam Fernandez-Izquierdo, Martin A. Ruiz-Molina, Igor Borges-Doren, Ross Haroldson and Manuel Quevedo-Lopez
Materials 2024, 17(20), 5123; https://doi.org/10.3390/ma17205123 - 21 Oct 2024
Viewed by 1099
Abstract
This paper investigates the optoelectronic properties of CsPbBr3, a lead-based perovskite, and Cs2AgBiBr6, a lead-free double perovskite, in composite thick films synthesized using mechanochemical and hot press methods, with poly(butyl methacrylate) as the matrix. Comprehensive characterization was [...] Read more.
This paper investigates the optoelectronic properties of CsPbBr3, a lead-based perovskite, and Cs2AgBiBr6, a lead-free double perovskite, in composite thick films synthesized using mechanochemical and hot press methods, with poly(butyl methacrylate) as the matrix. Comprehensive characterization was conducted, including X-ray diffraction (XRD), Raman spectroscopy, scanning electron microscopy (SEM), UV–visible spectroscopy (UV–Vis), and photoluminescence (PL). Results indicate that the polymer matrix does not significantly impact the crystalline structure of the perovskites but has a direct impact on the grain size and surface area, enhancing the interfacial charge transfer of the composites. Optical characterization indicates minimal changes in bandgap energies across all different phases, with CsPbBr3 exhibiting higher photocurrent than Cs2AgBiBr6. This is attributed to the CsPbBr3 superior charge carrier mobility. Both composites showed photoconductive behavior, with Cs2AgBiBr6 also demonstrating higher-energy (X-ray) photon detection. These findings highlight the potential of both materials for advanced photodetector applications, with Cs2AgBiBr6 offering an environmentally Pb-free alternative. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flow chart of device preparation for halide perovskite composites. Ball milling was used for the mechanochemical synthesis of perovskite materials. The hot press was used to obtain the thick film using a composite (perovskite:polymer—75:25 wt%). The photoresponse under dark and illuminated conditions (photons of 365 nm) was used to characterize the device.</p>
Full article ">Figure 2
<p>XRD diffractograms of (<b>a</b>) CsPbBr<sub>3</sub> and (<b>c</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub>. Raman spectra with 780 nm laser source of (<b>b</b>) CsPbBr<sub>3</sub> and (<b>d</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub>.</p>
Full article ">Figure 3
<p>SEM images for (<b>a</b>) CsPbBr<sub>3</sub> (powder), (<b>b</b>) CsPbBr<sub>3</sub> (composite), (<b>d</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub> (powder), (<b>e</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub> (composite), and cross-section images for (<b>c</b>) CsPbBr<sub>3</sub> (film) and (<b>f</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub> (film).</p>
Full article ">Figure 4
<p>UV–Vis spectroscopy of (<b>a</b>) CsPbBr<sub>3</sub> and (<b>c</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub>. PL spectroscopy of (<b>b</b>) CsPbBr<sub>3</sub> and (<b>d</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub>.</p>
Full article ">Figure 5
<p>I–V curves of (<b>a</b>) CsPbBr<sub>3</sub> and (<b>b</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub>.</p>
Full article ">Figure 6
<p>Bode diagram of (<b>a</b>) CsPbBr<sub>3</sub> and (<b>b</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub>. Theta phase vs. frequency of (<b>c</b>) CsPbBr<sub>3</sub> and (<b>d</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub>. Nyquist plot of (<b>e</b>) CsPbBr<sub>3</sub> and (<b>f</b>) Cs<sub>2</sub>AgBiBr<sub>6</sub>.</p>
Full article ">
10 pages, 1806 KiB  
Article
Design and Simulation for Minimizing Non-Radiative Recombination Losses in CsGeI2Br Perovskite Solar Cells
by Tingxue Zhou, Xin Huang, Diao Zhang, Wei Liu and Xing’ao Li
Nanomaterials 2024, 14(20), 1650; https://doi.org/10.3390/nano14201650 - 14 Oct 2024
Viewed by 1042
Abstract
CsGeI2Br-based perovskites, with their favorable band gap and high absorption coefficient, are promising candidates for the development of efficient lead-free perovskite solar cells (PSCs). However, bulk and interfacial carrier non-radiative recombination losses hinder the further improvement of power conversion efficiency and [...] Read more.
