[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

Journals

Article Types

Countries / Regions

Search Results (52)

Search Parameters:
Keywords = barium 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 (registering DOI) - 26 Dec 2024
Viewed by 283
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 ">
15 pages, 15469 KiB  
Article
Unveiling BaTiO3-SrTiO3 as Anodes for Highly Efficient and Stable Lithium-Ion Batteries
by Nischal Oli, Nawraj Sapkota, Brad R. Weiner, Gerardo Morell and Ram S. Katiyar
Nanomaterials 2024, 14(21), 1723; https://doi.org/10.3390/nano14211723 - 29 Oct 2024
Viewed by 1085
Abstract
Amidst the swift expansion of the electric vehicle industry, the imperative for alternative battery technologies that balance economic feasibility with sustainability has reached unprecedented importance. Herein, we utilized Perovskite-based oxide compounds barium titanate (BaTiO3) and strontium titanate (SrTiO3) nanoparticles [...] Read more.
Amidst the swift expansion of the electric vehicle industry, the imperative for alternative battery technologies that balance economic feasibility with sustainability has reached unprecedented importance. Herein, we utilized Perovskite-based oxide compounds barium titanate (BaTiO3) and strontium titanate (SrTiO3) nanoparticles as anode materials for lithium-ion batteries from straightforward and standard carbonate-based electrolyte with 10% fluoroethylene carbonate (FEC) additive [1M LiPF6 (1:1 EC: DEC) + 10% FEC]. SrTiO3 and BaTiO3 electrodes can deliver a high specific capacity of 80 mA h g−1 at a safe and low average working potential of ≈0.6 V vs. Li/Li+ with excellent high-rate performance with specific capacity of ~90 mA h g−1 at low current density of 20 mA g−1 and specific capacity of ~80 mA h g−1 for over 500 cycles at high current density of 100 mA g−1. Our findings pave the way for the direct utilization of perovskite-type materials as anode materials in Li-ion batteries due to their promising potential for Li+ ion storage. This investigation addresses the escalating market demands in a sustainable manner and opens avenues for the investigation of diverse perovskite oxides as advanced anodes for next-generation metal-ion batteries. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) XRD pattern SrTiO<sub>3</sub> (STO). (<b>b</b>) Crystal structure of STO. (<b>c</b>) Raman shifts of STO. (<b>d</b>) XRD pattern of BaTiO<sub>3</sub> (BTO). (<b>e</b>) Crystal structure of BTO. (<b>f</b>) Raman shifts of BTO.</p>
Full article ">Figure 2
<p>SEM images of BTO and STO. (<b>a</b>) BTO at 200 nm. (<b>b</b>) BTO at 500 nm. (<b>c</b>) STO at 200 nm. (<b>d</b>) STO at 500 nm.</p>
Full article ">Figure 3
<p>SEM mapping images of BTO and STO at 2 µm. (<b>a</b>–<b>d</b>) Ba, Ti, O, and BaTiO<sub>3</sub>, respectively. (<b>e</b>–<b>h</b>) Sr, Ti, O, and SrTiO<sub>3</sub>, respectively.</p>
Full article ">Figure 4
<p>Electrochemical performance of BTO at different voltage ranges in LIBs. (<b>a</b>) Galvanostatic charge–discharge (GCD) at 0.001–1.5 V. (<b>b</b>) Cyclic performance at 0.001–1.5 V. (<b>c</b>) Galvanostatic charge–discharge (GCD) at 0.001–2.0 V. (<b>d</b>) Cyclic performance at 0.001–2.0 V.</p>
Full article ">Figure 5
<p>Electrochemical performance of STO at different voltage ranges in LIBs. (<b>a</b>) Galvanostatic charge–discharge (GCD) at 0.001–1.5 V. (<b>b</b>) Cyclic performance at 0.001–1.5 V. (<b>c</b>) Galvanostatic charge–discharge (GCD) at 0.001–2.0 V. (<b>d</b>) Cyclic performance at 0.001–2.0 V.</p>
Full article ">Figure 6
<p>Rate performance of STO and BTO in 0.001–1.5 V. (<b>a</b>) STO rate performance in the range of 20 to 150 mA g<sup>−1</sup>. (<b>b</b>) STO rate curve profile at different current densities in the range of 20 to 150 mA g<sup>−1</sup>. (<b>c</b>) BTO rate performance in the range of 20 to 150 mA g<sup>−1</sup>. (<b>d</b>) BTO discharge/charge voltage profile at different current densities in the range of 20 to 150 mA g<sup>−1</sup>.</p>
Full article ">Figure 7
<p>Cyclic voltammetry (CV) curves of the BTO and STO at room temperature. (<b>a</b>) CV of BTO in the first 3 cycles at a scan rate of 0.1 mV s<sup>−1</sup>. (<b>b</b>) CV of BTO in the different scan rates of 0.1–1.2 mV s<sup>−1</sup>. (<b>c</b>) CV of STO in the first 3 cycles at a scan rate of 0.1 mV s<sup>−1</sup>. (<b>d</b>) CV of STO in the different scan rates of 0.1–1.2 mV s<sup>−1</sup>.</p>
Full article ">Figure 8
<p>(<b>a</b>) Peak current (i) vs. v<sup>1/2</sup> (square root of scan rate (v)) of BTO. (<b>b</b>) Log (i) (peak current (i)) vs. log (v) (scan rate (v)) of STO. (<b>c</b>) Log (i) vs. log (v) of BTO and STO.</p>
Full article ">Figure 9
<p>X-ray photoelectron spectroscopy (XPS) measurement. (<b>a</b>) Pristine STO, lithiation (discharge), lithiation–delithiation (discharge–charge), after 3 cycles. (<b>b</b>) O1s. (<b>c</b>) STO survey different conditions. (<b>d</b>) Pristine BTO, lithiation (discharge), lithiation–delithiation (discharge–charge), after 3 cycles. (<b>e</b>) O1s. (<b>f</b>) BTO survey different conditions.</p>
Full article ">Figure 10
<p>Electrochemical impedance spectroscopy (EIS) STO- and BTO-based electrodes. (<b>a</b>) STO before charging. (<b>b</b>) STO after 40 charge–discharge. (<b>c</b>) BTO before charging. (<b>d</b>) BTO after 40 cycles.</p>
Full article ">
27 pages, 2743 KiB  
Review
Exploring the Potential of Cold Sintering for Proton-Conducting Ceramics: A Review
by Andrea Bartoletti, Elisa Mercadelli, Angela Gondolini and Alessandra Sanson
Materials 2024, 17(20), 5116; https://doi.org/10.3390/ma17205116 - 19 Oct 2024
Viewed by 1897
Abstract
Proton-conducting ceramic materials have emerged as effective candidates for improving the performance of solid oxide cells (SOCs) and electrolyzers (SOEs) at intermediate temperatures. BaCeO3 and BaZrO3 perovskites doped with rare-earth elements such as Y2O3 (BCZY) are well known [...] Read more.
Proton-conducting ceramic materials have emerged as effective candidates for improving the performance of solid oxide cells (SOCs) and electrolyzers (SOEs) at intermediate temperatures. BaCeO3 and BaZrO3 perovskites doped with rare-earth elements such as Y2O3 (BCZY) are well known for their high proton conductivity, low operating temperature, and chemical stability, which lead to SOCs’ improved performance. However, the high sintering temperature and extended processing time needed to obtain dense BCZY-type electrolytes (typically > 1350 °C) to be used as SOC electrolytes can cause severe barium evaporation, altering the stoichiometry of the system and consequently reducing the performance of the final device. The cold sintering process (CSP) is a novel sintering technique that allows a drastic reduction in the sintering temperature needed to obtain dense ceramics. Using the CSP, materials can be sintered in a short time using an appropriate amount of a liquid phase at temperatures < 300 °C under a few hundred MPa of uniaxial pressure. For these reasons, cold sintering is considered one of the most promising ways to obtain ceramic proton conductors in mild conditions. This review aims to collect novel insights into the application of the CSP with a focus on BCZY-type materials, highlighting the opportunities and challenges and giving a vision of future trends and perspectives. Full article
(This article belongs to the Section Advanced and Functional Ceramics and Glasses)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of the cold sintering process (<b>a</b>) and the thermocompression apparatus (<b>b</b>).</p>
Full article ">Figure 2
<p>Schematic representation of mechanical–chemical effects during cold sintering process.</p>
Full article ">Figure 3
<p>Parameters affecting the cold sintering process [<a href="#B127-materials-17-05116" class="html-bibr">127</a>].</p>
Full article ">Figure 4
<p>A schematic representation of the stages involved in the cold sintering process [<a href="#B152-materials-17-05116" class="html-bibr">152</a>].</p>
