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

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,997)

Search Parameters:
Keywords = perovskite

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
11 pages, 2717 KiB  
Article
Vapor-Assisted Method to Deposit Compact (CH3NH3)3Bi2I9 Thin Films for Bismuth-Based Planar Perovskite Solar Cells
by Zihao Gao, Xinjie Wang, Zhen Sun, Ping Song, Xiyuan Feng and Zhixin Jin
Micromachines 2025, 16(2), 218; https://doi.org/10.3390/mi16020218 - 14 Feb 2025
Abstract
Bismuth-based perovskite derivatives, (CH3NH3)3Bi2I9 (MBI), are promising non-toxic light-absorbing materials widely used in various photoelectric devices because of their excellent stability. However, MBI-based perovskite solar cells (PSCs) are limited by poor film quality, and [...] Read more.
Bismuth-based perovskite derivatives, (CH3NH3)3Bi2I9 (MBI), are promising non-toxic light-absorbing materials widely used in various photoelectric devices because of their excellent stability. However, MBI-based perovskite solar cells (PSCs) are limited by poor film quality, and the performance of such a device is far behind that of lead-based PSCs. In this work, the crystal structure and morphological properties of MBI films were compared across different preparation methods. The two-step vapor-assisted method can prepare continuous dense MBI films because MBI crystal nucleation is induced by the BiI3 seed layer. The MBI film grown by this method is better for the production of excellent PSCs compared to the film prepared by the solution method. The best photovoltaic device based on the MBI film could obtain a power conversion efficiency of 1.13%. An MBI device is stored in the glove box for 60 days, and the device’s performance is maintained at 99%. These results indicate that the vapor-assisted deposition of MBI films can be an effective method to improve the performance of bismuth-based planar PSCs. Full article
(This article belongs to the Special Issue Energy Conversion Materials/Devices and Their Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Preparation procedure of the MBI thin film by the SC method. (<b>b</b>) TSS method. (<b>c</b>) TSVA method.</p>
Full article ">Figure 2
<p>The surface morphologies of MBI films were prepared by different methods. (<b>a</b>) SC method, (<b>b</b>) TSS, (<b>c</b>) TSVA method.</p>
Full article ">Figure 3
<p>(<b>a</b>) XRD patterns. The “*” symbols represent the signals of the FTO substrates. (<b>b</b>) UV–visible spectra, (<b>c</b>) Tauc plot versus the light energy, (<b>d</b>) photoluminescence spectra of MBI films grown via different methods.</p>
Full article ">Figure 4
<p>(<b>a</b>) Configuration of the planar MBI photovoltaic device; (<b>b</b>) schematic band diagram of MBI film grown via different methods.</p>
Full article ">Figure 5
<p>(<b>a</b>) <span class="html-italic">J</span>-<span class="html-italic">V</span> curves of MBI devices with different growth methods. (<b>b</b>) IPCE spectra and the integrated current for the different devices. (<b>c</b>) Stability of MBI devices with different preparation methods in a glove box. (<b>d</b>) PCE box statistics of 20 photovoltaic devices prepared by different methods.</p>
Full article ">Figure 6
<p>(<b>a</b>) C-V curves of different devices yielded by Mott–Schottky testing; (<b>b</b>) SCLC curves for hole-only devices prepared via different methods.</p>
Full article ">
9 pages, 1685 KiB  
Article
Optimal Methylammounium Chloride Additive for High-Performance Perovskite Solar Cells
by Qinghua Cao, Hui Liu, Jiangping Xing, Bing’e Li, Chuangping Liu, Fobao Xie, Xiaoli Zhang and Weiren Zhao
Nanomaterials 2025, 15(4), 292; https://doi.org/10.3390/nano15040292 - 14 Feb 2025
Abstract
Organic–inorganic lead halide perovskite solar cells (PSCs) have presented promising improvements within recent years due to the superior photophysical properties of perovskites. The efficiency of PSCs is closely related to the quality of the of the perovskite film. Additive engineering is an effective [...] Read more.
Organic–inorganic lead halide perovskite solar cells (PSCs) have presented promising improvements within recent years due to the superior photophysical properties of perovskites. The efficiency of PSCs is closely related to the quality of the of the perovskite film. Additive engineering is an effective strategy to regulate the crystallization of perovskite film. Therefore, in this work, we introduce methylammounium chloride (MACl) into a perovskite precursor as an additive to improve the crystallization of perovskite film and to suppress the formation of defects to achieve high-performance PSCs. By meticulously investigating and studying the influence of different percentages of MACl additives on perovskite film quality, we obtain that the best amount of incorporated MACl is 10%. Thanks the employment of the optimal amount of MACl, the perovskite film shows a significantly improved morphology with larger grains, a smoother surface, and suppressed defects. Finally, the target PSCs with the addition of 10% MACl present the highest PCE of 23.61%, which is much higher than the value (16.72%) of the control device. Full article
(This article belongs to the Special Issue Perovskite Nanomaterials for Photovoltaic and Optoelectronic Devices)
Show Figures