CsGeI2Br-based perovskites, with their favorable band gap and high absorption coefficient, are promising candidates for the development of efficient lead-free perovskite solar cells (PSCs). However, bulk and interfacial carrier non-radiative recombination losses hinder the further improvement of power conversion efficiency and stability in PSCs. To overcome this challenge, the photovoltaic potential of the device is unlocked by optimizing the optical and electronic parameters through rigorous numerical simulation, which include tuning perovskite thickness, bulk defect density, and series and shunt resistance. Additionally, to make the simulation data as realistic as possible, recombination processes, such as Auger recombination, must be considered. In this simulation, when the Auger capture coefficient is increased to 10−29 cm6 s−1, the efficiency drops from 31.62% (without taking Auger recombination into account) to 29.10%. Since Auger recombination is unavoidable in experiments, carrier losses due to Auger recombination should be included in the analysis of the efficiency limit to avoid significantly overestimating the simulated device performance. Therefore, this paper provides valuable insights for designing realistic and efficient lead-free PSCs. Full article
(This article belongs to the Special Issue Perovskite Nanostructures: Synthesis, Properties and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of the perovskite device used for this simulation. (<b>b</b>) Energy−band alignment of ETL and HTL with CsGeI<sub>2</sub>Br perovskite and FTO.</p>
Full article ">Figure 2
<p>Effects of the various perovskite thicknesses on device performance: (<b>a</b>) <span class="html-italic">V<sub>oc</sub></span>, (<b>b</b>) FF, (<b>c</b>) PCE, (<b>d</b>) <span class="html-italic">V<sub>mpp</sub></span> and <span class="html-italic">J<sub>mpp</sub></span>. (<b>e</b>) Simulation of QFLS and <span class="html-italic">V<sub>oc</sub></span> of PSCs. (<b>f</b>) Energy−band diagram of the Fermi energy levels of electrons and holes that change with various perovskite thicknesses.</p>
Full article ">Figure 3
<p>(<b>a</b>) Plots of the current–voltage (J−V) curves with different N<sub>t</sub> and (<b>b</b>) the corresponding EQE curves. (<b>c</b>) The impedance of optimized PSCs with different N<sub>t</sub> (In the magnified image, magenta represents 10<sup>12</sup> cm<sup>−3</sup>, pink represents 10<sup>13</sup> cm<sup>−3</sup>, and orange represents 10<sup>14</sup> cm<sup>−3</sup>).</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>b</b>) Effect of the variation in R<sub>s</sub> on device performance and (<b>c</b>) the corresponding impedance. (<b>d</b>–<b>f</b>) Effect of the variation in R<sub>sh</sub> on device performance and (<b>c</b>) the corresponding impedance.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>d</b>) Effect of the variation Auger recombination coefficient on device performance.</p>
Full article ">
17 pages, 7553 KiB  
Article
Microwave-Assisted Fabrication and Characterization of Carbon Fiber-Sodium Bismuth Titanate Composites
by Fareeha Azam, Muhammad Asif Rafiq, Furqan Ahmed, Adnan Moqbool, Osama Fayyaz, Zerfishan Imran, Muhammad Salman Habib and Rana Abdul Shakoor
Crystals 2024, 14(9), 798; https://doi.org/10.3390/cryst14090798 - 10 Sep 2024
Viewed by 765
Abstract
Lead-based piezoelectric materials cause many environmental problems, regardless of their exceptional performance. To overcome this issue, a lead-free piezoelectric composite material was developed by incorporating different percentages of carbon fiber (CF) into the ceramic matrix of Bismuth Sodium Titanate (BNT) by employing the [...] Read more.
Lead-based piezoelectric materials cause many environmental problems, regardless of their exceptional performance. To overcome this issue, a lead-free piezoelectric composite material was developed by incorporating different percentages of carbon fiber (CF) into the ceramic matrix of Bismuth Sodium Titanate (BNT) by employing the microwave sintering technique. The aim of this study was also to evaluate the impact of microwave sintering on the microstructure and the electrical behavior of the carbon-fiber-reinforced Bi0.5Na0.5TiO3 composite (BNT-CF). A uniform distribution of the CF and increased densification of the BNT-CF was achieved, leading to improved piezoelectric properties. X-ray diffraction (XRD) showed the formation of a phase-pure crystalline perovskite structure consisting of CF and BNT. A Field Emission Scanning electron microscope (FESEM) revealed that utilizing microwave sintering at lower temperatures and shorter dwell times results in a superior densification of the BNT-CF. Raman Spectroscopy confirmed the perovskite structure of the BNT-CF and the presence of a Morphotropic Phase Boundary (MPB). An analysis of nanohardness indicated that the hardness of the BNT-CF increases with the increasing amount of CF. It is also revealed that the electrical conductivity of the BNT-CF at a low frequency is significantly influenced by the amount of CF and the temperature. Moreover, an increase in the carbon fiber concentration resulted in a decrease in dielectric properties. Finally, a lead-free piezoelectric BNT-CF showing dense and uniform microstructure was developed by the microwave sintering process. The promising properties of the BNT-CF make it attractive for many industrial applications. Full article
(This article belongs to the Special Issue Structural and Characterization of Composite Materials)
Show Figures

Figure 1

Figure 1
<p>A schematic representation of BNT–carbon fiber composite formation.</p>
Full article ">Figure 2
<p>(<b>a</b>) X-ray diffraction pattern of BNT-CFs microwave-sintered at 1000 °C for 20 min in 2θ ranges from 20 to 80°; (<b>b</b>) the magnifying image of splitting peaks in the range of 45–48°.</p>
Full article ">Figure 3
<p>(<b>a</b>) FE-SEM images of BNT powder synthesized using a ball mill; (<b>b</b>) particle-size analysis of BNT powder.</p>
Full article ">Figure 4
<p>FE-SEM micrographs of surfaces for (<b>a</b>) BNT, (<b>b</b>) BNT-0.5CF, (<b>c</b>) BNT-1CF, (<b>d</b>) BNT-2CF, and (<b>e</b>) BNT-5CF composites sintered at 1000 °C for 20 min with varying compositions of carbon fibers.</p>
Full article ">Figure 5
<p>(<b>a</b>) BNT, (<b>b</b>) BNT-0.5CF, (<b>c</b>) BNT-1CF, (<b>d</b>) BNT-2CF, (<b>e</b>) BNT-5CF, and (<b>f</b>) the reduction in grain size with increasing carbon fiber content.</p>
Full article ">Figure 6
<p>EDX analysis of (<b>a</b>) BNT and (<b>b</b>) BNT-2CF composites.</p>
Full article ">Figure 7
<p>Raman spectra for the synthesized compositions of BNT-CF at room temperature.</p>
Full article ">Figure 8
<p>The AFM images and roughness value of BNT-CF compositions: (<b>a</b>) BNT; (<b>b</b>) BNT-0.5CF; and (<b>c</b>) BNT-5CF.</p>
Full article ">Figure 9
<p>Nanohardness value of BNT-CF with varying compositions of CF.</p>
Full article ">Figure 10
<p>Variation in AC conductivity with respect to frequency at various temperatures.</p>
Full article ">Figure 11
<p>Arrhenius plot of ln resistance vs. 1/temperature.</p>
Full article ">Figure 12
<p>Dielectric properties of BNT-CF containing different concentrations of CF with varying frequency and temperature ranges.</p>
Full article ">
24 pages, 3999 KiB  
Review
Single Crystal Sn-Based Halide Perovskites
by Aditya Bhardwaj, Daniela Marongiu, Valeria Demontis, Angelica Simbula, Francesco Quochi, Michele Saba, Andrea Mura and Giovanni Bongiovanni
Nanomaterials 2024, 14(17), 1444; https://doi.org/10.3390/nano14171444 - 4 Sep 2024
Viewed by 2304
Abstract
Sn-based halide perovskites are expected to be the best replacement for toxic lead-based counterparts, owing to their similar ionic radii and the optimal band gap for use in solar cells, as well as their versatile use in light-emitting diodes and photodetection applications. Concerns, [...] Read more.