Full article ">Figure 5
<p>STEM micrographs and corresponding EDS maps (the scale bars are 5 nm) of a GB in a sample produced by cold sintering + PA (<b>a</b>) and conventionally sintered (<b>b</b>). Atomic Probe Tomography investigation of grain boundaries (scale bar is 50 nm) in the cold-sintered +PA sample (<b>c</b>) and the relative composition profile (orange arrow in (<b>c</b>)) of a random GB (<b>d</b>,<b>e</b>) [<a href="#B149-materials-17-05116" class="html-bibr">149</a>].</p>
Full article ">Figure 6
<p>Microstructures of the BZCY-BZCY/NiO half-cell after cold sintering and post-annealing (<b>a</b>,<b>b</b>) and after subsequent reduction treatment (<b>c</b>,<b>d</b>); detail 1 shows the reduced Ni metal structure and detail 2 the remaining intact BZCY network [<a href="#B203-materials-17-05116" class="html-bibr">203</a>].</p>
Full article ">
33 pages, 3665 KiB  
Review
Role of Sintering Aids in Electrical and Material Properties of Yttrium- and Cerium-Doped Barium Zirconate Electrolytes
by Shivesh Loganathan, Saheli Biswas, Gurpreet Kaur and Sarbjit Giddey
Processes 2024, 12(10), 2278; https://doi.org/10.3390/pr12102278 - 18 Oct 2024
Viewed by 1039
Abstract
Ceramic proton conductors have the potential to lower the operating temperature of solid oxide cells (SOCs) to the intermediate temperature range of 400–600 °C. This is attributed to their superior ionic conductivity compared to oxide ion conductors under these conditions. However, prominent proton-conducting [...] Read more.
Ceramic proton conductors have the potential to lower the operating temperature of solid oxide cells (SOCs) to the intermediate temperature range of 400–600 °C. This is attributed to their superior ionic conductivity compared to oxide ion conductors under these conditions. However, prominent proton-conducting materials, such as yttrium-doped barium cerates and zirconates with specified compositions like BaCe1−xYxO3−δ (BCY), BaZr1−xYxO3−δ (BZY), and Ba(Ce,Zr)1−yYyO3−δ (BCZY), face significant challenges in achieving dense electrolyte membranes. It is suggested that the incorporation of transition and alkali metal oxides as sintering additives can induce liquid phase sintering (LPS), offering an efficient method to facilitate the densification of these proton-conducting ceramics. However, current research underscores that incorporating these sintering additives may lead to adverse secondary effects on the ionic transport properties of these materials since the concentration and mobility of protonic defects in a perovskite are highly sensitive to symmetry change. Such a drop in ionic conductivity, specifically proton transference, can adversely affect the overall performance of cells. The extent of variation in the proton conductivity of the perovskite BCZY depends on the type and concentration of the sintering aid, the nature of the sintering aid precursors used, the incorporation technique, and the sintering profile. This review provides a synopsis of various potential sintering techniques, explores the influence of diverse sintering additives, and evaluates their effects on the densification, ionic transport, and electrochemical properties of BCZY. We also report the performance of most of these combinations in an actual test environment (fuel cell or electrolysis mode) and comparison with BCZY. Full article
(This article belongs to the Section Chemical Processes and Systems)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic showing working principle of proton-conducting solid oxide fuel cell (H-SOFCs) (<b>a</b>) and proton-conducting solid oxide electrolytic cell (H-SOECs) (<b>b</b>).</p>
Full article ">Figure 2
<p>Depiction of bulk, grain, and electrode resistances in Nyquist plot for BCZY (<b>a</b>), reproduced with permission from Elsevier [<a href="#B54-processes-12-02278" class="html-bibr">54</a>]; BCZY electrolyte film thickness and porosity as function of sintering temperature replotted [<a href="#B44-processes-12-02278" class="html-bibr">44</a>] (<b>b</b>); X-ray diffractogram showing peak shift of BaCe<sub>0.3</sub>Zr<sub>0.55</sub>Y<sub>0.15</sub>O<sub>3−δ</sub> (BCZY35) upon adding NiO, CuO, and ZnO sintering aids, reproduced with permission from Elsevier [<a href="#B54-processes-12-02278" class="html-bibr">54</a>] (<b>c</b>).</p>
Full article ">Figure 3
<p>Arrhenius plots of total (<b>a</b>), grain boundary (<b>b</b>), and grain bulk (<b>c</b>) conductivity of 1 wt% NiO-, ZnO-, and CuO-added BaCe<sub>0.3</sub>Zr<sub>0.55</sub>Y<sub>0.15</sub>O<sub>3−δ</sub> (BCZY35) in 3% humid H<sub>2</sub> reproduced with permission from Elsevier [<a href="#B54-processes-12-02278" class="html-bibr">54</a>]; Arrhenius plots of grain interior (GI), grain boundary (GB), and total conductivity of BZY10 + 0.2 wt%NiO sintered at 1500 (<b>d</b>) and BZY10 sintered at 1600 (<b>e</b>) under 2.7% humidified H<sub>2</sub> and Ar (1:19 <span class="html-italic">v</span>/<span class="html-italic">v</span>), replotted from [<a href="#B95-processes-12-02278" class="html-bibr">95</a>].</p>
Full article ">Figure 4
<p>Cross-sectional FE-SEM images for BCZY63 pellets doped with (<b>A</b>) CuO—1450 °C, (<b>B</b>) ZnO—1450 °C, (<b>C</b>) Fe<sub>2</sub>O<sub>3</sub>—1600 °C, (<b>D</b>) Cr<sub>2</sub>O<sub>3</sub>—1600 °C, (<b>E</b>) PdO—1600 °C, and (<b>F</b>) Control—1650 °C, reproduced with permission from Elsevier [<a href="#B109-processes-12-02278" class="html-bibr">109</a>]. Protonic conductivity as a function of temperature for BCZY doped with ZnO and CuO (<b>G</b>), Fe<sub>2</sub>O<sub>3</sub> and MnO<sub>2</sub> (<b>H</b>), and Cr<sub>2</sub>O<sub>3</sub> and PdO (<b>I</b>) under different atmospheres, replotted from an original paper [<a href="#B109-processes-12-02278" class="html-bibr">109</a>]. The schematic shows that no sintering aid leads to porosity, whereas an excess of one causes the accumulation of unwanted secondary phases along the grain boundaries (<b>J</b>).</p>
Full article ">Figure 5
<p>Relative density of Ba<sub>1.03</sub>Ce<sub>0.5</sub>Zr<sub>0.4</sub>Y<sub>0.1</sub>O<sub>3−<span class="html-italic">δ</span></sub> with various ZnO concentrations sintered at different temperatures (<b>a</b>) and SEM images of Ba<sub>1.03</sub>Ce<sub>0.5</sub>Zr<sub>0.4</sub>Y<sub>0.1</sub>O<sub>3−<span class="html-italic">δ</span></sub> without (<b>b</b>) and with 1 wt% (<b>c</b>) ZnO sintered at 1300 °C for 10 h, reproduced with permission from Elsevier [<a href="#B115-processes-12-02278" class="html-bibr">115</a>]; BaCe<sub>0.8</sub>Zr<sub>0.1</sub>Y<sub>0.1</sub>O<sub>3−δ</sub> densification mechanism through ZnO.BaO eutectic formation reproduced with permission from Elsevier [<a href="#B116-processes-12-02278" class="html-bibr">116</a>] (<b>d</b>); thermogravimetric analysis (TGA) curve of pristine BaCe<sub>0.3</sub>Zr<sub>0.55</sub>Y<sub>0.15</sub>O<sub>3−δ</sub> [<a href="#B54-processes-12-02278" class="html-bibr">54</a>] in pure CO<sub>2</sub> environment (<b>e</b>), and TGA curve of BaCe<sub>0.55</sub>Zr<sub>0.3</sub>Y<sub>0.15</sub>O<sub>3−d</sub> and BaCe<sub>0.35</sub>Zr<sub>0.5</sub>Y<sub>0.15</sub>O<sub>3−d</sub> in CO<sub>2</sub> and air (<b>f</b>) [<a href="#B121-processes-12-02278" class="html-bibr">121</a>], replotted from original papers.</p>
Full article ">Figure 6
<p>The temperature dependence of the conductivity of different BCZY10 and BCZY27 samples under a wet reducing atmosphere (9% H<sub>2</sub>N<sub>2</sub>, P<sub>H2O</sub> = 0.015 atm) with (<b>a</b>) Co doping and (<b>b</b>) Ni doping [<a href="#B132-processes-12-02278" class="html-bibr">132</a>]; Arrhenius plots of the conductivities in 3% H<sub>2</sub>O of the BZCY, BZCY-2, BZCY-5, and BZCY-10 samples sintered at 1400 °C for 5 h, reproduced with permission from Elsevier [<a href="#B152-processes-12-02278" class="html-bibr">152</a>] (<b>c</b>).</p>
Full article ">Figure 7
<p>Various strategies that require further in-depth investigation to achieve BCZY densification at lower temperatures using sintering aids without any deterioration of protonic conductivity.</p>