Figure 1

Figure 1
<p>Statistical distribution of (<b>a</b>) PCE, (<b>b</b>) Voc, (<b>c</b>) FF, and (<b>d</b>) Jsc of PSCs with different percentages of MACl additives.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) control perovskite film and (<b>b</b>) target perovskite film; AFM images of (<b>c</b>) control perovskite film and (<b>d</b>) target perovskite film.</p>
Full article ">Figure 3
<p>(<b>a</b>) XRD spectra, (<b>b</b>) absorbance spectra, (<b>c</b>) Tauc plots, and (<b>d</b>) PL spectra of control and target perovskite films.</p>
Full article ">Figure 4
<p>(<b>a</b>) Diagram of structure of PSCs; (<b>b</b>) J-V curves of best-performing target PSCs fabricated with 10% MACl additive and control device; (<b>c</b>) dark current curves of control and target PSCs; (<b>d</b>) light intensity dependence of Voc of control and target PSCs.</p>
Full article ">
8 pages, 1717 KiB  
Article
Analyzing Efficiency of Perovskite Solar Cells Under High Illumination Intensities by SCAPS Device Simulation
by Heng Li, Yongtao Huang, Muyan Zhu, Pingyuan Yan and Chuanxiang Sheng
Nanomaterials 2025, 15(4), 286; https://doi.org/10.3390/nano15040286 - 13 Feb 2025
Abstract
The perovskite solar cell (PSC) is undergoing intense study to meet sustainable energy and environmental demands. However, large-sized solar cells will degrade the power conversion efficiency, thus concentrating light on small-size devices would be a solution. Here, we report the performance of a [...] Read more.
The perovskite solar cell (PSC) is undergoing intense study to meet sustainable energy and environmental demands. However, large-sized solar cells will degrade the power conversion efficiency, thus concentrating light on small-size devices would be a solution. Here, we report the performance of a p–i–n structured device using CH3NH3PbI3 (MAPbI3) as the active layer with an area of 6 mm2. We prove that the power output would be up to 4.2 mW under 10 Suns compared to the 0.9 mW obtained under 1 Sun; however, this results in an actual efficiency drop of the PSC. Further, using a SCAPS device simulation, we found that the intrinsic properties, such as mobility and defect density, of MAPbI3 has no profound influence on the relationship between light intensity and power conversion efficiency (PCE), but the series resistance is the dominant limiting factor on the performance of the PSC under high illumination intensities. Our work suggests the potential of perovskite in concentrating photovoltaics and makes recommendations for future development. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Typical J-V curve of a MAPbI3 perovskite solar cell with the forward and reverse scan direction under a simulated AM 1.5 G illumination. (<b>b</b>) J-V curves of a MAPbI<sub>3</sub> perovskite solar cell excited using a continuous wave laser of 532 nm at various intensities. (<b>c</b>–<b>f</b>): J<sub>sc</sub>, V<sub>oc</sub>, FF number, and PCE of a solar cell as the function of the intensity of the excitation laser.</p>
Full article ">Figure 2
<p>(<b>a</b>) Device structure diagram. (<b>b</b>) Device structure and energy band diagram of each layer. (<b>c</b>) Simulated J-V curves at various defect intensities.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>d</b>): V<sub>oc</sub>, J<sub>sc</sub>, FF number, and PCE as a function of light intensities at various defect densities.</p>
Full article ">Figure 4
<p>(<b>a</b>) Simulated PCE as the function of light intensity at two groups of electron and hole mobility. (<b>b</b>) Simulated PCE as the function of light intensity at two shunt resistances, and the series resistance is fixed at 4.5 Ω·cm<sup>2</sup>. (<b>c</b>,<b>d</b>): Simulated PCE and FF number at three groups of R<sub>s</sub> and R<sub>sh</sub>.</p>
Full article ">
14 pages, 10847 KiB  
Article
Promoting Effect of Copper Doping on LaMO3 (M = Mn, Fe, Co, Ni) Perovskite-Supported Gold Catalysts for Selective Gas-Phase Ethanol Oxidation
by Lijun Yue, Jie Wang and Peng Liu
Catalysts 2025, 15(2), 176; https://doi.org/10.3390/catal15020176 - 13 Feb 2025
Abstract
Developing more effective gold–support synergy is essential for enhancing the catalytic performance of supported gold nanoparticles (AuNPs) in the gas-phase oxidation of ethanol to acetaldehyde (AC) at lower temperatures. This study demonstrates a significantly improved Au–support synergy achieved by copper doping in LaMO [...] Read more.
Developing more effective gold–support synergy is essential for enhancing the catalytic performance of supported gold nanoparticles (AuNPs) in the gas-phase oxidation of ethanol to acetaldehyde (AC) at lower temperatures. This study demonstrates a significantly improved Au–support synergy achieved by copper doping in LaMO3 (M = Mn, Fe, Co, Ni) perovskites. Among the various Au/LaMCuO3 catalysts, Au/LaMnCuO3 exhibited exceptional catalytic activity, achieving an AC yield of up to 91% and the highest space-time yield of 764 gAC gAu−1 h−1 at 225 °C. Notably, this catalyst showed excellent hydrothermal stability, maintaining performance for at least 100 h without significant deactivation when fed with 50% aqueous ethanol. Comprehensive characterization reveals that Cu doping facilitates the formation of surface oxygen vacancies on the Au/LaMCuO3 catalysts and enhances Au–support interactions. The LaMnCuO3 perovskite stabilizes the crucial Cu+ species, resulting in a stable Au-Mn-Cu synergy within the Au/LaMnCuO3 catalyst, which facilitates the activation of O2 and ethanol at lower temperatures. The optimization of the reaction conditions further improves AC productivity. Kinetic studies indicate that the cleavages of both the O-H bond and the α-C-H bond of ethanol are the rate-controlling steps. Full article