Sn-based halide perovskites are expected to be the best replacement for toxic lead-based counterparts, owing to their similar ionic radii and the optimal band gap for use in solar cells, as well as their versatile use in light-emitting diodes and photodetection applications. Concerns, however, exist about their stability under ambient conditions, an issue that is exacerbated in polycrystalline films because grain boundaries present large concentrations of defects and act as entrance points for oxygen and water, causing Sn oxidation. A current thriving research area in perovskite materials is the fabrication of perovskite single crystals, promising improved optoelectronic properties due to excellent uniformity, reduced defects, and the absence of grain boundaries. This review summarizes the most recent advances in the fabrication of single crystal Sn-based halide perovskites, with emphasis on synthesis methods, compositional engineering, and formation mechanisms, followed by a discussion of various challenges and appropriate strategies for improving their performance in optoelectronic applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photographs of the single crystals of Sn-based halide perovskites grown by top-seeded solution growth method: (<b>A</b>) (a) MASnI<sub>3</sub> and (b) FASnI<sub>3</sub> by optimization of growth conditions (reproduced with permission [<a href="#B26-nanomaterials-14-01444" class="html-bibr">26</a>], Copyright Wiley). (<b>B</b>) (a–e) Pb-Sn mixed-halide MAPb<sub>x</sub>Sn<sub>1−x</sub>Br<sub>3</sub> (MA-CH<sub>3</sub>NH<sub>3</sub>) single crystals (reproduced with permission [<a href="#B27-nanomaterials-14-01444" class="html-bibr">27</a>], Copyright, ACS Publications), hydrothermal method: (<b>C</b>) Color changes of Cs<sub>2</sub>SnCl<sub>6−x</sub>Br<sub>x</sub> from transparent to yellow to red (reproduced with permission [<a href="#B28-nanomaterials-14-01444" class="html-bibr">28</a>], Copyright Wiley). (<b>D</b>) Bridgman method for the growth of mixed-halide perovskite single crystals of Cs(Pb<sub>0.75</sub>Sn<sub>0.25</sub>)(Br<sub>1.00</sub>Cl<sub>2.00</sub>) (a) ingot after crystal growth, (b) polished crystal from cleaved part of the ingot (reproduced with permission [<a href="#B31-nanomaterials-14-01444" class="html-bibr">31</a>], Copyright Springer Nature. (<b>E</b>) Inverse temperature crystallization for the growth of mixed-halides (MAPbI<sub>3</sub>)<sub>x</sub>(FASnI<sub>3</sub>)<sub>1−x</sub> (x is 0.8, 0.5, 0.2) (reproduced with permission [<a href="#B34-nanomaterials-14-01444" class="html-bibr">34</a>], Copyright ACS Publications). (<b>F</b>) Space-confined method for the growth of mixed-halide (FASnI<sub>3</sub>)<sub>0.1</sub>(MAPbI<sub>3</sub>)<sub>0.9</sub> perovskites (reproduced with permission [<a href="#B36-nanomaterials-14-01444" class="html-bibr">36</a>], Copyright Royal Society of Chemistry).</p>
Full article ">Figure 2
<p>(<b>A</b>) Reductant engineering method using oxalic acid for the growth of FPEA<sub>2</sub>SnI<sub>4</sub> single crystals (reproduced with permission [<a href="#B37-nanomaterials-14-01444" class="html-bibr">37</a>], Copyright Wiley). (<b>B</b>) Ethylene glycol and HI-assisted modified temperature lowering method for fabrication of 2D (4-FPEA)<sub>2</sub>SnI<sub>4</sub> tin halide perovskite single crystals (reproduced with permission [<a href="#B38-nanomaterials-14-01444" class="html-bibr">38</a>], Copyright ACS Publications). (<b>C</b>) Synthesis process for fabrication of 1D perovskite single crystal-MDASn<sub>2</sub>I<sub>6</sub> and its corresponding image and dimensions (reproduced with permission [<a href="#B49-nanomaterials-14-01444" class="html-bibr">49</a>], Copyright ACS Publications). (<b>D</b>) Single crystals of (a) (R/S-<span class="html-italic">α</span>-PEA)SnCl<sub>3</sub> and (b) (R/S-<span class="html-italic">α</span>-PEA)SnBr<sub>3</sub> of sizes-6 mm, 10 mm by bottom seeded solution growth (reproduced with permission [<a href="#B50-nanomaterials-14-01444" class="html-bibr">50</a>], Copyright Wiley). (<b>E</b>) Single crystals of B-γ CsSnI<sub>3</sub> by solvent volatilization method (reproduced with permission [<a href="#B51-nanomaterials-14-01444" class="html-bibr">51</a>], Copyright Royal Society of Chemistry).</p>
Full article ">Figure 3
<p>(<b>A</b>,<b>B</b>) Energy band structures and DOS of CsSnCl<sub>3</sub>, CsSnBr<sub>3</sub>, CsSnI<sub>3</sub> (reproduced with permission [<a href="#B55-nanomaterials-14-01444" class="html-bibr">55</a>], Copyright MDPI). (<b>C</b>) Band-bowing in mixed-halide perovskites of MA(Pb<sub>1−x</sub>Sn<sub>x</sub>)I<sub>3</sub> (reproduced with permission [<a href="#B57-nanomaterials-14-01444" class="html-bibr">57</a>], Copyright ACS). (<b>D</b>) Partial density of states of p-orbital electron in MASn<sub>a</sub>Pb<sub>1-a</sub>I<sub>3</sub> mixed-halide perovskites (a-0, 0.25, 0.5, 0.75, 1) (reproduced with permission [<a href="#B58-nanomaterials-14-01444" class="html-bibr">58</a>], Copyright Elsevier).</p>