Full article ">
18 pages, 4088 KiB  
Article
Enhancing the Performance of BaxMnO3 (x = 1, 0.9, 0.8 and 0.7) Perovskites as Catalysts for CO Oxidation by Decreasing the Ba Content
by Á. Díaz-Verde and M. J. Illán-Gómez
Nanomaterials 2024, 14(16), 1334; https://doi.org/10.3390/nano14161334 - 10 Aug 2024
Cited by 1 | Viewed by 1229
Abstract
Mixed oxides featuring perovskite-type structures (ABO3) offer promising catalytic properties for applications focused on the control of atmospheric pollution. In this work, a series of BaxMnO3 (x = 1, 0.9, 0.8 and 0.7) samples have been synthesized, characterized [...] Read more.
Mixed oxides featuring perovskite-type structures (ABO3) offer promising catalytic properties for applications focused on the control of atmospheric pollution. In this work, a series of BaxMnO3 (x = 1, 0.9, 0.8 and 0.7) samples have been synthesized, characterized and tested as catalysts for CO oxidation reaction in conditions close to that found in the exhausts of last-generation automotive internal combustion engines. All samples were observed to be active as catalysts for CO oxidation during CO-TPRe tests, with Ba0.7MnO3 (B0.7M) being the most active one, as it presents the highest amount of oxygen vacancies (which act as active sites for CO oxidation) and Mn (IV), which features the highest levels of reducibility and the best redox properties. B0.7M has also showcased a high stability during reactions at 300 °C, even though a slightly lower CO conversion is achieved during the second consecutive reaction cycle. This performance appears to be related to the decrease in the Mn (IV)/Mn (III) ratio. Full article
(This article belongs to the Special Issue Synthesis and Applications of Perovskite Nanocrystals)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of BxM samples.</p>
Full article ">Figure 2
<p>XPS spectra of the Mn 2p<sup>3/2</sup> (<b>a</b>), Mn 3p (<b>b</b>), Ba 3d<sup>5/2</sup> (<b>c</b>) and O 1s (<b>d</b>) transitions.</p>
Full article ">Figure 3
<p>H<sub>2</sub>-TPR consumption profiles for BxM samples, for Mn<sub>2</sub>O<sub>3</sub> and MnO<sub>2</sub> references (<b>a</b>), and for H<sub>2</sub> consumption (mL H<sub>2</sub> (g of cat)<sup>−1</sup>) (<b>b</b>).</p>
Full article ">Figure 4
<p>CO conversion profiles of the first (solid lines) and the second cycle (dotted lines) of CO-TPR tests for BxM samples.</p>
Full article ">Figure 5
<p>O<sub>2</sub>-TPD profiles for BxM samples.</p>
Full article ">Figure 6
<p>CO conversion profiles for BxM and 1% Pt/Al<sub>2</sub>O<sub>3</sub> samples in the 0.1% CO/1% O<sub>2</sub>/He (<b>a</b>); 1% CO/1% O<sub>2</sub>/He (<b>b</b>); and 1% CO/10% O<sub>2</sub>/He (<b>c</b>) reactant mixtures.</p>
Full article ">Figure 7
<p>CO conversion profiles of B0.7M at 300 °C in the 1% CO/He reactant mixture.</p>
Full article ">Figure 8
<p>XPS spectra of the Mn 2p<sub>3/2</sub> (<b>a</b>) and O 1s (<b>b</b>) transitions for the fresh and the spent B0.7M samples.</p>
Full article ">
15 pages, 8638 KiB  
Article
Effect of Rh Doping on Optical Absorption and Oxygen Evolution Reaction Activity on BaTiO3 (001) Surfaces
by Talgat M. Inerbaev, Aisulu U. Abuova, Zhadyra Ye. Zakiyeva, Fatima U. Abuova, Yuri A. Mastrikov, Maksim Sokolov, Denis Gryaznov and Eugene A. Kotomin
Molecules 2024, 29(11), 2707; https://doi.org/10.3390/molecules29112707 - 6 Jun 2024
Cited by 1 | Viewed by 885
Abstract
In the present work, we investigate the potential of modified barium titanate (BaTiO3), an inexpensive perovskite oxide derived from earth-abundant precursors, for developing efficient water oxidation electrocatalysts using first-principles calculations. Based on our calculations, Rh doping is a way of making [...] Read more.
In the present work, we investigate the potential of modified barium titanate (BaTiO3), an inexpensive perovskite oxide derived from earth-abundant precursors, for developing efficient water oxidation electrocatalysts using first-principles calculations. Based on our calculations, Rh doping is a way of making BaTiO3 absorb more light and have less overpotential needed for water to oxidize. It has been shown that a TiO2-terminated BaTiO3 (001) surface is more promising from the point of view of its use as a catalyst. Rh doping expands the spectrum of absorbed light to the entire visible range. The aqueous environment significantly affects the ability of Rh-doped BaTiO3 to absorb solar radiation. After Ti→Rh replacement, the doping ion can take over part of the electron density from neighboring oxygen ions. As a result, during the water oxidation reaction, rhodium ions can be in an intermediate oxidation state between 3+ and 4+. This affects the adsorption energy of reaction intermediates on the catalyst’s surface, reducing the overpotential value. Full article
(This article belongs to the Special Issue Chemistry of Materials for Energy and Environmental Sustainability)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) TiO<sub>2</sub>- and (<b>b</b>) BaO-terminated (001) surfaces of tetragonal BaTiO<sub>3</sub>.</p>
Full article ">Figure 2
<p>(<b>Left</b>): Top view of the outermost layer of TiO<sub>2</sub>-terminated (<b>a</b>) undoped and (<b>b</b>) Rh-doped surfaces. The numbers indicate the distance (Å) between the (<b>a</b>) Ti and (<b>b</b>) Rh atoms and the nearest surface oxygen atoms (O1). (<b>c</b>) Side view of a doped TiO<sub>2</sub>-terminated surface (Ba ions omitted); the numbers indicate the interatomic distance between the metal atoms (Ti: black, Rh: pink) and subsurface oxygen (O2). (<b>Right</b>): Top view of the two upper layers of BaO-terminated (<b>d</b>) undoped and (<b>e</b>) Rh-doped surfaces. Side view of a doped BaO-terminated surface (<b>f</b>).</p>
Full article ">Figure 3
<p>Total and partial densities of states for bare and doped TiO<sub>2</sub>- and BaO-terminated surfaces. Top: (<b>a</b>) undoped TiO<sub>2</sub>-terminated surface; (<b>b</b>) Rh-doped TiO<sub>2</sub>-terminated surface. The contribution of the surface nearest to the oxygen atoms of Rh, O(Rh), is highlighted. Bottom: (<b>c</b>) undoped BaO-terminated surface; (<b>d</b>) Rh-doped BaO-terminated surface. <span class="html-italic">E</span><sub>F</sub>: Fermi energy.</p>
Full article ">Figure 4
<p>Optical absorption of undoped and Rh-doped (<b>a</b>) TiO<sub>2</sub>- and (<b>b</b>) BaO-terminated surfaces. Black and blue lines correspond to dry and wet surfaces, respectively. The solid lines illustrate total optical absorption, while dashed and dotted lines correspond to the contributions of spin-up (UP) and spin-down (DW) electronic states. Orange lines refer to experimental data adapted from Ref. [<a href="#B32-molecules-29-02707" class="html-bibr">32</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Change in the arrangement of ions in the slab after Ba was replaced with Rh; (<b>b</b>) electronic DOS for relaxed slab; (<b>c</b>) optical absorption spectrum for the model investigated. Dashed and dotted lines represent optical absorption by spin-up and spin-down states. The solid line illustrates total absorption.</p>
Full article ">Figure 6
<p>Standard free energy diagram for the OER at zero potential (U = 0, dotted lines) and equilibrium potential for oxygen evolution (<span class="html-italic">U</span> = 1.23 V, solid lines) at pH = 0 and T = 298 K. Black and blue lines show data for dry and wet surfaces, respectively. Dashed lines correspond to the ideal catalyst.</p>
Full article ">Figure 7
<p>Equation (5) calculates the charge transfer between the TiO<sub>2</sub>-termiated catalyst surface and the intermediate reaction products. A side view of the surface of the top two layers is presented. OH adsorbed on (<b>a</b>) undoped and (<b>b</b>) Rh-doped surfaces; O adsorbed on (<b>c</b>) undoped and (<b>d</b>) Rh-doped surfaces; and HOO adsorbed on (<b>e</b>) undoped and (<b>f</b>) Rh-doped surfaces. The yellow and blue clouds indicate the isocontours of positive and negative values of the electron charge density, respectively.</p>