(This article belongs to the Special Issue New Insights into Synergistic Dual Catalysis)
Show Figures

Figure 1

Figure 1
<p>XRD spectra of (<b>A</b>) Au/LaMO<sub>3</sub> and (<b>B</b>) Au/LaMCuO<sub>3</sub>.</p>
Full article ">Figure 2
<p>TEM images and Au particle size distributions for (<b>a</b>) Au/LaMnO<sub>3</sub>, (<b>b</b>) Au/LaFeO<sub>3</sub>, (<b>c</b>) Au/LaCoO<sub>3</sub>, (<b>d</b>) Au/LaNiO<sub>3</sub>, (<b>e</b>) Au/LaMnCuO<sub>3</sub>, (<b>f</b>) Au/LaFeCuO<sub>3</sub>, (<b>g</b>) Au/LaCoCuO<sub>3</sub>, and (<b>h</b>) Au/LaNiCuO<sub>3</sub>.</p>
Full article ">Figure 2 Cont.
<p>TEM images and Au particle size distributions for (<b>a</b>) Au/LaMnO<sub>3</sub>, (<b>b</b>) Au/LaFeO<sub>3</sub>, (<b>c</b>) Au/LaCoO<sub>3</sub>, (<b>d</b>) Au/LaNiO<sub>3</sub>, (<b>e</b>) Au/LaMnCuO<sub>3</sub>, (<b>f</b>) Au/LaFeCuO<sub>3</sub>, (<b>g</b>) Au/LaCoCuO<sub>3</sub>, and (<b>h</b>) Au/LaNiCuO<sub>3</sub>.</p>
Full article ">Figure 3
<p>XPS spectra of the Au/LaMCuO<sub>3</sub> catalysts. (<b>A</b>) Au 4f, (<b>B</b>) Cu 2p, and (<b>C</b>) O 1s.</p>
Full article ">Figure 4
<p>H<sub>2</sub>-TPR profiles of various Au/LaMO<sub>3</sub> (<b>A</b>) and Au/LaMCuO<sub>3</sub> (<b>B</b>) catalysts (the dashed line represents the profile of the corresponding support).</p>
Full article ">Figure 5
<p>Catalytic performance of Au/LaMO<sub>3</sub> (<b>A</b>) and Au/LaMCuO<sub>3</sub> (<b>B</b>) catalysts. Arrhenius curve of Au/LaMCuO<sub>3</sub> for ethanol oxidation (<b>C</b>). Catalytic stability of Au/LaMnCuO<sub>3</sub> using anhydrous ethanol as the feed at 225 °C (<b>D</b>) and 50 wt.% aqueous ethanol at 200 °C (<b>E</b>). (Reaction conditions: 0.06 g catalyst; 5 μL min<sup>−1</sup> ethanol; ethanol/O<sub>2</sub>/N<sub>2</sub> = 1/3/36; GHSV = 100,000 mL g<sub>cat</sub><sup>−1</sup> h<sup>−1</sup>.)</p>
Full article ">Figure 6
<p>Effects of oxidant (<b>A</b>,<b>B</b>), concentration of O<sub>2</sub> (<b>C</b>) and ethanol (<b>D</b>), GHSV (<b>E</b>), and temperature (<b>F</b>) on the catalytic performance of ethanol oxidation over Au/LaMnCuO<sub>3</sub> catalysts. (General reaction conditions: 0.06 g catalyst; 0–20 vol% O<sub>2</sub>; 1–10 vol% ethanol; GHSV = 50,000–200,000 h<sup>−1</sup>. (<b>A</b>,<b>B</b>): 2.5 vol% ethanol; ethanol/N<sub>2</sub>O/N<sub>2</sub> = 1/3/36; ethanol/O<sub>2</sub>/H<sub>2</sub>/N<sub>2</sub> = 1/3/1/35; ethanol/O<sub>2</sub>/N<sub>2</sub> = 1/3/36; ethanol/N<sub>2</sub> = 1/39; GHSV = 100,000 h<sup>−1</sup>. (<b>C</b>): 2.5 vol% ethanol; GHSV = 100,000 h<sup>−1</sup>. (<b>D</b>): ethanol/O<sub>2</sub> = 1/3; GHSV = 100,000 h<sup>−1</sup>. (<b>E</b>): 2.5 vol% ethanol; ethanol/O<sub>2</sub> = 1/3. (<b>F</b>): 2.5 vol% ethanol; ethanol/O<sub>2</sub> = 1/3; GHSV = 200,000 h<sup>−1</sup>).</p>
Full article ">
27 pages, 6383 KiB  
Review
A Review of Measurement and Characterization of Film Layers of Perovskite Solar Cells by Spectroscopic Ellipsometry
by Liyuan Ma, Xipeng Xu, Changcai Cui, Tukun Li, Shan Lou, Paul J. Scott, Xiangqian Jiang and Wenhan Zeng
Nanomaterials 2025, 15(4), 282; https://doi.org/10.3390/nano15040282 - 13 Feb 2025
Abstract
This article aims to complete a review of current literature describing the measurement and characterization of photoelectric and geometric properties of perovskite solar cell (PSC) film layer materials using the spectroscopic ellipsometry (SE) measurement technique. Firstly, the influence of film quality on the [...] Read more.
This article aims to complete a review of current literature describing the measurement and characterization of photoelectric and geometric properties of perovskite solar cell (PSC) film layer materials using the spectroscopic ellipsometry (SE) measurement technique. Firstly, the influence of film quality on the performance of PSCs is combed and analyzed. Secondly, SE measurement technology is systematically introduced, including the measurement principle and data analysis. Thirdly, a detailed summary is provided regarding the characterization of the geometric and optoelectronic properties of the substrate, electron transport layer (ETL), perovskite layer, hole transport layer (HTL), and metal electrode layer using SE. The oscillator models commonly used in fitting film layer materials in PSCs are comprehensively summarized. Fourthly, the application of SE combined with various measurement techniques to assess the properties of film layer materials in PSCs is presented. Finally, the noteworthy direction of SE measurement technology in the development of PSCs is discussed. The review serves as a valuable reference for further enhancing the application of SE in PSCs, ultimately contributing to the commercialization of PSCs. Full article
Show Figures