Full article ">Figure 4
<p>(<b>A</b>) Calculated defect formation energies in CsSnI<sub>3</sub> as a function of the chemical potentials of electrons, Cs, and Sn. Points A and E correspond to Sn-rich growth conditions, while points B, C, and E correspond to Sn-poor growth conditions. Empty (solid) circles represent acceptor (donor) defects. (<b>B</b>) Calculated transition energies for various defects, with acceptor (donor) defect levels shown in red (blue). The number of empty (solid) circles denotes the number of holes (electrons) released following defect ionization. (Reproduced with permission [<a href="#B66-nanomaterials-14-01444" class="html-bibr">66</a>], Copyright ACS).</p>
Full article ">Figure 5
<p>(<b>A</b>) (a) Schematic of Cs<sub>2</sub>SnCl<sub>6−x</sub>Br<sub>x</sub> single crystal photodetector device, (b) mechanism of the photodetector, where I<sub>p</sub>, I<sub>e</sub>, I<sub>D</sub> denote penetration, electron diffusion, and drift length; (c) current vs. time response at −20 V bias with illumination wavelengths of 590, 610, and 560 nm with light intensity of 1.3 mW/cm<sup>2</sup> (reproduced with permission [<a href="#B28-nanomaterials-14-01444" class="html-bibr">28</a>], Copyright Wiley). (<b>B</b>) (a–c) current vs. time response of (MAPbI<sub>3</sub>)<sub>x</sub>(FASnI<sub>3</sub>)<sub>1−x</sub> (x = 0.8, 0.5, and 0.2) single crystals at 0 V bias with channel width of 30, 50, and 100 m in wavelength range of 405–1064 nm with light intensity of 0.60 mW/cm<sup>2</sup> (reproduced with permission [<a href="#B34-nanomaterials-14-01444" class="html-bibr">34</a>], Copyright ACS). (<b>C</b>) Device structure of 2D Cs<sub>2</sub>SnI<sub>6</sub> single crystals and current vs. time response at different power densities at 1 V bias under illumination by light of wavelength 405 nm (reproduced with permission [<a href="#B72-nanomaterials-14-01444" class="html-bibr">72</a>], Copyright Elsevier). (<b>D</b>) (a–c) current vs. voltage curves of (C<sub>8</sub>H<sub>9</sub>F<sub>3</sub>N)<sub>2</sub>Pb<sub>1−x</sub>Sn<sub>x</sub>I<sub>4</sub> (x = 0, 0.5, and 1) single-crystal photodetectors, where the insets depict the device structure and optical microscope images in dark and illumination with light of wavelengths 405, 650 nm with light intensity of 0.006, 6.37, and 70.7 mW/mm<sup>2</sup>. The current vs. time curves were obtained at 5, 10 V bias (reproduced with permission [<a href="#B25-nanomaterials-14-01444" class="html-bibr">25</a>], Copyright Royal Society of Chemistry).</p>
Full article ">Figure 6
<p>(<b>A</b>) (a) Cross-section scanning electron microscope image of C-CsSnI<sub>3</sub>, (b) photocurrent density vs. voltage curve, (c) stabilized power output, and (d) external quantum efficiency curves of C-CsSnI<sub>3</sub>, S-CsSnI<sub>3</sub> perovskite solar cells (reproduced with permission [<a href="#B51-nanomaterials-14-01444" class="html-bibr">51</a>], Copyright Royal Society of Chemistry). (<b>B</b>,<b>C</b>) Mobility-lifetime values, sensitivity, of X-ray detectors of FPEA<sub>2</sub>PbI<sub>4</sub> and heterojunction FPEA<sub>2</sub>PbI<sub>4</sub>-FPEA<sub>2</sub>SnI<sub>4</sub> tin single crystal, X-ray photocurrent signals, operational stability of FPEA<sub>2</sub>PbI<sub>4</sub>-FPEA<sub>2</sub>SnI<sub>4</sub> (reproduced with permission [<a href="#B37-nanomaterials-14-01444" class="html-bibr">37</a>], Copyright Wiley). (<b>D</b>) (a,d) Microphotoluminescence spectra, (b,e) FWHM, integrated intensity of emission peaks as function of pump density, (c,f) time-resolved photoluminescence spectra below and above pump fluence threshold of (TEA)<sub>2</sub>SnI<sub>4</sub> and (TEA)<sub>2</sub>(MA)Sn<sub>2</sub>I<sub>7</sub> tin perovskite single crystals (reproduced with permission [<a href="#B42-nanomaterials-14-01444" class="html-bibr">42</a>], Copyright Royal Society of Chemistry). (<b>E</b>) (a) FET device structure, (b) transfer curve, (c) current vs. voltage curve, (d) on–off switching curves showing good stability of 4AMPSnI<sub>4</sub> Dion–Jacobson tin perovskite single crystals (Reproduced with permission [<a href="#B43-nanomaterials-14-01444" class="html-bibr">43</a>], Copyright ACS).</p>
Full article ">
18 pages, 4690 KiB  
Article
Preparation and Properties of Nb5+-Doped BCZT-Based Ceramic Thick Films by Scraping Process
by Yang Zou, Bijun Fang, Xiaolong Lu, Shuai Zhang and Jianning Ding
Materials 2024, 17(17), 4348; https://doi.org/10.3390/ma17174348 - 2 Sep 2024
Cited by 1 | Viewed by 833
Abstract
A bottleneck characterized by high strain and low hysteresis has constantly existed in the design process of piezoelectric actuators. In order to solve the problem that actuator materials cannot simultaneously exhibit large strain and low hysteresis under relatively high electric fields, Nb5+ [...] Read more.