Full article ">Figure 8
<p>Equation (6) calculates the charge transfer between wet and dry TiO<sub>2</sub>-termiated catalyst surfaces. Top view of the upper layer of the (<b>a</b>) undoped and (<b>b</b>) Rh-doped surfaces; side view of the two upper layers of the (<b>c</b>) undoped and (<b>d</b>) Rh-doped surfaces. The yellow and blue clouds indicate the isocontours of positive and negative values of the electron charge density, respectively.</p>
Full article ">
18 pages, 31596 KiB  
Article
Synthesis of BaZrS3 and BaS3 Thin Films: High and Low Temperature Approaches
by Tim Freund, Sumbal Jamshaid, Milad Monavvar and Peter Wellmann
Crystals 2024, 14(3), 267; https://doi.org/10.3390/cryst14030267 - 9 Mar 2024
Cited by 2 | Viewed by 2132
Abstract
Current research efforts in the field of the semiconducting chalcogenide perovskites are directed towards the fabrication of thin films and subsequently determine their performance in the photovoltaic application. These efforts are motivated by the outstanding properties of this class of materials in terms [...] Read more.
Current research efforts in the field of the semiconducting chalcogenide perovskites are directed towards the fabrication of thin films and subsequently determine their performance in the photovoltaic application. These efforts are motivated by the outstanding properties of this class of materials in terms of stability, high absorption coefficient near the band edge and no significant health concerns compared to their halide counterparts. The approach followed here is to use stacked precursor layers and is adopted from other chalcogenide photovoltaic materials like the kesterites and chalcopyrites. The successful synthesis of BaZrS3 from stacked layers of BaS and Zr and annealing at high temperatures (~1100 °C) with the addition of elemental sulfur is demonstrated. However, the film shows the presence of secondary phases and a flawed surface. As an alternative to this, BaS3 could be used as precursor due to its low melting point of 554 °C. Previously, the fabrication of BaS3 films was demonstrated, but in order to utilize them in the fabrication of BaZrS3 thin films, their microstructure and processing are further improved in this work by reducing the synthesis temperature to 300 °C, resulting in a smoother surface. This work lays the groundwork for future research in the fabrication of chalcogenide perovskites utilizing stacked layers and BaS3. Full article
(This article belongs to the Special Issue Perovskites – New and Old Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration showing the general approach of stacked layers (<b>a</b>), the high temperature approach (<b>b</b>) and the BaS<sub>3</sub> based approach (<b>c</b>).</p>
Full article ">Figure 2
<p>Schematic of the oven setup used for sulfurization showing the positions of the various instruments.</p>
Full article ">Figure 3
<p>XRD of Sample 1 annealed at high temperature (1100 °C) showing a match with BZS and other secondary phases. The inlet is a macroscopic image of the same sample showing the ablated parts and the sub-optimal surface structure.</p>
Full article ">Figure 4
<p>Comparison of microscope images of previously fabricated BaS<sub>3</sub> samples ((<b>a</b>–<b>d</b>), reused with permission from [<a href="#B49-crystals-14-00267" class="html-bibr">49</a>]) and BaS<sub>3</sub> samples fabricated as part of this study (<b>e</b>–<b>h</b>). The microstructures of Samples 2 (<b>a</b>), 3 (<b>b</b>) and 4 (<b>c</b>,<b>d</b>) have larger defects and bulges than Samples 5 (<b>e</b>), 6 (<b>f</b>), 7 (<b>g</b>) and 8 (<b>h</b>). The different colors in (<b>e</b>–<b>h</b>) are not accurate to reality and are caused by differences in the white balance.</p>
Full article ">Figure 5
<p>SEM images of Sample 10 (<b>a</b>,<b>c</b>) and Sample 9 (<b>b</b>,<b>d</b>) showing the surface of BaS and BaS<sub>3</sub> films, respectively, in top (<b>a</b>,<b>b</b>) and cross-sectional (<b>c</b>,<b>d</b>) view.</p>
Full article ">Figure 6
<p>The XRD pattern (black) of Sample 12 annealed for 10 min shows a good match with the BaS<sub>2</sub> reference pattern in red (<b>a</b>), whereas Sample 13 annealed for 15 min matches with BaS<sub>3</sub> (<b>b</b>). The Temperature curves for Samples 12 and 9 annealed for 10 (<b>c</b>) and 30 min (<b>d</b>), respectively, demonstrate, that a temperature slightly above 300 °C is needed to obtain a BaS<sub>3</sub> film. The sample temperatures are displayed in blue, whereas the oven temperature is displayed in orange.</p>
Full article ">Figure 7
<p>(<b>a</b>) XRD spectrum of Sample 16 annealed in rapid thermal processing, showing good match with the BaS<sub>3</sub> reference pattern (red lines). (<b>b</b>) Temperature curve for the same sample (blue line) showing the rise in temperature up to 420 °C even after exertion of the sample after 1:30 min. The oven temperature (orange line) was set to 660 °C and only slightly drops during the process.</p>
Full article ">Figure 8
<p>The microscope images of rapidly annealed BaS<sub>3</sub> film samples show a similar microstructure to those annealed for 30 min. (<b>a</b>,<b>b</b>) show Sample 15, (<b>c</b>,<b>d</b>) show Samples 16 and 17, respectively.</p>
Full article ">Figure A1
<p>BaS film (Sample 10, (<b>a</b>)) and BaS<sub>3</sub> film (Sample 9, (<b>b</b>)) cross-sections additionally used for the determination of the film thicknesses.</p>
Full article ">Figure A2
<p>Areas used for EDX measurements (<b>a</b>,<b>b</b>) and the resulting spectra (<b>c</b>,<b>d</b>) for a BaS film (Sample 10, (<b>a</b>,<b>c</b>)) and a BaS<sub>3</sub> film (Sample 9, (<b>b</b>,<b>d</b>)).</p>
Full article ">Figure A3
<p>XRD patterns of BaS3 Samples 15, 17 and 18 annealed in the rapid processing route (<b>a</b>) and their respective temperature curves with the sample temperature in blue and the oven temperature in orange (<b>b</b>–<b>d</b>).</p>
Full article ">
12 pages, 4122 KiB  
Article
Ferroelectric Phase Transition in Barium Titanate Revisited with Ab Initio Molecular Dynamics
by Christian Ludt, Dirk C. Meyer and Matthias Zschornak
Materials 2024, 17(5), 1023; https://doi.org/10.3390/ma17051023 - 23 Feb 2024
Cited by 1 | Viewed by 1715
Abstract
The ferroelectric phase transition of the perovskite barium titanate as well as its technical importance regarding the switching of respective polar properties is well known and has been thoroughly studied, both experimentally and on theoretical grounds. While details about the phase diagram as [...] Read more.
The ferroelectric phase transition of the perovskite barium titanate as well as its technical importance regarding the switching of respective polar properties is well known and has been thoroughly studied, both experimentally and on theoretical grounds. While details about the phase diagram as well as transition temperatures are experimentally well known, the theoretical approaches still face difficulties in contributing a detailed description of these phase transitions. Within this work, a new methodological approach is introduced to revisit the ferroelectric phase transition with first-principles methods. With the chosen ab initio molecular dynamics (AIMD) method in combination with the applied NpT ensemble, we are able to join the accuracy of density functional theory (DFT) with ambient conditions, realized using a thermostat and barostat in an MD simulation. The derived phase diagram confirms recent corrections in the theoretical models and reproduces the phase boundary pressure dependence of TC. In conclusion of the statistical atomistic dynamics, the nature of the transition can be described in a more detailed way. In addition, this work paves the way towards locally patterned piezoelectrica by means of acoustic standing waves as well as piezoelectrically induced acoustic resonators. Full article
Show Figures