Figure 1

Figure 1
<p>Measurement and characterization of photoelectric and geometric properties for each film layer of a typical SnO<sub>2</sub>-based PSC by SE.</p>
Full article ">Figure 2
<p>Three common structures of PSCs.</p>
Full article ">Figure 3
<p>Influencing factors of power conversion efficiency of PSCs.</p>
Full article ">Figure 4
<p>The basic principle of SE measurement and analysis.</p>
Full article ">Figure 5
<p>Fitting strategies of PSC multilayer films stack structure.</p>
Full article ">Figure 6
<p>Characterization of SnO<sub>2</sub> films by SE. (<b>a</b>) Different substrates [<a href="#B54-nanomaterials-15-00282" class="html-bibr">54</a>]. (<b>b</b>) Different temperature [<a href="#B53-nanomaterials-15-00282" class="html-bibr">53</a>]. (<b>c</b>) Different thickness [<a href="#B51-nanomaterials-15-00282" class="html-bibr">51</a>]. (<b>d</b>) Different technology [<a href="#B52-nanomaterials-15-00282" class="html-bibr">52</a>]. (<b>e</b>) Different doping ratios [<a href="#B55-nanomaterials-15-00282" class="html-bibr">55</a>].</p>
Full article ">Figure 7
<p>Analysis procedure of geometric and photoelectric properties of SnO<sub>2</sub> films by SE.</p>
Full article ">Figure 8
<p>Perovskite materials and devices. (<b>a</b>) Element [<a href="#B16-nanomaterials-15-00282" class="html-bibr">16</a>]. (<b>b</b>) Structure [<a href="#B75-nanomaterials-15-00282" class="html-bibr">75</a>]. (<b>c</b>) Device [<a href="#B76-nanomaterials-15-00282" class="html-bibr">76</a>]. (<b>d</b>) Principle [<a href="#B77-nanomaterials-15-00282" class="html-bibr">77</a>].</p>
Full article ">Figure 9
<p>Preparation process of perovskite film [<a href="#B80-nanomaterials-15-00282" class="html-bibr">80</a>]. (<b>a</b>) One-step coating. (<b>b</b>) Two-step coating.</p>
Full article ">Figure 10
<p>Influence factors of perovskite films. (<b>a</b>) Rough layer [<a href="#B82-nanomaterials-15-00282" class="html-bibr">82</a>,<a href="#B83-nanomaterials-15-00282" class="html-bibr">83</a>]. (<b>b</b>) Ion doping [<a href="#B86-nanomaterials-15-00282" class="html-bibr">86</a>]. (<b>c</b>) Interfacial layer [<a href="#B20-nanomaterials-15-00282" class="html-bibr">20</a>]. (<b>d</b>) Void ratio [<a href="#B61-nanomaterials-15-00282" class="html-bibr">61</a>].</p>
Full article ">Figure 11
<p>Influence of the external environment on perovskite films [<a href="#B87-nanomaterials-15-00282" class="html-bibr">87</a>].</p>
Full article ">Figure 12
<p>Influence of external environment on the properties of the perovskite film characterized by SE. (<b>a</b>) Humidity [<a href="#B89-nanomaterials-15-00282" class="html-bibr">89</a>]. (<b>b</b>) Temperature [<a href="#B90-nanomaterials-15-00282" class="html-bibr">90</a>].</p>
Full article ">Figure 13
<p>(<b>a</b>) Optical constant of the Spiro-OMeTAD film [<a href="#B113-nanomaterials-15-00282" class="html-bibr">113</a>]. (<b>b</b>) Optical constant of PEDOT: PSS, Cu<sub>2</sub>O, and CuI films [<a href="#B114-nanomaterials-15-00282" class="html-bibr">114</a>]. (<b>c</b>) Optical constant of the NiOx film [<a href="#B115-nanomaterials-15-00282" class="html-bibr">115</a>].</p>
Full article ">Figure 14
<p>Assist technique in spectroscopic ellipsometry. (<b>a</b>) Scanning electron microscopy (SEM) [<a href="#B61-nanomaterials-15-00282" class="html-bibr">61</a>]. (<b>b</b>) Photoluminescence (PL) [<a href="#B46-nanomaterials-15-00282" class="html-bibr">46</a>]. (<b>c</b>) Atomic force microscopy (AFM) [<a href="#B61-nanomaterials-15-00282" class="html-bibr">61</a>]. (<b>d</b>) Ultraviolet-visible spectroscopy (UV-Vis) [<a href="#B127-nanomaterials-15-00282" class="html-bibr">127</a>].</p>
Full article ">
16 pages, 6263 KiB  
Article
Stabilizing Perovskite Solar Cells by Methyltriphenylphosphonium Iodide Studied with Maximum Power Point Tracking
by Niklas Manikowsky, Zekarias Teklu Gebremichael, Chikezie Williams Ugokwe, Bashudev Bhandari, Steffi Stumpf, Ulrich S. Schubert and Harald Hoppe
Crystals 2025, 15(2), 176; https://doi.org/10.3390/cryst15020176 - 13 Feb 2025
Abstract
The use of organic halide salts to passivate metal halide perovskite (MHP) surface defects has been studied extensively. Passivating the surface defects of the MHP is of critical importance for realizing highly efficient and stable perovskite solar cells (PSCs). Here, the successful application [...] Read more.
The use of organic halide salts to passivate metal halide perovskite (MHP) surface defects has been studied extensively. Passivating the surface defects of the MHP is of critical importance for realizing highly efficient and stable perovskite solar cells (PSCs). Here, the successful application of a multifunctional organic salt, methyltriphenylphosphonium iodide (MTPPI), used as a passivation additive for grain boundary defects and as a molecular sealing layer in terms of stabilization, has been used to stabilize the mixed cation perovskite RbCsMAFA-PbIBr. To assess the passivating and stabilizing effects of MTPPI on RbCsMAFA-PbIBr PSCs, maximum power point tracking (MPPT) was applied as the most realistic and closest-to-application condition for the ageing test. Here, perovskite solar cells were aged under a light source yielding an excitation intensity corresponding to one sun with maximum power point tracking, which was interrupted periodically by current–voltage sweeps. This allowed for the extraction of all photovoltaic parameters necessary for a proper understanding of the ageing process. The MTPPI additive can donate iodine anions to halide vacancies and compensate a negative surface excess charge with cation interactions. On top of this, the large and bulky methyltriphenylphosphonium (MTPP+) cation may block both the escape of volatile perovskite components and the ingress of oxygen and water vapour. These collective roles of MTPPI have improved both the efficiency and stability of the solar cells compared to the reference without passivation additives. Full article
(This article belongs to the Special Issue Preparation and Characterization of Optoelectronic Functional Films)
Show Figures