A bottleneck characterized by high strain and low hysteresis has constantly existed in the design process of piezoelectric actuators. In order to solve the problem that actuator materials cannot simultaneously exhibit large strain and low hysteresis under relatively high electric fields, Nb5+-doped 0.975(Ba0.85Ca0.15)[(Zr0.1Ti0.9)0.999Nb0.001]O3-0.025(Bi0.5Na0.5)ZrO3 (BCZTNb0.001-0.025BiNZ) ceramic thick films were prepared by a film scraping process combined with a solid-state twin crystal method, and the influence of sintering temperature was studied systematically. All BCZTNb0.001-0.025BiNZ ceramic thick films sintered at different sintering temperatures have a pure perovskite structure with multiphase coexistence, dense microstructure and typical dielectric relaxation behavior. The conduction mechanism of all samples at high temperatures is dominated by oxygen vacancies confirmed by linear fitting using the Arrhenius law. As the sintering temperature elevates, the grain size increases, inducing the improvement of dielectric, ferroelectric and field-induced strain performance. The 1325 °C sintered BCZTNb0.001-0.025BiNZ ceramic thick film has the lowest hysteresis (1.34%) and relatively large unipolar strain (0.104%) at 60 kV/cm, showing relatively large strain and nearly zero strain hysteresis compared with most previously reported lead-free piezoelectric ceramics and presenting favorable application prospects in the actuator field. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of preparing ceramic thick films by scraping process.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures; (<b>b</b>) locally amplified XRD patterns of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films around (200) peak.</p>
Full article ">Figure 3
<p>XRD Rietveld refinement results of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures: (<b>a</b>) 1305 °C; (<b>b</b>) 1325 °C; (<b>c</b>) 1345 °C.</p>
Full article ">Figure 3 Cont.
<p>XRD Rietveld refinement results of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures: (<b>a</b>) 1305 °C; (<b>b</b>) 1325 °C; (<b>c</b>) 1345 °C.</p>
Full article ">Figure 4
<p>SEM images and grain size statistics (insets) of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures: (<b>a</b>) 1305 °C; (<b>b</b>) 1325 °C; (<b>c</b>) 1345 °C.</p>
Full article ">Figure 5
<p>(<b>a</b>) Dielectric performance-temperature curves of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures at 10 kHz; (<b>b</b>) curves of T<sub>m</sub> and ε<sub>m</sub> with sintering temperature at 10 kHz.</p>
Full article ">Figure 6
<p>Dielectric performance-temperature curves at several frequencies of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures: (<b>a</b>) 1305 °C; (<b>b</b>) 1315 °C; (<b>c</b>) 1325 °C; (<b>d</b>) 1335 °C; (<b>e</b>) 1345 °C.</p>
Full article ">Figure 7
<p>Dielectric exponential law fitting of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures using 10 kHz data: (<b>a</b>) 1305 °C; (<b>b</b>) 1315 °C; (<b>c</b>) 1325 °C; (<b>d</b>) 1335 °C; (<b>e</b>) 1345 °C.</p>
Full article ">Figure 8
<p>Z″ versus Z′ curves and the fitted equivalent circuits at 430 °C (insets) of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures: (<b>a</b>) 1305 °C; (<b>b</b>) 1325 °C; (<b>c</b>) 1345 °C.</p>
Full article ">Figure 9
<p>σ<sub>dc</sub> and τ versus 1/T curves of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures: (<b>a</b>) 1305 °C; (<b>b</b>) 1325 °C; (<b>c</b>) 1345 °C.</p>
Full article ">Figure 10
<p>(<b>a</b>) P-E hysteresis loops at 60 kV/cm and 1 Hz and (<b>b</b>) bipolar S-E curves at 60 kV/cm and 10 Hz of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures; (<b>c</b>) curves of polarization and coercive field with sintering temperature of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films.</p>
Full article ">Figure 11
<p>(<b>a</b>) Unipolar S-E curves at 60 kV/cm and 10 Hz of the BCZTNb<sub>0.001</sub>-0.025BiNZ ceramic thick films sintered at different sintering temperatures; (<b>b</b>) curves of S<sub>unipolar</sub>, d<sub>33</sub><sup>*</sup> and Hys with sintering temperature.</p>
Full article ">
15 pages, 6959 KiB  
Article
The Influence of Lanthanum Admixture on Microstructure and Electrophysical Properties of Lead-Free Barium Iron Niobate Ceramics
by Dariusz Bochenek, Dagmara Brzezińska, Przemysław Niemiec and Lucjan Kozielski
Materials 2024, 17(15), 3666; https://doi.org/10.3390/ma17153666 - 25 Jul 2024
Viewed by 672
Abstract
This article presents the research results of lead-free Ba1−3/2xLax(Fe0.5Nb0.5)O3 (BFNxLa) ceramic materials doped with La (x = 0.00–0.06) obtained via the solid-state reaction method. The tests of the BFNx [...] Read more.