Figure 1

Figure 1
<p>AIMD simulation of the BaTiO<sub>3</sub> 2 × 2 × 2 supercell at 600 (<b>a</b>), 700 (<b>b</b>), 800 (<b>c</b>), and 2000 K (<b>d</b>) and 0 GPa external pressure: The area of the first picosecond is highlighted gray to mark the relaxation time, which is needed to approach the thermodynamic equilibrium. Lattice parameters <span class="html-italic">a, b,</span> and <span class="html-italic">c</span> given in red, blue and green, indicate the tetragonal phase in the cases of 600, 700, and 800 K. At 2000 K, the lattice parameters approach each other, indicating a cubic phase. The mean displacements of titanium ions in comparison to the respective oxygen planes in <span class="html-italic">z</span>-direction <math display="inline"><semantics> <mrow> <mo>⌀</mo> <mo>Δ</mo> <msub> <mi>z</mi> <mrow> <mtext>Ti-O</mtext> </mrow> </msub> </mrow> </semantics></math> show different behavior depending on temperature: In case of the 600 K simulation, <math display="inline"><semantics> <mrow> <mo>⌀</mo> <mo>Δ</mo> <msub> <mi>z</mi> <mrow> <mtext>Ti-O</mtext> </mrow> </msub> </mrow> </semantics></math> is stable at approximately 23 pm, reflecting that the phase is polar, while at 700 K a polarization switching occurs, which is given by the changed sign of <math display="inline"><semantics> <mrow> <mo>⌀</mo> <mo>Δ</mo> <msub> <mi>z</mi> <mrow> <mtext>Ti-O</mtext> </mrow> </msub> </mrow> </semantics></math>. At 800 K, the behavior is alike, but for a short time a centrosymmetric plateau can be established. In the case of 2000 K, the displacement parameter <math display="inline"><semantics> <mrow> <mo>⌀</mo> <mo>Δ</mo> <msub> <mi>z</mi> <mrow> <mtext>Ti-O</mtext> </mrow> </msub> </mrow> </semantics></math> switches frequently.</p>
Full article ">Figure 2
<p>AIMD simulation of BaTiO<sub>3</sub> 2 × 2 × 2 supercells at 400 K for different Hubbard U values (10 eV (<b>a</b>), 12 eV (<b>b</b>) and 14 eV (<b>c</b>)). While the simulation at 10 eV shows tetragonal behavior over the whole simulation time, regarding lattice parameters (<span class="html-italic">a</span> in red, <span class="html-italic">b</span> in blue, <span class="html-italic">c</span> in green) as well as the displacements, cubic behavior is indicated at 12 eV and 14 eV.</p>
Full article ">Figure 3
<p>AIMD simulation of the BaTiO<sub>3</sub> 3 × 3 × 3 supercell at 50 (<b>a</b>), 150 (<b>b</b>), 300 (<b>c</b>), and 400 K (<b>d</b>) and 0 GPa external pressure: The noise in the monitored quantities is reduced significantly compared to the 2 × 2 × 2 supercell. Lattice parameters <span class="html-italic">a</span>, <span class="html-italic">b</span>, and <span class="html-italic">c</span> given in red, blue and green, indicate the tetragonal phase for 50 and 150 K, while at 150 K, short cubic-like behavior appears. For the elevated temperatures of 300 and 400 K, the approaching lattice parameters indicate a predominantly cubic unit cell. The mean displacement <math display="inline"><semantics> <mrow> <mo>⌀</mo> <mo>Δ</mo> <msub> <mi>z</mi> <mrow> <mtext>Ti-O</mtext> </mrow> </msub> </mrow> </semantics></math> reflects a polar phase at 50 K. Switching begins to be established at 150 K and is more frequent as the temperature increases up to 400 K. Here, a rather constant rapid switching indicates the first appearance of a centrosymmetric phase.</p>
Full article ">Figure 4
<p>AIMD simulation of the BaTiO<sub>3</sub> 4 × 4 × 4 supercell at 100 (<b>a</b>), 350 (<b>b</b>), and 400 K (<b>c</b>) and 0 GPa external pressure: Lattice parameters <span class="html-italic">a, b, c</span> given in red, blue and green, indicate the tetragonal phase for 100 and 350 K. For the elevated temperature of 400 K, the approaching lattice parameters indicate a cubic unit cell. The mean displacement <math display="inline"><semantics> <mrow> <mo>⌀</mo> <mo>Δ</mo> <msub> <mi>z</mi> <mrow> <mtext>Ti-O</mtext> </mrow> </msub> </mrow> </semantics></math> reflects a polar phase at 100 K and 350 K. At 400 K, a rather constant rapid switching indicates the first appearance of a centrosymmetric phase.</p>
Full article ">Figure 5
<p>AIMD simulation of the BaTiO<sub>3</sub> 3 × 3 × 3 supercell under the influence of chosen temperatures and external pressures: in (<b>a</b>) at 100 K and 2 GPa, (<b>b</b>) at 100 K and 3 GPa, (<b>c</b>) at 150 K and 5 GPa, (<b>d</b>) at 400 K at 10 GPa. Lattice parameters <span class="html-italic">a, b,</span> and <span class="html-italic">c</span> given in red, blue and green, indicate the cubic phase for all conditions. The mean displacement <math display="inline"><semantics> <mrow> <mo>⌀</mo> <mo>Δ</mo> <msub> <mi>z</mi> <mrow> <mtext>Ti-O</mtext> </mrow> </msub> </mrow> </semantics></math> reflects polar character for 100 K and 2 GPa, while for 100 K at 3 GPa the behavior represents a semi-state. For the other conditions, centrosymmetry is observed.</p>
Full article ">Figure 6
<p>Phase diagram of BaTiO<sub>3</sub>. The colored stars display the AIMD results for ferroelectric phase in blue and paraelectric phase in red, respectively. Mixed stars represent MD calculations, which show a transition state. The solid line indicates the tetragonal to cubic phase transition proposed experimentally by Ishidate et al. [<a href="#B11-materials-17-01023" class="html-bibr">11</a>]. The circles represent theoretical calculations by Íniguez and Vanderbilt [<a href="#B50-materials-17-01023" class="html-bibr">50</a>], while solid circles give classical Monte Carlo simulations, and open circles are generated using the path-integral quantum Monte Carlo (PI-QMC) technique. Colored backgrounds emphasize the respective regions in the phase diagram based on our calculated data.</p>
Full article ">Figure 7
<p>AIMD simulation of the BaTiO<sub>3</sub> 4 × 4 × 4 supercell under the influence of chosen temperatures and external pressures: in (<b>a</b>) at 100 K and 7 GPa, (<b>b</b>) at 200 K and 1 GPa, and (<b>c</b>) at 300 K and 1 GPa. Lattice parameters <span class="html-italic">a, b,</span> and <span class="html-italic">c</span> given in red, blue and green, indicate the cubic phase for 100 K and 7 GPa. The mean displacement <math display="inline"><semantics> <mrow> <mo>⌀</mo> <mo>Δ</mo> <msub> <mi>z</mi> <mrow> <mtext>Ti-O</mtext> </mrow> </msub> </mrow> </semantics></math> reflects polar character for 200 K and 1 GPa, while for 300 K at 1 GPa the behavior represents the transition state. For the other conditions, centrosymmetry is observed.</p>
Full article ">
16 pages, 3231 KiB  
Article
Construction of Spinel/Perovskite Heterojunction for Boosting Photocatalytic Performance for Polyacrylamide
by Qinghan Zhu, Yuxue Luo, Ke Yang, Guangbo Che, Haiwang Wang and Jian Qi
Catalysts 2023, 13(11), 1424; https://doi.org/10.3390/catal13111424 - 8 Nov 2023
Cited by 4 | Viewed by 1244
Abstract
The use of photocatalytic technology to degrade polyacrylamide in crude oil extraction wastewater is a promising approach, but there have been few reports so far. In this study, ZnFe2O4/Ba0.7Sr0.3TiO3 heterogeneous composite materials of a [...] Read more.
The use of photocatalytic technology to degrade polyacrylamide in crude oil extraction wastewater is a promising approach, but there have been few reports so far. In this study, ZnFe2O4/Ba0.7Sr0.3TiO3 heterogeneous composite materials of a spinel/perovskite type with different proportions were synthesized. The composite materials with 31% ZnFe2O4 content exhibited a maximum polyacrylamide degradation efficiency of 46.54%, which demonstrated the unique role of the spinel/perovskite heterogeneous structure. When Ag nanoparticles were grown in situ on the surface of ZnFe2O4/Ba0.7Sr0.3TiO3, the photocatalytic degradation efficiency reached 81.28%. The main reason was that the introduction of Ag nanoparticles not only increased the active sites and enhanced light absorption capacity but also accelerated the separation of photo-generated charges. This work provides new ideas for the construction of spinel/perovskite heterogeneous composite materials and has reference significance for the application of photocatalytic degradation in the treatment of wastewater-containing polymers. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of ZnFe<sub>2</sub>O<sub>4</sub>, Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub>, and ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub> with different components.</p>
Full article ">Figure 2
<p>FT-IR spectra of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub> (ZFO wt% = 31) sample, (<b>a</b>) before calcination and (<b>b</b>) after calcinations at 1100 °C.</p>
Full article ">Figure 3
<p>EDS spectra of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub> (ZFO wt% = 31).</p>
Full article ">Figure 4
<p>XPS spectra of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub> (ZFO wt% = 31), (<b>a</b>) Ba 3d, (<b>b</b>) Fe 2p, (<b>c</b>) O 1s, (<b>d</b>) Sr 3d, (<b>e</b>) Ti 2p and (<b>f</b>) Zn 2p.</p>
Full article ">Figure 5
<p>SEM images of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub> with different compositions, (<b>a</b>,<b>b</b>) ZFO wt% = 31 and (<b>c</b>,<b>d</b>) ZFO wt% = 35.</p>
Full article ">Figure 6
<p>PL spectra of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub> with different ZFO/BST ratios, (<b>a</b>) photoluminescence spectra; (<b>b</b>) partial enlarged view of (<b>a</b>).</p>
Full article ">Figure 7
<p>(<b>a</b>) Photodegradation of PAM aqueous solution with different components of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub>; (<b>b</b>) total degradation efficiency of different components.</p>
Full article ">Figure 8
<p>(<b>a</b>) The photocatalytic degradation performance of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub>/Ag catalyst with different Ag contents for PAM and (<b>b</b>) comparison of corresponding degradation efficiencies. (<b>c</b>) The photocatalytic degradation performance of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub>/Ag under different catalyst dosage for PAM and (<b>d</b>) comparison of corresponding degradation efficiencies. (<b>e</b>) The photocatalytic degradation performance of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub>/Ag catalyst under different initial concentration of PAM and (<b>f</b>) comparison of corresponding degradation efficiencies. (<b>g</b>) The photocatalytic degradation performance of ZnFe<sub>2</sub>O<sub>4</sub>/Ba<sub>0.7</sub>Sr<sub>0.3</sub>TiO<sub>3</sub>/Ag catalyst under different pH values and (<b>h</b>) comparison of corresponding degradation efficiencies.</p>
Full article ">
23 pages, 62179 KiB  
Article
Oxide Strontium-Barium Perovskites Ceramics: Examinations of Structural Phase Transitions and Potential Application as Oxygen Carriers
by Ewelina Ksepko, Rafal Lysowski and Miratul Alifah
Processes 2023, 11(7), 2144; https://doi.org/10.3390/pr11072144 - 18 Jul 2023
Viewed by 1380
Abstract
The structural properties of selected (Ba1−xSrx)PbO3 ceramics were examined at 14–1148 K using X-ray powder diffraction (XRD). These materials are attractive due to their variety of applications, such as, for example, high-temperature thermoelectric energy conversion. Attention was paid [...] Read more.
The structural properties of selected (Ba1−xSrx)PbO3 ceramics were examined at 14–1148 K using X-ray powder diffraction (XRD). These materials are attractive due to their variety of applications, such as, for example, high-temperature thermoelectric energy conversion. Attention was paid to this paper as a continuation of the previous examinations of higher Sr2+ concentrations. The type of perovskite distortion and temperatures of the structural phase transitions (SPTs) were determined from the splitting of certain pseudocubic lines. At this point, for example (Ba0.3Sr0.7)PbO3 showed three temperature-induced SPTs. When the amount of Sr increased in the samples, no phase transition was observed, which is contrary to the data previously demonstrated in the literature. The quality of the ceramics was examined by scanning electron microscopy-energy dispersion X-ray spectroscopy (SEM-EDS), demonstrating their homogeneity and uniform elements dispersion. As a result of profound crystal investigations, confirmed by thermogravimetric analysis and quadrupole mass spectroscopy (TGA-QMS), a phase diagram was prepared for the (Ba1−xSrx)PbO3 system based on our former and recent study. Also, the investigation of a new application for the (Ba1−xSrx)PbO3 family is presented in this paper for the first time. The TGA analysis was conducted on Illinois#6 hard coal to evaluate the capability of perovskites to be used in the chemical looping combustion (CLC) process in a range of temperatures 1073–1173 K. Due to its thermal stability and reactivity, Ba0.9Sr0.1PbO3 is the material with the greatest potential to be applied as an oxygen carrier. The combination of strontium and barium offers encouraging results compared to the pure barium and strontium lead oxide perovskites. Full article
(This article belongs to the Special Issue Advances in Ceramic Processing and Application of Ceramic Materials)
Show Figures