Figure 1

Figure 1
<p>SEM images of four different multi-crystalline RbCsMAFA-PbIBr perovskite films (<b>a</b>) containing 0.00 wt.%, (<b>b</b>) containing 0.05 wt.%, (<b>c</b>) containing 0.10 wt.%, and (<b>d</b>) containing 0.20 wt.% of MTPPI additive. The scale bar represents a length of 200 nm. Note: the bright particles in section “(<b>a</b>)” are typical for pristine PbI<sub>2</sub> (compare with [<a href="#B73-crystals-15-00176" class="html-bibr">73</a>,<a href="#B74-crystals-15-00176" class="html-bibr">74</a>]).</p>
Full article ">Figure 2
<p>Transmittance, reflectance (both <b>left</b>), and absorptance (<b>right</b>) of RbCsMAFA-PbIBr perovskite films cast on PEDOT:PSS-coated glass substrates, bearing different amounts of MTPPI processing agent.</p>
Full article ">Figure 3
<p>Photoluminescence (PL) intensity spectra for RbCsMAFA-PbIBr perovskite films showing virtually the same absorptance: with an increasing amount of MTPPI additive, the PL intensity was increasing superlinearly. The recovery of the radiative recombination is a strong indicator for the successful passivation of trap states, commonly located at the grain boundaries.</p>
Full article ">Figure 4
<p>Photovoltaic parameters as a function of the MTPPI mass fraction. (<b>a</b>) Short circuit density, (<b>b</b>) open circuit voltage, (<b>c</b>) fill factor, (<b>d</b>) power conversion efficiency, (<b>e</b>) series resistance, (<b>f</b>) parallel resistance. The full lines indicate the trend for the averaged data—in this case, from 4 to 8 solar cells for each data point—while the closed triangle indicates the maximal value achieved. Clearly, the performance rises in increasing MTPPI mass fractions.</p>
Full article ">Figure 5
<p>External quantum efficiency (EQE) spectra of the solar cells confirm a consistent increase in photocurrent with even small concentrations of MTPPI additive. The calculated total photocurrents for an AM 1.5G spectrum in fact yielded even higher values than the IV-characterization beforehand.</p>
Full article ">Figure 6
<p>Summary of the ageing of unsealed devices under continuous illumination, being tracked at the maximum power point for more than 800 h inside the glovebox. While all devices show considerable degradation, the device containing 0.2 wt.% mass fraction of MTPPI exhibited a considerably improved stability. Shown are the (<b>a</b>) short-circuit photocurrents, (<b>b</b>) open circuit voltage, (<b>c</b>) fill factor, (<b>d</b>) power conversion efficiency, (<b>e</b>) series resistance, and (<b>f</b>) parallel resistance.</p>
Full article ">Figure 7
<p>IV-characteristics of reverse IV-sweep development over time for increasing (from <b>top left</b> to <b>bottom right</b>) concentrations of MTPPI mass fractions (0.00, 0.05, 0.10, and 0.20 wt.%). While in all cases the photocurrent decrease is obvious, the forward currents remained nearly stable for the highest MTPPI concentration.</p>
Full article ">Scheme 1
<p>Layer stack with its materials (<b>left</b>) and the chemical structure of methyltriphenylphosphonium iodide (MTPPI) (<b>right</b>), the passivation additive, are displayed.</p>
Full article ">
24 pages, 8526 KiB  
Review
Research Progress of Halide Perovskite Nanocrystals in Biomedical Applications: A Review
by Guiyun Wang, Yanxia Qi, Zhiyan Zhou, Zhuang Liu and Ruowei Wang
Inorganics 2025, 13(2), 55; https://doi.org/10.3390/inorganics13020055 - 13 Feb 2025
Abstract
Halide perovskite nanocrystals have rapidly emerged as a prominent research topic in materials science over the past decade owing to their exceptional optoelectronic properties and tunability. Their distinctive characteristics, including high light absorption coefficients, high quantum yields, narrow-band emissions, low defect densities, and [...] Read more.
Halide perovskite nanocrystals have rapidly emerged as a prominent research topic in materials science over the past decade owing to their exceptional optoelectronic properties and tunability. Their distinctive characteristics, including high light absorption coefficients, high quantum yields, narrow-band emissions, low defect densities, and adjustable chemical compositions and sizes, position them as highly promising candidates for applications in optoelectronic devices, energy conversion units, and other related systems. However, due to the toxicity and instability of halide perovskite nanocrystals, their widespread application in the biomedical field has been limited in the past. In recent years, numerous innovative coating strategies have been reported to effectively enhance the stability of halide perovskite nanocrystals while confining their toxic metal ions within the coating layers, thereby significantly improving their biocompatibility. This review provides a comprehensive summary of the recent progress of halide perovskite nanocrystals in the field of biomedicine. It covers coating strategies to enhance stability and biocompatibility, as well as the applications of coated halide perovskite nanocrystals in biomedicine, with a particular focus on their unique advantages in bioimaging and chemical sensing. Finally, we address unresolved issues and challenges, such as the metabolic pathways and final products of halide perovskite nanocrystals in vivo. We hope to inspire researchers in the field and provide direction for future studies. Full article
(This article belongs to the Special Issue Advanced Inorganic Semiconductor Materials, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>A schematic illustration depicting the absorption, distribution, and excretion pathways of lead compounds within the human body. For each exposure route, the percentage indicates the proportion of lead absorbed. Lead not absorbed through gastrointestinal ingestion is expelled via feces, while unabsorbed lead from respiratory or dermal exposure is presumed to be exhaled or removed through natural bodily processes, respectively. Once absorbed, lead accumulates in three primary compartments of the body, where it is retained for durations specific to each organ before eventual excretion. Reprinted with permission from ref. [<a href="#B49-inorganics-13-00055" class="html-bibr">49</a>]. Copyright 2016, Springer Nature.</p>
Full article ">Figure 2
<p>The image illustrates mint plants cultivated in (<b>a</b>) uncontaminated control soil and (<b>b</b>) soil contaminated with Pb<sup>2+</sup> from perovskite materials. The lead concentrations detected in the leaves, stems, and roots are displayed adjacent to each corresponding section of the plants. Reprinted with permission from ref. [<a href="#B55-inorganics-13-00055" class="html-bibr">55</a>]. Copyright 2020, Springer Nature.</p>
Full article ">Figure 3
<p>(<b>a</b>) Survival rate of zebrafish, 4 days post-fertilization, plotted against toxicant concentration and pH. (<b>b</b>) Healthy zebrafish (control group). (<b>c</b>) Zebrafish exposed to PbI<sub>2</sub>, displaying brain hemovascular defects, heart edema, and dorsal curvature. (<b>d</b>) Zebrafish embryo exposed to SnI<sub>2</sub>, exhibiting failure to hatch at 96 h post-fertilization. Reprinted with permission from ref. [<a href="#B49-inorganics-13-00055" class="html-bibr">49</a>]. Copyright 2016, Springer Nature.</p>
Full article ">Figure 4
<p>(<b>a</b>) High-resolution transmission electron microscopy (HRTEM) image of CsPbBr<sub>3</sub> nanocrystals. (<b>b</b>) Schematic illustration depicting the fabrication strategies for microscale carriers loaded with CsPbBr<sub>3</sub> nanocrystals. (<b>c</b>) Confocal laser scanning microscopy image of microscale carriers loaded with CsPbBr<sub>3</sub> nanocrystals. (<b>d</b>) Confocal laser scanning microscopy image of CT-26 cells incubated with microscale carriers containing CsPbBr<sub>3</sub> nanocrystals for 24 h. (<b>e</b>) Confocal laser scanning microscopy image of muscle tissue following intramuscular injection of microscale carriers loaded with CsPbBr<sub>3</sub> nanocrystals. Reprinted with permission from ref. [<a href="#B80-inorganics-13-00055" class="html-bibr">80</a>]. Copyright 2022, American Chemical Society.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic diagram illustrating the water-assisted transformation process, which depicts the conversion of CsPbBr<sub>3</sub>/Cs<sub>4</sub>PbBr<sub>6</sub> composite nanocrystals into CsPbBr<sub>3</sub>/CsPb<sub>2</sub>Br<sub>5</sub> composite nanocrystals. (<b>b</b>) Photographs showcasing CsPbBr<sub>3</sub>/CsPb<sub>2</sub>Br<sub>5</sub> composite nanocrystals dispersed in toluene and water taken under UV light (365 nm). (<b>c</b>) Bright-field and fluorescence images of HeLa cells incubated with CsPbBr<sub>3</sub>/CsPb<sub>2</sub>Br<sub>5</sub> nanocrystals for 24 h, with nuclei stained in blue using 4,6-diamidino-2-phenylindole (DAPI) dye. The overlay images reveal the cellular uptake of the nanocrystals. Reprinted with permission from ref. [<a href="#B86-inorganics-13-00055" class="html-bibr">86</a>]. Copyright 2019, American Chemical Society.</p>