This article presents the research results of lead-free Ba1−3/2xLax(Fe0.5Nb0.5)O3 (BFNxLa) ceramic materials doped with La (x = 0.00–0.06) obtained via the solid-state reaction method. The tests of the BFNxLa ceramic samples included structural (X-ray), morphological (SEM, EDS, EPMA), DC electrical conductivity, and dielectric measurements. For all BFNxLa ceramic samples, the X-ray tests revealed a perovskite-type cubic structure with the space group Pm3¯m. In the case of the samples with the highest amount of lanthanum, i.e., for x = 0.04 (BFN4La) and x = 0.06 (BFN6La), the X-ray analysis also showed a small amount of pyrochlore LaNbO4 secondary phase. In the microstructure of BFNxLa ceramic samples, the average grain size decreases with increasing La content, affecting their dielectric properties. The BFN ceramics show relaxation properties, diffusion phase transition, and very high permittivity at room temperature (56,750 for 1 kHz). The admixture of lanthanum diminishes the permittivity values but effectively reduces the dielectric loss and electrical conductivity of the BFNxLa ceramic samples. All BFNxLa samples show a Debye-like relaxation behavior at lower frequencies; the frequency dispersion of the dielectric constant becomes weaker with increasing admixtures of lanthanum. Research has shown that using an appropriate amount of lanthanum introduced to BFN can obtain high permittivity values while decreasing dielectric loss and electrical conductivity, which predisposes them to energy storage applications. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of the BFN<span class="html-italic">x</span>La ceramic samples at RT.</p>
Full article ">Figure 2
<p>SEM images of the microstructure of fractures of the ceramic samples: (<b>a</b>,<b>a′</b>) BFN0La, (<b>b</b>,<b>b′</b>) BFN1La, (<b>c</b>,<b>c′</b>) BFN2La, (<b>d</b>,<b>d′</b>) BFN3La, (<b>e</b>,<b>e′</b>) BFN4La, and (<b>f</b>,<b>f′</b>) BFN6La, respectively. Next to it, grain size distribution diagrams.</p>
Full article ">Figure 3
<p>The EDS analysis of chemical elements of the ceramic samples: (<b>a</b>) BFN0La, (<b>b</b>) BFN1La, (<b>c</b>) BFN2La, (<b>d</b>) BFN3La, (<b>e</b>) BFN4La, and (<b>f</b>) BFN6La, respectively.</p>
Full article ">Figure 4
<p>EPMA test results for the BFN<span class="html-italic">x</span>La ceramics.</p>
Full article ">Figure 5
<p>The lnσ<sub>DC</sub>(1000/<span class="html-italic">T</span>) relationship of the BFN<span class="html-italic">x</span>La ceramic samples.</p>
Full article ">Figure 6
<p>Frequency dependence of real <span class="html-italic">ε</span>′ and imaginary <span class="html-italic">ε</span>″ parts of the dielectric constant of the BFN<span class="html-italic">x</span>La ceramics.</p>
Full article ">Figure 7
<p>Temperature dependencies of the (<b>a</b>) permittivity and (<b>b</b>) dielectric loss factor for BFN<span class="html-italic">x</span>La ceramics measured at 1 kHz.</p>
Full article ">
13 pages, 2970 KiB  
Article
Dual Light Emission of CsSnI3-Based Powders Synthesized via a Mechanochemical Process
by Xuan Huang, Xiaobing Tang, Xiyu Wen, Yuebin Charles Lu and Fuqian Yang
Materials 2024, 17(14), 3577; https://doi.org/10.3390/ma17143577 - 19 Jul 2024
Viewed by 797
Abstract
Lead toxicity has hindered the wide applications of lead halide perovskites in optoelectronics and bioimaging. A significant amount of effort has been made to synthesize lead-free halide perovskites as alternatives to lead halide perovskites. In this work, we demonstrate the feasibility of synthesizing [...] Read more.