Figure 1

Figure 1
<p>SEM images of (Ba<sub>1−x</sub>Sr<sub>x</sub>)PbO<sub>3</sub> series surface at 250× magnification.</p>
Full article ">Figure 2
<p>SEM image of SrPbO<sub>3</sub> powders registered at 800× magnification (<b>a</b>), example of distribution maps of Sr, Pb and O ions (<b>b</b>).</p>
Full article ">Figure 3
<p>Splitting of the main reflections <span class="html-italic">{220}<sub>c</sub></span> and <span class="html-italic">{222}<sub>c</sub></span> in the function of temperature for SrPbO<sub>3</sub>.</p>
Full article ">Figure 4
<p>Superstructure lines observed for SrPbO<sub>3</sub>.</p>
Full article ">Figure 5
<p>SrPbO<sub>3</sub> cell with <span class="html-italic">Pbnm</span> orthorhombic symmetry at 298 K. Projections on <span class="html-italic">[101]</span>, <span class="html-italic">[011]</span>, and <span class="html-italic">[110]</span> planes are shown. Red spheres—oxygen, grey spheres—lead, green spheres—strontium.</p>
Full article ">Figure 6
<p>Rietveld analysis results of X-ray powder diffraction pattern for SrPbO<sub>3</sub> collected at RT.</p>
Full article ">Figure 7
<p>Pseudocubic cell parameters in the function of temperature for SrPbO<sub>3</sub>. The inlet shows the temperature evolution of the cell volume.</p>
Full article ">Figure 8
<p>Rietveld analysis results of X-ray powder diffraction pattern for SrPbO<sub>3</sub> collected at 14 K.</p>
Full article ">Figure 9
<p>Splitting of the main reflections <span class="html-italic">{200}<sub>c</sub></span> and <span class="html-italic">{222}<sub>c</sub></span> for (Ba<sub>0.2</sub>Sr<sub>0.8</sub>)PbO<sub>3</sub> at RT and 1093 K.</p>
Full article ">Figure 10
<p>Pseudocubic cell parameters in the function of temperature for (Ba<sub>0.2</sub>Sr<sub>0.8</sub>)PbO<sub>3</sub>.</p>
Full article ">Figure 11
<p>Splitting of the main reflections <span class="html-italic">{220}<sub>c</sub></span> and <span class="html-italic">{200}<sub>c</sub></span> in different phases for (Ba<sub>0.3</sub>Sr<sub>0.7</sub>)PbO<sub>3</sub>.</p>
Full article ">Figure 12
<p>Pseudocubic cell parameters in the function of temperature for (Ba<sub>0.3</sub>Sr<sub>0.7</sub>)PbO<sub>3</sub>.</p>
Full article ">Figure 13
<p>The pseudocubic cell parameters in function of composition for (Ba<sub>1−x</sub>Sr<sub>x</sub>)PbO<sub>3</sub> system.</p>
Full article ">Figure 14
<p>Phase diagram for (Ba<sub>1−x</sub>Sr<sub>x</sub>)PbO<sub>3</sub> system.</p>
Full article ">Figure 15
<p>STA analysis for selected examples of BaPbO<sub>3</sub> and SrPbO<sub>3</sub> perovskites.</p>
Full article ">Figure 16
<p>TGA-QMS analysis for selected samples of SrPbO<sub>3</sub>, Ba<sub>0.9</sub>Sr<sub>0.1</sub>PbO<sub>3</sub> and BaPbO<sub>3</sub> perovskites.</p>
Full article ">
15 pages, 6689 KiB  
Article
Enhanced Photocatalytic and Anticancer Activity of Zn-Doped BaTiO3 Nanoparticles Prepared through a Green Approach Using Banana Peel Extract
by Maqusood Ahamed and M. A. Majeed Khan
Catalysts 2023, 13(6), 985; https://doi.org/10.3390/catal13060985 - 8 Jun 2023
Cited by 9 | Viewed by 2366
Abstract
Perovskite barium titanate (BaTiO3) has received a lot of interest due to its extraordinary dielectric and ferroelectric properties, along with its moderate biocompatibility. Here, we investigated how Zn doping tuned the physicochemical characteristics, photocatalytic activity, and anticancer potential of BaTiO3 [...] Read more.
Perovskite barium titanate (BaTiO3) has received a lot of interest due to its extraordinary dielectric and ferroelectric properties, along with its moderate biocompatibility. Here, we investigated how Zn doping tuned the physicochemical characteristics, photocatalytic activity, and anticancer potential of BaTiO3 nanoparticles synthesized from banana peel extract. XRD, TEM, SEM, EDS, XPS, BET, Raman, and PL were utilized to characterize the as-synthesized pure and Zn (1 and 3 mol%)-doped BaTiO3 nanoparticles. All of the synthesized samples showed evidence of the BaTiO3 tetragonal phase, and the XRD patterns of the Zn-doped BaTiO3 nanoparticles showed the presence of a Zn peak. The particle size of BaTiO3 decreased with increasing levels of Zn doping without morphological changes. After Zn doping, the PL intensity of BaTiO3 decreased, suggesting a lower electron–hole recombination rate. BET analysis found that the surface area of Zn-doped BaTiO3 nanoparticles was higher than that of pure BaTiO3. Under visible irradiation, the photocatalytic activity of pure and Zn-doped BaTiO3 nanoparticles was compared, and a remarkable 85% photocatalytic activity of Zn (3%)-doped BaTiO3 nanoparticles was measured. As a result, Zn-doped BaTiO3 nanoparticles are recognized as excellent photocatalysts for degrading organic pollutants. According to cytotoxicity data, Zn (3%)-doped BaTiO3 nanoparticles display four-fold greater anticancer activity against human lung carcinoma (A549) than pure BaTiO3 nanoparticles. It was also observed that Zn-doped BaTiO3 nanoparticles kill cancer cells by increasing the intracellular level of reactive oxygen species. Furthermore, compared to pure BaTiO3, the Zn-doped BaTiO3 nanostructure showed better cytocompatibility in non-cancerous human lung fibroblasts (IMR90). The Zn-doped BaTiO3 nanoparticles have a reduced particle size, increased surface area, and a lower electron–hole recombination rate, which are highly beneficial for enhanced photocatalytic and anticancer activity. Overall, current data showed that green-fabricated Zn-BaTiO3 nanoparticles have superior photocatalytic and anticancer effects along with improved biocompatibility compared to those of pure BaTiO3. This work underlines the significance of utilizing agricultural waste (e.g., fruit peel) for the fabrication of BaTiO3-based nanostructures, which hold great promise for biomedical and environmental applications. Full article
Show Figures

Figure 1

Figure 1
<p>XRD spectra of pure and Zn-doped BaTiO<sub>3</sub> nanoparticles (<b>A</b>) and peak shifts (<b>B</b>).</p>
Full article ">Figure 1 Cont.
<p>XRD spectra of pure and Zn-doped BaTiO<sub>3</sub> nanoparticles (<b>A</b>) and peak shifts (<b>B</b>).</p>
Full article ">Figure 2
<p>TEM particle size distribution (<b>A</b>–<b>C</b>), TEM micrographs of low (<b>D</b>–<b>F</b>) and high (<b>G</b>–<b>I</b>) resolutions of pure BaTiO<sub>3</sub>, 1% Zn–BaTiO<sub>3</sub>, and 3% Zn–BaTiO<sub>3</sub> nanoparticles, respectively.</p>
Full article ">Figure 3
<p>SEM micrographs of pure BaTiO<sub>3</sub> (<b>A</b>), 1% Zn–BaTiO<sub>3</sub> (<b>B</b>), and 3% Zn–BaTiO<sub>3</sub> (<b>C</b>) nanoparticles. (<b>D</b>) Elemental composition of 3% Zn–BaTiO<sub>3</sub> nanoparticles analyzed via EDS.</p>
Full article ">Figure 4
<p>SEM micrograph (<b>A</b>) and elemental mapping (<b>B</b>) of 3% Zn–BaTiO<sub>3</sub> nanoparticles.</p>
Full article ">Figure 5
<p>XPS survey spectrum of Zn (3%)-doped BaTiO<sub>3</sub> nanoparticles (<b>A</b>), high resolution signal of Ba 3d (<b>B</b>), Ti 2p (<b>C</b>), Zn 2p (<b>D</b>), and O 1s (<b>E</b>).</p>
Full article ">Figure 6
<p>BET adsorption–desorption isotherms for pure BaTiO<sub>3</sub> (<b>A</b>) and Zn (3%)-doped BaTiO<sub>3</sub> nanoparticles (<b>B</b>). Insets represent the pore diameter distribution.</p>
Full article ">Figure 6 Cont.
<p>BET adsorption–desorption isotherms for pure BaTiO<sub>3</sub> (<b>A</b>) and Zn (3%)-doped BaTiO<sub>3</sub> nanoparticles (<b>B</b>). Insets represent the pore diameter distribution.</p>
Full article ">Figure 7
<p>Raman spectroscopy of pure and Zn-doped BaTiO<sub>3</sub> nanoparticles.</p>
Full article ">Figure 8
<p>Photoluminescence spectra of pure and Zn-doped BaTiO<sub>3</sub> nanoparticles.</p>
Full article ">Figure 9
<p>Time-dependent absorption spectra of MB dye solution under visible-light irradiation with pure BaTiO<sub>3</sub> (<b>A</b>), 1% Zn–BaTiO<sub>3</sub> (<b>B</b>), 3% Zn–BaTiO<sub>3</sub> (<b>C</b>), and photocatalytic degradation (C/C<sub>o</sub> versus time plot) of MB dye solution with the same nanoparticles (<b>D</b>).</p>
Full article ">Figure 10
<p>Anticancer activity (<b>A</b>) and ROS generation (<b>B</b>) of pure and Zn-doped BaTiO<sub>3</sub> nanoparticles in human lung cancer A549 cells. Cytocompatibility of the same nanoparticles in non-cancerous human lung fibroblasts (IMR90) (<b>C</b>). The results are shown as the mean and standard deviation of three separate experiments (<span class="html-italic">n</span> = 3). * denotes significant difference from the control (<span class="html-italic">p</span> &lt; 0.05 level). GraphPad Prism (version 6.05) was used for the statistical analysis.</p>
Full article ">Figure 11
<p>A graphical illustration of the green synthesis of Zn-doped BaTiO<sub>3</sub> nanoparticles from banana peel extract.</p>
Full article ">
15 pages, 6959 KiB  
Article
Yttrium and Niobium Elements Co-Doping and the Formation of Double Perovskite Structure Ba2YNbO6 in BCZT
by Runyu Mao, Deyi Zheng, Qiyun Wu, Yuying Wang and Chang Liu
Materials 2023, 16(11), 4044; https://doi.org/10.3390/ma16114044 - 29 May 2023
Viewed by 1723
Abstract
The (Ba0.85Ca0.15) (Ti0.90Zr0.10)O3 + x Y3+ + x Nb5+ (abbreviated as BCZT-x(Nb + Y), x = 0 mol%, 0.05 mol%, 0.1 mol%, 0.2 mol%, 0.3 mol%) lead-free piezoceramics samples were [...] Read more.
The (Ba0.85Ca0.15) (Ti0.90Zr0.10)O3 + x Y3+ + x Nb5+ (abbreviated as BCZT-x(Nb + Y), x = 0 mol%, 0.05 mol%, 0.1 mol%, 0.2 mol%, 0.3 mol%) lead-free piezoceramics samples were prepared by a traditional solid-state sintering method. And the effects of Yttrium and Niobium elements (Y3+ and Nb5+) co-doping on the defect, phase and structure, microstructure, and comprehensive electrical properties have been investigated. Research results show that the Y and Nb elements co-doping can dramatically enhance piezoelectric properties. It is worth noting that XPS defect chemistry analysis, XRD phase analysis and TEM results together show that a new phase of double perovskite structure Barium Yttrium Niobium Oxide (Ba2YNbO6) is formed in the ceramic, and the XRD Rietveld refinement and TEM results show the coexistence of the R-O-T phase. Both these two reasons together lead to significant performance improvements of piezoelectric constant (d33) and planar electro-mechanical coupling coefficient (kp). The functional relation between temperature and dielectric constant testing results present that the Curie temperature increases slightly, which shows the same law as the change of piezoelectric properties. The ceramic sample reaches an optimal performance at x = 0.1% of BCZT-x(Nb + Y), where d33 = 667 pC/N, kp = 0.58, εr = 5656, tanδ = 0.022, Pr = 12.8 μC/cm2, EC = 2.17 kV/cm, TC =92 °C, respectively. Therefore, they can be used as potential alternative materials to lead based piezoelectric ceramics. Full article
(This article belongs to the Special Issue Piezoelectric and Ferroelectric Materials)
Show Figures