Full article ">Figure 6
<p>(<b>a</b>) A schematic illustration outlines the room-temperature fabrication process for polymer capsules containing CsPbBr<sub>3</sub> nanocrystals. This procedure involves encapsulating the nanocrystals within a water-soluble polymer matrix, ensuring both protection and enhanced stability. The photographs in (<b>b</b>,<b>c</b>) showcase aged samples of Cs–oleate-coated and didodecyl dimethylammonium bromide (DDAB)-coated CsPbBr<sub>3</sub> nanocrystals, respectively, encapsulated in polymer capsules and dispersed in deionized water. Images taken under both daylight and UV light demonstrate that both samples retain strong photoluminescence, highlighting their superior emission stability even after prolonged incubation in water. (<b>d</b>) A magnified TEM image provides a detailed view of a single polymer capsule. The embedded sketches illustrate the random distribution of CsPbBr<sub>3</sub> nanocrystals within the capsule matrix, as well as the presence of aggregated structures. These aggregates, identified as nanocrystal clusters, are highlighted with yellow frames, emphasizing their spatial arrangement within the polymer shell. (<b>e</b>) The graph plots the PLQY of polymer-encapsulated CsPbBr<sub>3</sub> nanocrystals dispersed in water over several months. Dotted lines represent the PLQY of un-encapsulated nanocrystals in toluene prior to encapsulation. (<b>f</b>) Confocal fluorescence images show cells incubated for 24 h in a medium containing CsPbBr<sub>3</sub> nanocrystals encapsulated with Cs–oleate coating. The images were captured using an excitation wavelength of 400 nm. Cell nuclei were stained with DAPI dye, which emits a blue fluorescence signal, providing a clear contrast against the bright green fluorescence emitted by the encapsulated nanocrystals. Reprinted with permission from ref. [<a href="#B91-inorganics-13-00055" class="html-bibr">91</a>]. Copyright 2022, American Chemical Society.</p>
Full article ">Figure 7
<p>Schematic representation of (<b>a</b>) the cancer detection process and (<b>b</b>) the structure of CsPbBr<sub>3</sub>–SiO<sub>2</sub>@SiO<sub>2</sub>–Ab nanoparticles. TEM images of CsPbBr<sub>3</sub>–SiO<sub>2</sub>@SiO<sub>2</sub> nanoparticles captured at (<b>c</b>) high and (<b>d</b>) low magnifications. (<b>e</b>) Photographs of CsPbBr<sub>3</sub>–SiO<sub>2</sub>@SiO<sub>2</sub> nanoparticle powder under UV light and ambient daylight. (<b>f</b>) Confocal laser scanning microscopy images of Panc-1 cells treated with CsPbBr<sub>3</sub>–SiO<sub>2</sub>@SiO<sub>2</sub>–Ab nanoparticles. Cell nuclei are stained with blue DAPI, while the nanoparticles are visualized in green. (<b>g</b>) Cellular uptake efficiencies assessed via photoluminescence of Panc-1 cells treated with varying concentrations of nanoparticles. (<b>h</b>) Photoluminescence spectra of nanoparticle solutions and Panc-1 cells treated with the same solution. (<b>i</b>) Cell viability, as determined by the WST-1 assay, following treatment with different concentrations of nanoparticle solutions. (<b>j</b>) In vivo photographic and X-Ray images of pancreatic tumor-bearing mice taken 2 h after intravenous injection of CsPbBr<sub>3</sub>–SiO<sub>2</sub>@SiO<sub>2</sub>–Ab nanoparticles. Reprinted with permission from ref. [<a href="#B100-inorganics-13-00055" class="html-bibr">100</a>]. Copyright 2021, Wiley-VCH GmbH.</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic illustration of CsPbBr<sub>3</sub>@SiO<sub>2</sub> core–shell nanocrystals engineered for two-photon bioimaging applications. (<b>b</b>) TEM images of the as-prepared CsPbBr<sub>3</sub>@SiO<sub>2</sub> nanocrystals. (<b>c</b>) Schematic representation depicting the two-photon absorption and subsequent emission process in CsPbBr<sub>3</sub>@SiO<sub>2</sub> nanocrystals. (<b>d</b>) Fluorescence image of HepG2 cells from the experimental group under 800 nm NIR illumination, demonstrating enhanced luminescent activity. Reprinted with permission from ref. [<a href="#B123-inorganics-13-00055" class="html-bibr">123</a>]. Copyright 2022, Elsevier.</p>
Full article ">Figure 9
<p>Schematic representation of the trinity strategy that empowers CsPbCl<sub>3</sub> nanocrystals to function as hydrophilic and highly efficient fluorescent nanozymes for the development of biomarker reporting platforms. Reprinted with permission from ref. [<a href="#B130-inorganics-13-00055" class="html-bibr">130</a>]. Copyright 2024, American Chemical Society.</p>
Full article ">Figure 10
<p>Schematic representation of SERS-active composites featuring Au–Ag Janus nanoparticles/CsPbBr<sub>3</sub> for immunoassays targeting Staphylococcus aureus enterotoxins. Reprinted with permission from ref. [<a href="#B133-inorganics-13-00055" class="html-bibr">133</a>]. Copyright 2022, American Chemical Society.</p>
Full article ">Figure 11
<p>Liquid-phase dual-functional chiral perovskites have been developed for both hydrogen sulfide (H<sub>2</sub>S) detection and antibacterial applications against <span class="html-italic">Escherichia coli</span>. Reprinted with permission from ref. [<a href="#B134-inorganics-13-00055" class="html-bibr">134</a>]. Copyright 2024, Elsevier.</p>
Full article ">Figure 12
<p>(<b>a</b>) Schematic of the bio-catalytic activity of PM-CsPbX<sub>3</sub> nanocrystals. (<b>b</b>) Photographs of oxidase/PM-CsPbX<sub>3</sub> nanocrystals materials with different substrates under UV light. Reprinted with permission from ref. [<a href="#B136-inorganics-13-00055" class="html-bibr">136</a>]. Copyright 2021, Wiley-VCH GmbH.</p>
Full article ">
18 pages, 1761 KiB  
Article
Oxides for Pt Capture in the Ammonia Oxidation Process—A Screening Study
by Julie Hessevik, Cathinka S. Carlsen, Oskar K. Bestul, David Waller, Helmer Fjellvåg and Anja O. Sjåstad
Reactions 2025, 6(1), 13; https://doi.org/10.3390/reactions6010013 - 11 Feb 2025
Abstract
Metallic Pd/Ni gauzes, located downstream of the Pt/Rh ammonia oxidation catalyst nets in the Ostwald process, is the current technology for capturing volatile gas phase platinum and rhodium species lost from the Pt/Rh combustion catalyst through evaporation. In this screening study, we explore [...] Read more.
Metallic Pd/Ni gauzes, located downstream of the Pt/Rh ammonia oxidation catalyst nets in the Ostwald process, is the current technology for capturing volatile gas phase platinum and rhodium species lost from the Pt/Rh combustion catalyst through evaporation. In this screening study, we explore four oxide families, ABO3 perovskites, (ABO3)n(AO) Ruddlesden–Popper (RP) phases, AO rock salt, and A2O3 sesquioxide type oxides, as alternative materials for platinum capture. It was found that all the tested nickelates, LaNiO3, NdNiO3, La2NiO4, and La4Ni3O10, captured platinum well and formed A2NiPtO6. In contrast, La0.85Sr0.15FeO3, LaFeO3, and LaCoO3 did not capture platinum. CaO, SrO, and Nd2O3 formed low-dimensional platinates such as CaxPt3O4, Sr4PtO6, and a newly discovered neodymium platinate, Nd10.67Pt4O24. Gd2O3 did not capture platinum in bench-scale experiments in dry air, but did, however, seem to capture platinum under pilot plant conditions, likely due to the co-capture of Co lost from the N2O abatement catalyst. The catalytic activity of both oxides and platinum-containing products were studied, toward NOx and N2O decomposition. None of the oxides showed significant activity toward NOx decomposition, and all showed activity toward N2O decomposition, but to different extents. An overall assessment of the screened oxides with respect to potential use in industrial Ostwald conditions is provided. All tested oxides except CaO and SrO withstood industrial conditions. From our assessments, the nickelates and A2O3 (A = Nd, Gd) stand out as superior oxides for platinum capture. Full article
(This article belongs to the Special Issue Feature Papers in Reactions in 2024)
Show Figures