Lead toxicity has hindered the wide applications of lead halide perovskites in optoelectronics and bioimaging. A significant amount of effort has been made to synthesize lead-free halide perovskites as alternatives to lead halide perovskites. In this work, we demonstrate the feasibility of synthesizing CsSnI3-based powders mechanochemically with dual light emissions under ambient conditions from CsI and SnI2 powders. The formed CsSnI3-based powders are divided into CsSnI3-dominated powders and CsSnI3-contained powders. Under the excitation of ultraviolet light of 365 nm in wavelength, the CsSnI3-dominated powders emit green light with a wavelength centered at 540 nm, and the CsSnI3-contained powders emit orange light with a wavelength centered at 608 nm. Both the CsSnI3-dominated and CsSnI3-contained powders exhibit infrared emission with the peak emission wavelengths centered at 916 nm and 925 nm, respectively, under a laser of 785 nm in wavelength. From the absorbance spectra, we obtain bandgaps of 2.32 eV and 2.08 eV for the CsSnI3-dominated and CsSnI3-contained powders, respectively. The CsSnI3-contained powders exhibit the characteristics of thermal quenching and photoelectrical response under white light. Full article
Show Figures

Figure 1

Figure 1
<p>Optical images of the prepared powders under white light and UV light of 365 nm in wavelength: (<b>a</b>) CsSnI<sub>3</sub>-dominated powders, and (<b>b</b>) CsSnI<sub>3</sub>-contained powders.</p>
Full article ">Figure 2
<p>SEM images of green-emitting powders: (<b>a</b>,<b>b</b>) plate-like structure, (<b>c</b>) Cs<sub>2</sub>SnI<sub>6</sub> octahedral microcrystals, and (<b>d</b>) spherical nanocrystals; SEM images of orange-emitting powders: (<b>e</b>,<b>f</b>) spherical nanocrystals, (<b>g</b>) octahedral microcrystals, and (<b>h</b>) rod-like microcrystal.</p>
Full article ">Figure 3
<p>XRD patterns of (<b>a</b>) freshly prepared CsSnI<sub>3</sub>-dominated powders and the one stored in a vacuum chamber for one week, and (<b>b</b>) freshly prepared CsSnI<sub>3</sub>-contained powders and the one stored under ambient conditions for one week.</p>
Full article ">Figure 4
<p>PL spectra of (<b>a</b>) freshly prepared CsSnI<sub>3</sub>-dominated and CsSnI<sub>3</sub>-contained powders under UV light of 365 nm in wavelength, and (<b>b</b>) freshly prepared CsSnI<sub>3</sub>-dominated and CsSnI<sub>3</sub>-contained powders under a laser of 785 nm in wavelength.</p>
Full article ">Figure 5
<p>Absorbance spectra and Tauc plots (insets) of freshly prepared (<b>a</b>) CsSnI<sub>3</sub>-dominated powders and (<b>b</b>) CsSnI<sub>3</sub>-contained powders.</p>
Full article ">Figure 6
<p>Long-term stability of the CsSnI<sub>3</sub>-based powders over a period of 7 days: (<b>a</b>) CsSnI<sub>3</sub>-dominated powders in a vacuum chamber, and (<b>b</b>) CsSnI<sub>3</sub>-contained powder under ambient conditions.</p>
Full article ">Figure 7
<p>Temperature effects on the PL characteristics of the CsSnI<sub>3</sub>-contained powders: (<b>a</b>) PL spectra at different temperatures, (<b>b</b>) PL intensity vs. temperature, and (<b>c</b>) photon energy vs. temperature.</p>
Full article ">Figure 8
<p>Photo-response of the CsSnI<sub>3</sub>-dominated powders during a voltage sweeping from 0 V to 200 V (<b>a</b>) with and without the illumination of white light and (<b>b</b>) under a 20 s light on-and-off cycle for 385 s; photo-response of the CsSnI<sub>3</sub>-contained powders during a voltage sweeping from 0 V to 200 V, (<b>c</b>) with and without the illumination of white light and (<b>d</b>) under a 20 s light on-and-off cycle for 385 s.</p>
Full article ">
8 pages, 1252 KiB  
Article
Theoretical Design of Tellurium-Based Two-Dimensional Perovskite Photovoltaic Materials
by Chunhong Long and Peihao Huang
Molecules 2024, 29(13), 3155; https://doi.org/10.3390/molecules29133155 - 2 Jul 2024
Viewed by 1134
Abstract
In recent years, the photoelectric conversion efficiency of three–dimensional (3D) perovskites has seen significant improvements. However, the commercial application of 3D perovskites is hindered by stability issues and the toxicity of lead. Two–dimensional (2D) perovskites exhibit good stability but suffer from low efficiency. [...] Read more.
In recent years, the photoelectric conversion efficiency of three–dimensional (3D) perovskites has seen significant improvements. However, the commercial application of 3D perovskites is hindered by stability issues and the toxicity of lead. Two–dimensional (2D) perovskites exhibit good stability but suffer from low efficiency. Designing efficient and stable lead–free 2D perovskite materials remains a crucial unsolved scientific challenge. This study, through structural prediction combined with first–principles calculations, successfully predicts a 2D perovskite, CsTeI5. Theoretical calculations indicate that this compound possesses excellent stability and a theoretical efficiency of up to 29.3%, showing promise for successful application in thin–film solar cells. This research provides a new perspective for the design of efficient and stable lead-free 2D perovskites. Full article
(This article belongs to the Special Issue Novel Two-Dimensional Energy-Environmental Materials)
Show Figures

Figure 1

Figure 1
<p>The crystal structure and stability of CsTeI<sub>5</sub>. (<b>a</b>) The crystal structure view along the a–axis of CsTeI<sub>5</sub>. The red frame is the primitive cell. (<b>b</b>) The crystal structure view along the c–axis of CsTeI<sub>5</sub>. (<b>c</b>) The formation energy of CsTeI<sub>5</sub> relative to CsI and TeI<sub>4</sub>. (<b>d</b>) Mean square displacements (MSDs) of CsTeI<sub>5</sub> at 500 K.</p>
Full article ">Figure 2
<p>The electronic properties of CsTeI<sub>5</sub>. (<b>a</b>) Band structure and transition dipole moment of CsTeI<sub>5</sub>. The VBM and CBM are highlighted as red dots. (<b>b</b>) Electronic density of states of CsTeI<sub>5</sub>. (<b>c</b>) Projected density of states for CsTeI<sub>5</sub>. (<b>d</b>) Joint density of states for CsTeI<sub>5</sub>.</p>
Full article ">Figure 3
<p>Optical properties. (<b>a</b>) Absorption coefficients. (<b>b</b>) Theoretical photoelectric conversion efficiency.</p>
Full article ">
16 pages, 5467 KiB  
Article
Novel Sol-Gel Synthesis Route for Ce- and V-Doped Ba0.85Ca0.15Ti0.9Zr0.1O3 Piezoceramics
by Larissa S. Marques, Michelle Weichelt, Michel Kuhfuß, Carlos R. Rambo and Tobias Fey
Materials 2024, 17(13), 3228; https://doi.org/10.3390/ma17133228 - 1 Jul 2024
Viewed by 850
Abstract
To meet the current demand for lead-free piezoelectric ceramics, a novel sol-gel synthesis route is presented for the preparation of Ba0.85Ca0.15Ti0.9Zr0.1O3 doped with cerium (Ce = 0, 0.01, and 0.02 mol%) and vanadium (V [...] Read more.