Figure 1

Figure 1
<p>(<bold>a</bold>) the XPS results of BCZT and BCZT-<italic>x</italic>(Nb + Y) samples. (<bold>b</bold>–<bold>e</bold>) XPS spectra of O 1s peaks binding states for BCZT-<italic>x</italic>(Nb + Y).</p>
Full article ">Figure 2
<p>(<bold>a1</bold>–<bold>a3</bold>) XRD patterns of BCZT-<italic>x</italic>(Nb + Y) ceramics. (<bold>b1</bold>,<bold>b2</bold>) Ba<sub>2</sub>YNbO<sub>6</sub> Schematic diagram of double perovskite structure (<bold>c</bold>) XRD Synchrotron Rietveld refinement of the <italic>x</italic> = 0.1 mol% ceramics. (<bold>d</bold>) phase fractions of BCZT-<italic>x</italic>(Nb + Y) ceramics. (<bold>e</bold>) c/a of BCZT-0.1 mol%(Nb + Y) ceramics from the Rietveld refinement.</p>
Full article ">Figure 3
<p>(<bold>a1</bold>–<bold>a5</bold>) SEM microscopic topography of BCZT-<italic>x</italic>(Nb + Y) ceramics, (<bold>b1</bold>–<bold>b5</bold>) grain size distribution of ceramics, (<bold>c1</bold>–<bold>c5</bold>) TEM of x = 0.1 mol% ceramics, (<bold>c1</bold>) topography, (<bold>c2</bold>) selected region electron diffraction, (<bold>c3</bold>) high-resolution lattice fringes. (<bold>c4</bold>,<bold>c5</bold>) fast Fourier transform of (<bold>c3</bold>). (<bold>d</bold>) density for BCZT and BCZT-<italic>x</italic>(Nb + Y). (<bold>e</bold>) average grain size distribution of BCZT-<italic>x</italic>(Nb + Y) ceramics.</p>
Full article ">Figure 4
<p>(<bold>a1</bold>−<bold>a5</bold>) P-E hysteresis loop of BCZT-<italic>x</italic>(Nb + Y) ceramic at 3 kV, 10 Hz and room temperature. (<bold>b</bold>) The <italic>P</italic><sub>r</sub> and <italic>E</italic><sub>C</sub>. (<bold>c</bold>) <italic>d</italic><sub>33</sub> and <italic>k</italic><sub>p</sub>, (<bold>d</bold>) <italic>ε</italic><sub>r</sub> and tan<italic>δ</italic> of BCZT-<italic>x</italic>(Nb + Y) ceramic samples. (<bold>e</bold>) Comparison of the <italic>d</italic><sub>33</sub> and <italic>T</italic><sub>C</sub> in reported BCZT-based piezoelectric ceramics in this work [<xref ref-type="bibr" rid="B4-materials-16-04044">4</xref>,<xref ref-type="bibr" rid="B6-materials-16-04044">6</xref>,<xref ref-type="bibr" rid="B7-materials-16-04044">7</xref>,<xref ref-type="bibr" rid="B8-materials-16-04044">8</xref>,<xref ref-type="bibr" rid="B9-materials-16-04044">9</xref>,<xref ref-type="bibr" rid="B10-materials-16-04044">10</xref>].</p>
Full article ">Figure 5
<p>(<bold>a</bold>–<bold>f</bold>) The relationship between temperature and dielectric constant of BCZT-<italic>x</italic>(Nb + Y) ceramic samples.</p>
Full article ">Figure 6
<p>(<bold>a–f</bold>) The 1/<italic>ε</italic><sub>r</sub> and temperature for the BCZT-<italic>x</italic>(Nb + Y) samples at 100 kHz. Insets is ln(1/<italic>ε</italic> − 1/<italic>ε<sub>m</sub></italic>)/ln(<italic>T</italic> − <italic>T</italic><sub>m</sub>).</p>
Full article ">
14 pages, 2665 KiB  
Article
Highly Conductive Cerium- and Neodymium-Doped Barium Zirconate Perovskites for Protonic Ceramic Fuel Cells
by Serdar Yilmaz, Bekir Kavici, Prakash Ramakrishnan, Cigdem Celen and Bahman Amini Horri
Energies 2023, 16(11), 4318; https://doi.org/10.3390/en16114318 - 25 May 2023
Cited by 2 | Viewed by 1363
Abstract
The rare-earth-doped zirconia-based solid electrolytes have gained significant interest in protonic ceramic fuel cell (PCFC) applications due to their high ionic conductivity. However, these solid electrolytes are susceptible to low conductivity and chemical stability at low operating temperatures, which are of interest in [...] Read more.
The rare-earth-doped zirconia-based solid electrolytes have gained significant interest in protonic ceramic fuel cell (PCFC) applications due to their high ionic conductivity. However, these solid electrolytes are susceptible to low conductivity and chemical stability at low operating temperatures, which are of interest in commercializing ceramic fuel cells. Thus, tailoring the structural properties of these electrolytes towards gaining high ionic conductivity at low/intermediate temperatures is crucial. In this study, Ce (cerium) and Nd (neodymium) co-doped barium zirconate perovskites, BaZr(0.80-x-y)CexNdyY0.10Yb0.10O3-δ (BZCNYYO) of various doping fractions (x, y: 0, 0.5, 0.10, 0.15), were synthesized (by the Pechini method) to systematically analyze their structural and conductivity properties. The X-ray diffraction patterns showed a significant lattice strain, and the stress inferences for each co-doped BZCNYYO sample were compared with Nd-cation-free reference samples, BaZrO3 and BaZr(0.80-x-y-z)CexYyYbzO3-δ (x: 0, 0.70; y: 0.20, 0.10; z: 0, 0.10). The comparative impedance investigation at low-to-intermediate temperatures (300–700 °C) showed that BaZr0.50Ce0.15Nd0.15Y0.10Yb0.10O3-δ offers the highest lattice strain and stress characteristics with an ionic conductivity (σ) of 0.381 mScm−1 at 500 °C and activation energy (Ea) of 0.47 eV. In addition, this σ value was comparable to the best reference sample BaZr0.10Ce0.70Y0.10Yb0.10O3-δ (0.404 mScm−1) at 500 °C, and it outperformed all the reference samples when the set temperature condition was ≥600 °C. The result of this study suggests that Ce- and Nd-doped BZCNYYO solid electrolytes will be a specific choice of interest for developing intermediate-temperature PCFC applications with high ionic conductivity. Full article
(This article belongs to the Section D2: Electrochem: Batteries, Fuel Cells, Capacitors)
Show Figures

Figure 1

Figure 1
<p>The XRD patterns of the BZCNYYO samples with various Ce dopants (Ce<span class="html-italic"><sub>x</sub></span> = 0, 0.05, 0.10, 0.15 mol) and Nd<span class="html-italic"><sub>y</sub></span> dopant concentration for (<b>a</b>) <span class="html-italic">y</span> = 0 mol, (<b>b</b>) <span class="html-italic">y</span> = 0.05, (<b>c</b>) <span class="html-italic">y</span> = 0.10, and (<b>d</b>) <span class="html-italic">y</span> = 0.15 mol.</p>
Full article ">Figure 2
<p>The XRD patterns of the reference samples: (<b>a</b>) A1 and A2; (<b>b</b>) A3, A4, and A5.</p>
Full article ">Figure 3
<p>Relationship between lattice parameters and (<b>a</b>) Ce-dopant concentration and (<b>b</b>) Nd-dopant concentration and (<b>c</b>) Ce-dopant concentration and lattice strain <span class="html-italic">ε</span>.</p>
Full article ">Figure 4
<p>The Nyquist impedance plots of various compositions of BZCNYYO samples at 500 °C: (<b>a</b>) Ce<span class="html-italic"><sub>x</sub></span> (<span class="html-italic">x</span> = 0, 0.05, 0.10, 0.15 mol) without Nd<sub>y</sub>-dopant (<span class="html-italic">y</span> = 0 mol); (<b>b</b>) Nd<sub>y</sub> (<span class="html-italic">y</span> = 0, 0.05, 0.10, 0.15 mol) without Ce<span class="html-italic"><sub>x</sub></span>-dopant (<span class="html-italic">x</span> = 0 mol); and (<b>c</b>) Ce<span class="html-italic"><sub>x</sub></span> and Nd<sub>y</sub> co-doped samples (<span class="html-italic">x:y</span> = 1 molar ratio).</p>
Full article ">Figure 4 Cont.
<p>The Nyquist impedance plots of various compositions of BZCNYYO samples at 500 °C: (<b>a</b>) Ce<span class="html-italic"><sub>x</sub></span> (<span class="html-italic">x</span> = 0, 0.05, 0.10, 0.15 mol) without Nd<sub>y</sub>-dopant (<span class="html-italic">y</span> = 0 mol); (<b>b</b>) Nd<sub>y</sub> (<span class="html-italic">y</span> = 0, 0.05, 0.10, 0.15 mol) without Ce<span class="html-italic"><sub>x</sub></span>-dopant (<span class="html-italic">x</span> = 0 mol); and (<b>c</b>) Ce<span class="html-italic"><sub>x</sub></span> and Nd<sub>y</sub> co-doped samples (<span class="html-italic">x:y</span> = 1 molar ratio).</p>
Full article ">Figure 5
<p>The Nyquist impedance plots of B1 (<b>a</b>) and B16 (<b>b</b>) samples at 500 °C and their equivalent circuit.</p>
Full article ">Figure 6
<p>The Arrhenius plots of Nd-dopant (Nd<span class="html-italic"><sub>y</sub></span>: 0 mol) (<b>a</b>) and Ce-dopant (Ce<span class="html-italic"><sub>x</sub></span>: 0 mol) (<b>b</b>) BZCNNYO samples.</p>
Full article ">
23 pages, 4518 KiB  
Article
Enhancement of Photocatalytic Activity and Microstructural Growth of Cobalt-Substituted Ba1−xCoxTiO3 {x = 0, …, 1} Heterostructure
by Sana Jebali, Mahdi Meftah, Chadha Mejri, Abdesslem Ben Haj Amara and Walid Oueslati
ChemEngineering 2023, 7(3), 43; https://doi.org/10.3390/chemengineering7030043 - 1 May 2023
Cited by 4 | Viewed by 3116
Abstract
The photocatalytic degradation process and absorption kinetics of the aqueous solution of the Cibacron Brilliant Yellow 3G-P dye (Y) were investigated under UV-Vis light. Pure barium titanate BaTiO3 (BT) and cobalt ion-substituted barium Ba1−xCoxTiO3 (x = 0, [...] Read more.
The photocatalytic degradation process and absorption kinetics of the aqueous solution of the Cibacron Brilliant Yellow 3G-P dye (Y) were investigated under UV-Vis light. Pure barium titanate BaTiO3 (BT) and cobalt ion-substituted barium Ba1−xCoxTiO3 (x = 0, …, 1) nano-compound powders (BCT) were synthesized using the sol–gel method and colloidal solution destabilization, and utilized as photocatalysts. The powder X-ray diffraction (PXRD) crystal structure analysis of the BT nanoparticles (NPs) revealed a prominent reflection corresponding to the perovskite structure. However, impurities and secondary phase distributions were qualitatively identified in the PXRD patterns for x ≥ 0.2 of cobalt substitution rate. Rietveld refinements of the PXRD data showed that the BCT nano-compound series undergoes a transition from perovskite structure to isomorphous ilmenite-type rhombohedral CoTiO3 (CT) ceramic. The nanoparticles produced displayed robust chemical interactions, according to a Fourier transform infrared spectroscopy (FTIR) analysis. The BT and BCT nanoparticles had secondary hexagonal phases that matched the PXRD results and small aggregated, more spherically shaped particles with sizes ranging from 30 to 114 nm, according to transmission electron microscopy (TEM). Following a thorough evaluation of BCT nano-compounds with (x = 0.6), energy-dispersive X-ray (EDX) compositional elemental analysis revealed random distributions of cobalt ions. Through optical analysis of the photoluminescence spectra (PL), the electronic structure, charge carriers, defects, and energy transfer mechanisms of the compounds were examined. Due to the cobalt ions being present in the BT lattice, the UV-visible absorption spectra of BCT showed a little red-shift in the absorption curves when compared to pure BT samples. The electrical and optical characteristics of materials, such as their photon absorption coefficient, can be gathered from their UV-visible spectra. The photocatalytic reaction is brought about by the electron–hole pairs produced by this absorption. The estimated band gap energies of the examined compounds, which are in the range of 3.79 to 2.89 eV, are intriguing and require more investigation into their potential as UV photocatalysts. These nano-ceramics might be able to handle issues with pollution and impurities, such as the breakdown of organic contaminants and the production of hydrogen from water. Full article
Show Figures