Figure 1

Figure 1
<p>NO<sub>x</sub> decomposition (%) versus temperature over (<b>a</b>) LaNiO<sub>3</sub>, La<sub>2</sub>NiO<sub>4</sub>, La<sub>4</sub>Ni<sub>3</sub>O<sub>10</sub>, and La<sub>2</sub>NiPtO<sub>6</sub>; (<b>b</b>) NdNiO<sub>3</sub> and Nd<sub>2</sub>NiPtO<sub>6</sub>; (<b>c</b>) Nd<sub>2</sub>O<sub>3</sub> and Nd<sub>10.67</sub>Pt<sub>4</sub>O<sub>24</sub>; (<b>d</b>) CaO and Ca<sub>x</sub>Pt<sub>3</sub>O<sub>4</sub>/CaPt<sub>2</sub>O<sub>4</sub>/Pt; (<b>e</b>) background measurement of empty reactor. The gas feed composition was 1000 ppm NO<sub>x</sub> diluted in 5% O<sub>2</sub> in N<sub>2</sub>. The results are not corrected for the contribution of the empty reactor.</p>
Full article ">Figure 2
<p>N<sub>2</sub>O decomposition activity [10<sup>−5</sup>mol/m<sup>2</sup>min] of platinum capture oxides and platinum-containing counterparts at 800 °C. Note, the activity of CaO, Ca<sub>x</sub>Pt<sub>3</sub>O<sub>4</sub>/CaPt<sub>2</sub>O<sub>4</sub>/Pt, and LaNiO<sub>3</sub> are likely underestimated due to nearly 100% decomposition of N<sub>2</sub>O (see SI).</p>
Full article ">Figure 3
<p>Schematic illustration of Nd<sub>2</sub>O<sub>3</sub>-capturing gaseous PtO<sub>2</sub> and the subsequent transformation into Nd<sub>10.67</sub>Pt<sub>4</sub>O<sub>24</sub>, together with some identified evaluation steps needed for translating from lab-scale to full-scale platinum capture technology.</p>
Full article ">Scheme 1
<p>Schematic drawing of the bench-scale setup for platinum capture in dry air (444 mL/min) in a 6-zone furnace with a tube-in-a-tube concept held together with Swagelok (Solon, OH, USA) couplings and Teflon ferules. Three parallel experiments could be run at the same time. Long outer quartz tubes ran through the furnace from gas inlet to gas outlet. Pt coils were placed upstream at 1000 °C in zone 2 by inserting inner quartz tube on the inlet side, while oxide pellets were placed downstream in zones 4, 5, and 6 at 900, 800, and 700 °C, respectively, by inserting inner quartz tubes on the outlet side.</p>
Full article ">Scheme 2
<p>Schematic drawing of the pilot plant setup for platinum capture at ammonia oxidation process conditions. Two rounds of different durations (21 and 12 days, respectively) and different configurations were carried out. The oxide samples were placed downstream of the Pt-Rh ammonia oxidation catalyst, the Pd-Ni catchment gauzes, and the laughing gas abatement catalyst in round 1, but placed just downstream of the Pt-Rh catalyst in round 2.</p>
Full article ">Scheme 3
<p>Schematic drawing of the setup for bench-scale catalytic performance testing. Inlet gas contained 5 vol.% O<sub>2</sub> in N<sub>2</sub> and 1000 ppm NO or N<sub>2</sub>O. The sample was placed as a bed (approximately 8 mm in height) of small granules (d = 1–2 mm) on top of a glass sinter in the inner reactor. The temperature was measured by K-type thermocouples placed 1 cm above and below the sample bed. Measurement of NO<sub>x</sub>/N<sub>2</sub>O of the inlet gas and outlet gas was alternated, and the pressure drop was measured between the inlet and outlet pressure.</p>
Full article ">
12 pages, 1742 KiB  
Article
Simulation of Lead-Free Perovskite Solar Cells with Improved Performance
by Saood Ali, Praveen Kumar, Khursheed Ahmad and Rais Ahmad Khan
Crystals 2025, 15(2), 171; https://doi.org/10.3390/cryst15020171 - 10 Feb 2025
Abstract
At present, lead halide PVSKSCs are promising photovoltaic cells but have some limitations, including their low stability in ambient conditions and the toxicity of lead. Thus, it will be of great significance to explore lead-free perovskite materials as an alternative absorber layer. In [...] Read more.
At present, lead halide PVSKSCs are promising photovoltaic cells but have some limitations, including their low stability in ambient conditions and the toxicity of lead. Thus, it will be of great significance to explore lead-free perovskite materials as an alternative absorber layer. In recent years, the numerical simulation of perovskite solar cells (PVSKSCs) via the solar cell capacitance simulation (SCAPS) method has attracted the attention of the scientific community. In this work, we adopted SCAPS for the theoretical study of lead (Pb)-free PVSKSCs. A cesium bismuth iodide (CsBi3I10; CBI) perovskite-like material was used as an absorber layer. The thickness of the CBI layer was optimized. In addition, different electron transport layers (ETLs), such as titanium dioxide (TiO2), tin oxide (SnO2), zinc oxide (ZnO), and zinc selenide (ZnSe), and different hole transport layers, such as spiro-OMeTAD (2,2,7,7-tetrakis(N,N-di(4-methoxyphenylamine)-9,9′-spirobifluorene), poly(3-hexylthiophene-2,5-diyl) (P3HT), poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine (PTAA), and copper oxide (Cu2O), were explored for the simulation of CBI-based PVSKSCs. A device structure of FTO/ETL/CBI/HTL/Au was adopted for simulation studies. The simulation studies showed the improved photovoltaic performance of CBI-based PVSKSCs using spiro-OMeTAD and TiO2 as the HTL and ETL, respectively. An acceptable PCE of 11.98% with a photocurrent density (Jsc) of 17.360258 mA/cm2, a fill factor (FF) of 67.10%, and an open-circuit voltage (Voc) of 1.0282 V were achieved under the optimized conditions. It is expected that the present study will be beneficial for researchers working towards the development of CBI-based PVSKSCs. Full article
(This article belongs to the Section Materials for Energy Applications)
Show Figures