To meet the current demand for lead-free piezoelectric ceramics, a novel sol-gel synthesis route is presented for the preparation of Ba0.85Ca0.15Ti0.9Zr0.1O3 doped with cerium (Ce = 0, 0.01, and 0.02 mol%) and vanadium (V = 0, 0.3, and 0.4 mol%). X-ray diffraction patterns reveal the formation of a perovskite phase (space group P4mm) for all samples after calcination at 800 °C and sintering at 1250, 1350, and 1450 °C, where it is proposed that both dopants occupy the B site. Sintering studies show that V doping allows the sintering temperature to be reduced to at least 1250 °C. Undoped BCZT samples sintered at the same temperature show reduced functional properties compared to V-doped samples, i.e., d33 values increase by an order of magnitude with doping. The dissipation factor tan δ decreases with increasing sintering temperature for all doping concentrations, while the Curie temperature TC increases for all V-doped samples, reaching 120 °C for high-concentration co-doped samples. All results indicate that vanadium doping can facilitate the processing of BCZT at lower sintering temperatures without compromising performance while promoting thermal property stability. Full article
(This article belongs to the Special Issue Properties of Ceramic Composites)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the synthesis process for pure BCZT, V-doped BCZT, Ce-doped BCZT, and Ce and V co-doped BCZT sols.</p>
Full article ">Figure 2
<p>X-ray diffraction patterns of calcinated samples at 800 °C for 5 h from (<b>a</b>) 15 to 90° for undoped and doped samples; (<b>b</b>) 30 to 47° for undoped and doped samples; (<b>c</b>) 15 to 90° for co-doped samples; and (<b>d</b>) 30 to 47° for co-doped samples.</p>
Full article ">Figure 2 Cont.
<p>X-ray diffraction patterns of calcinated samples at 800 °C for 5 h from (<b>a</b>) 15 to 90° for undoped and doped samples; (<b>b</b>) 30 to 47° for undoped and doped samples; (<b>c</b>) 15 to 90° for co-doped samples; and (<b>d</b>) 30 to 47° for co-doped samples.</p>
Full article ">Figure 3
<p>Density values as a function of sintering temperature for (<b>a</b>) Ce-doped BCZT samples; (<b>b</b>) V-doped BCZT samples; and (<b>c</b>) Ce and V co-doped BCZT samples.</p>
Full article ">Figure 4
<p>X-ray diffraction patterns of sintered samples regarding (<b>a</b>) Ce-doped samples compared to undoped BCZT samples, 15°–90°; (<b>b</b>) magnified peaks of Ce-doped samples compared to BCZT samples, 31°–47°; (<b>c</b>) V-doped samples compared to undoped BCZT samples, 15°–90°; (<b>d</b>) magnified peaks of V-doped samples compared to undoped BCZT samples, 15°–90°; (<b>e</b>) co-doped samples compared to BCZT samples, 15°–90°; and (<b>f</b>) magnified peaks of co-doped samples, 31°–47°.</p>
Full article ">Figure 5
<p>Photographs of selected sintered samples in pellet and powder form.</p>
Full article ">Figure 6
<p>Piezoelectric measurements of all sintered samples as a function of sintering temperature: (<b>a</b>) d<sub>33</sub> values for Ce-doped samples; (<b>b</b>) d<sub>33</sub> values for V-doped samples; and (<b>c</b>) d<sub>33</sub> values for co-doped samples.</p>
Full article ">Figure 7
<p>SEM pictures of sintered samples: (<b>a</b>) BCZT-U sample sintered at 1450 °C; (<b>b</b>) BCZT-A sample sintered at 1350 °C; (<b>c</b>) BCZT-AB sample sintered at 1350 °C.</p>
Full article ">Figure 8
<p>Dielectric and piezoelectric measurements of all sintered samples as a function of sintering temperature: (<b>a</b>) d<sub>33</sub> values for Ce-doped samples; (<b>b</b>) capacitance values for Ce-doped samples; (<b>c</b>) d<sub>33</sub> values for V-doped samples; (<b>d</b>) capacitance values for V-doped samples; (<b>e</b>) d<sub>33</sub> values for co-doped samples; (<b>f</b>) capacitance values for co-doped samples.</p>
Full article ">
Back to TopTop