Figure 1

Figure 1
<p>Characteristic parts of PXRD patterns of Ba<sub>(1−x)</sub>Co<sub>x</sub>TiO<sub>3</sub> (x = {0, …, 1}) Ba/Co-substituted barium titanate.</p>
Full article ">Figure 2
<p>Results of Rietveld refinement of BaTiO<sub>3</sub> structures. Short vertical bars indicate the positions of diffraction maxima in the major tetragonal phase.</p>
Full article ">Figure 3
<p>FTIR spectra of Ba/Co-substituted barium titanate samples.</p>
Full article ">Figure 4
<p>(<b>a</b>) TEM micrographs and high-resolution images for Ba<sub>1−x</sub>Co<sub>x</sub>TiO<sub>3</sub> sample with (x = 0.6). EDX spectrum of: (<b>b</b>) pure BT and (<b>c</b>) Ba<sub>0.4</sub>Co<sub>0.6</sub>TiO<sub>3</sub> (x = 0.6) nano-heterostructure.</p>
Full article ">Figure 4 Cont.
<p>(<b>a</b>) TEM micrographs and high-resolution images for Ba<sub>1−x</sub>Co<sub>x</sub>TiO<sub>3</sub> sample with (x = 0.6). EDX spectrum of: (<b>b</b>) pure BT and (<b>c</b>) Ba<sub>0.4</sub>Co<sub>0.6</sub>TiO<sub>3</sub> (x = 0.6) nano-heterostructure.</p>
Full article ">Figure 5
<p>(<b>a</b>) UV-Vis absorption spectral of synthesized BCT nanoparticles (<b>b</b>) Photoluminescence spectra of all BCT nanoparticles.</p>
Full article ">Figure 6
<p>(<b>a</b>) UV-Vis DRS spectra of Ba<sub>0.4</sub>Co<sub>0.6</sub>TiO<sub>3</sub> solid material for optical band-gap determination and (<b>b</b>) variation of the gap energy as a function of the cobalt substitution rate.</p>
Full article ">Figure 6 Cont.
<p>(<b>a</b>) UV-Vis DRS spectra of Ba<sub>0.4</sub>Co<sub>0.6</sub>TiO<sub>3</sub> solid material for optical band-gap determination and (<b>b</b>) variation of the gap energy as a function of the cobalt substitution rate.</p>
Full article ">Figure 7
<p>(<b>a</b>) Photocatalytic degradation rate of CBY3G-P dye over Ba<sub>1−x</sub>Co<sub>x</sub>TiO<sub>3</sub> NPs as catalysis under 180 min of irradiation and (<b>b</b>) correlation graph showing the variation of photocatalytic degradation of the CBY3G-P as function of cobalt substitution rate.</p>
Full article ">Figure 8
<p>Schematic ELD of Ba<sub>0.4</sub>Co<sub>0.6</sub>TiO<sub>3</sub> with respect to potential for the generation of (O<sub>2</sub>•−) (E°(OH/H<sub>2</sub>O)) and O<sub>2</sub>•− (EO<sub>2</sub>/O<sub>2</sub>•−) radicals.</p>
Full article ">
12 pages, 6637 KiB  
Article
Phase Instability, Oxygen Desorption and Related Properties in Cu-Based Perovskites Modified by Highly Charged Cations
by Roman A. Shishkin, Alexey Yu. Suntsov and Mikhael O. Kalinkin
Ceramics 2023, 6(2), 968-979; https://doi.org/10.3390/ceramics6020057 - 11 Apr 2023
Viewed by 1676
Abstract
The rock-salt ordered A2CuWO6 (A = Sr, Ba) with I4/m space group and disordered SrCu0.5M0.5O3−δ (M = Ta, Nb) with Pm3m space group perovskites were successfully obtained via a solid-state reaction [...] Read more.
The rock-salt ordered A2CuWO6 (A = Sr, Ba) with I4/m space group and disordered SrCu0.5M0.5O3−δ (M = Ta, Nb) with Pm3m space group perovskites were successfully obtained via a solid-state reaction route. Heat treatment of Ba2CuWO6 over 900 °C in air leads to phase decomposition to the barium tungstate and copper oxide. Thermogravimetric measurements reveal the strong stoichiometric oxygen content and specific oxygen capacity (ΔWo) exceeding 2.5% for Ba2CuWO6. At the same time, oxygen content reveals Cu3+ content in SrCu0.5Ta0.5O3−δ. Under the following reoxidation of Ba2CuWO6, step-like behavior in weight changes was observed, corresponding to possible Cu+ ion formation at 900 °C; in contrast, no similar effect was detected for M5+ cations. The yellow color of Ba2CuWO6 enables to measure the band gap 2.59 eV. SrCu0.5Ta0.5O3−δ due to high oxygen valance concentration has a low thermal conductivity 1.28 W·m−1·K−1 in the temperature range 25–400 °C. Full article
Show Figures

Figure 1

Figure 1
<p>The XRD patterns of (<b>a</b>) samples synthesized at different temperatures; (<b>b</b>) single-phase Ba<sub>2</sub>CuWO<sub>6−δ</sub> obtained after 900 °C; (<b>c</b>) image of the blackened sample after 950 °C.</p>
Full article ">Figure 2
<p>(<b>a</b>) BEC image of cross-section of Ba<sub>2</sub>CuWO<sub>6−δ</sub> synthesized at 950 °C (<b>b</b>) mapping of decomposition border: Ba—green; Cu—red; W—blue.</p>
Full article ">Figure 3
<p>(<b>a</b>) Isothermal dependence of mass change in Ba<sub>2</sub>CuWO<sub>6−δ</sub> at the variable atmosphere; (<b>b</b>) XRD patterns of the sample after TGA.</p>
Full article ">Figure 4
<p>The XRD pattern of (<b>a</b>) SrCu<sub>0.5</sub>Ta<sub>0.5</sub>O<sub>3−δ</sub> with the results of Rietveld refinement; (<b>b</b>) samples synthesized at different temperatures: arrow—SrO·CuO; diamond—Sr<sub>2</sub>Ta<sub>2</sub>O<sub>7</sub>.</p>
Full article ">Figure 5
<p>The SEM images of SrCu<sub>0.5</sub>M<sub>0.5</sub>O<sub>3−δ</sub> for M = (<b>a</b>) Ta, (<b>b</b>) Nb.</p>
Full article ">Figure 6
<p>EDX point analysis of Sr-Cu-Ta-O system calcinated at 1000 °C for 24 h.</p>
Full article ">Figure 7
<p>Isothermal dependence of mass change in SrCu<sub>0.5</sub>Ta<sub>0.5</sub>O<sub>3−δ</sub> at the variable atmosphere.</p>
Full article ">Figure 8
<p>Plot of (<b>a</b>) (R·hυ)2 as a function of photon energy for the estimation of direct energy band gap; (<b>b</b>) thermal expansion of SrCu<sub>0.5</sub>Ta<sub>0.5</sub>O<sub>3−δ</sub>.</p>
Full article ">
Back to TopTop