Figure 1

Figure 1
<p>J-V results for FTO (500 nm)/TiO<sub>2</sub> (100 nm)/CBI (150 nm)/spiro-MeOTAD (150 nm)/Au. Inset shows device configuration.</p>
Full article ">Figure 2
<p>(<b>a</b>) J-V curves and (<b>b</b>–<b>e</b>) photovoltaic parameters of the simulated device using different thicknesses of CBI.</p>
Full article ">Figure 3
<p>(<b>a</b>) J-V curves and (<b>b</b>–<b>e</b>) photovoltaic parameters of the simulated device using different HTL layers.</p>
Full article ">Figure 4
<p>(<b>a</b>) J-V curves and (<b>b</b>–<b>e</b>) photovoltaic parameters of the simulated device using different ETL layers.</p>
Full article ">Figure 5
<p>(<b>a</b>) J-V curves and (<b>b</b>–<b>e</b>) photovoltaic parameters of the simulated device using different thicknesses of the TiO<sub>2</sub> layer.</p>
Full article ">Figure 6
<p>Schematic representation of the band alignments of the absorber layer, different ETLs, and HTLs.</p>
Full article ">
28 pages, 6510 KiB  
Review
[MxLy]n[MwXz]m Non-Perovskite Hybrid Halides of Coinage Metals Templated by Metal–Organic Cations: Structures and Photocatalytic Properties
by Piotr W. Zabierowski
Solids 2025, 6(1), 6; https://doi.org/10.3390/solids6010006 - 8 Feb 2025
Abstract
This review provides an analysis of non-perovskite hybrid halides of coinage metals templated by metal–organic cations (CCDC November 2023). These materials display remarkable structural diversity, from zero-dimensional molecular complexes to intricate three-dimensional frameworks, allowing fine-tuning of their properties. A total of 208 crystal [...] Read more.
This review provides an analysis of non-perovskite hybrid halides of coinage metals templated by metal–organic cations (CCDC November 2023). These materials display remarkable structural diversity, from zero-dimensional molecular complexes to intricate three-dimensional frameworks, allowing fine-tuning of their properties. A total of 208 crystal structures, comprising haloargentates, mixed-metal haloargentates, and halocuprates, are categorized and examined. Their potential in photocatalysis is discussed. Special attention is given to the structural adaptability of these materials for the generation of functional interfaces. This review highlights key compounds and aims to inspire further research into optimizing hybrid halides for advanced technological applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Some of the characteristic structures of hybrid haloargentates from the CSD. The lines correspond to the unit cell dimensions and the letter codes to CSD codes. The colors of the atoms in a ball and stick representation: carbon—dark gray, hydrogen—light gray, chlorine—bright green, bromine—dark orange, iodine—purple, nitrogen—blue, oxygen—red, nickel—green, copper—dark orange, silver—gray, zinc—dark blue.</p>
Full article ">Figure 2
<p>Some of the characteristic structures of hybrid haloargentates from the CSD. The lines correspond to the unit cell dimensions and the letter codes to CSD codes. The colors of the atoms in a ball and stick representation: carbon—dark gray, hydrogen—light gray, chlorine—bright green, bromine—dark orange, iodine—purple, nitrogen—blue, oxygen—red, sulfur—yellow, silver—gray, zinc—dark blue, iron—light red, manganese—light purple, ruthenium—turquoise.</p>
Full article ">Figure 3
<p>Some of the characteristic structures of hybrid haloargentates from the CSD. The lines correspond to the unit cell dimensions and the letter codes to CSD codes. The colors of the atoms in a ball and stick representation: carbon—dark gray, hydrogen—light gray, bromine—dark orange, iodine—purple, nitrogen—blue, oxygen—red, sulfur—yellow, copper—dark orange, silver—gray, zinc—dark blue, vanadium—gray, iron—light red, neodymium—light green.</p>
Full article ">Figure 4
<p>Some of the characteristic structures of hybrid haloargentates from the CSD. The lines correspond to the unit cell dimensions and the letter codes to CSD codes. The colors of the atoms in a ball and stick representation: carbon—dark gray, hydrogen—light gray, chlorine—bright green, bromine—dark orange, iodine—purple, nitrogen—blue, oxygen—red, sulfur—yellow, aluminum—pink, copper—dark orange, silver—gray, zinc—dark blue, lead—dark gray, potassium—dark violet, barium—green, lanthanum—light blue, dysprosium—light green.</p>
Full article ">Figure 5
<p>Some of the characteristic structures of hybrid haloargentates from the CSD. The lines correspond to the unit cell dimensions and the letter codes to CSD codes. The colors of the atoms in a ball and stick representation: carbon—dark gray, hydrogen—light gray, chlorine—bright green, bromine—dark orange, iodine—purple, nitrogen—blue, oxygen—red, sulfur—yellow, nickel—green, silver—gray (polyhedron), cobalt—violet, iron—light red, manganese—light purple, barium—green (polyhedron), lead—dark gray, praseodymium—light green (polyhedron), dysprosium—light green.</p>
Full article ">Figure 6
<p>Some of the characteristic structures of mixed-metal haloargentate structures. The lines correspond to the unit cell dimensions and the letter codes to CSD codes. The colors of the atoms in a ball and stick representation: carbon—dark gray, hydrogen—light gray, bromine—dark orange, iodine—purple, nitrogen—blue, oxygen—red, sulfur—yellow, nickel—green, iron—dark orange, silver—gray, zinc—dark blue, lead—dark gray, barium—green, potassium—dark violet, neodymium—light green.</p>
Full article ">Figure 7
<p>Some of the characteristic structures of halocuprate structures. The lines correspond to the unit cell dimensions and the letter codes to CSD codes. For the structure of VAHWAD the hydrogen atoms were omitted. The colors of the atoms in a ball and stick representation: carbon—dark gray, hydrogen—light gray, chlorine—bright green, bromine—dark orange, iodine—purple, nitrogen—blue, oxygen—red, sulfur—yellow, copper—dark orange, barium—green, lithium—pink.</p>
Full article ">Figure 8
<p>Some of the characteristic structures of halocuprate compounds. The lines correspond to the unit cell dimensions and the letter codes to CSD codes. The colors of the atoms in a ball and stick representation: carbon—dark gray, hydrogen—light gray, chlorine—bright green, bromine—dark orange, iodine—purple nitrogen—blue, oxygen—red, copper—dark orange, cobalt—violet, manganese—light purple.</p>
Full article ">
14 pages, 7012 KiB  
Article
The Effect of A-Cation and X-Anion Substitutions on the Electronic and Structural Properties of A2ZrX6 ‘Defect’ Perovskite Materials: A Theoretical Density Functional Theory Study
by Christina Kolokytha, Nektarios N. Lathiotakis, Andreas Kaltzoglou, Ioannis D. Petsalakis and Demeter Tzeli
Materials 2025, 18(3), 726; https://doi.org/10.3390/ma18030726 - 6 Feb 2025
Abstract
In the present work, nine ‘defect’ perovskites with the chemical formula A2ZrX6 have been studied, where the A-site cations are a methylammonium cation, formamidinium cation, and trimethyl-sulfonium cation and the X-site anions are halogen, X = Cl, Br, and I. [...] Read more.
In the present work, nine ‘defect’ perovskites with the chemical formula A2ZrX6 have been studied, where the A-site cations are a methylammonium cation, formamidinium cation, and trimethyl-sulfonium cation and the X-site anions are halogen, X = Cl, Br, and I. We employ periodic DFT calculations using GGA-PBE, MBJ, HSEsol, and HSE06 functionals. All studied compounds exhibit a wide-bandgap energy that ranges from 5.22 eV to 2.11 eV, while for some cases, geometry optimization led to significant structural modification. It was found that the increase in the halogen size resulted in a decrease in the bandgap energy. The choice of the organic A-site cation affects the bandgap as well, which is minimal for the methylammonium cation. Such semiconductors with organic cations may be utilized in optoelectronic devices, given the substantial benefit of solution processability and thin film formation compared to purely inorganic analogs, such as Cs2ZrX6. Full article
(This article belongs to the Section Advanced Nanomaterials and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>Calculated band structure and DOS of the (<b>a</b>) experimental crystal structure TMS<sub>2</sub>ZrCl<sub>6</sub> and (<b>b</b>) energetically optimized crystal structure TMS<sub>2</sub>ZrCl<sub>6</sub>.</p>
Full article ">Figure 2
<p>Calculated DOS of TMS<sub>2</sub>ZrCl<sub>6</sub> using three different functionals: (<b>a</b>) Meta-GGA functional mBJ, (<b>b</b>) screened hybrid functional HSEsol, and (<b>c</b>) screened hybrid functional HSE06.</p>
Full article ">Figure 3
<p>The optimized cubic crystal structures are shown for (<b>a</b>) TMS<sub>2</sub>ZrCl<sub>6</sub>, (<b>b</b>) TMS<sub>2</sub>ZrBr<sub>6</sub>, and (<b>c</b>) TMS<sub>2</sub>ZrI<sub>6</sub>. The images were created using VESTA software, version 3.90.1a [<a href="#B35-materials-18-00726" class="html-bibr">35</a>]. Color assignment: yellow for S, black for carbon, white for hydrogen, grey for Zr, green for Cl, brown for Br, and violet for I.</p>
Full article ">Figure 4
<p>The calculated GGA-PBE band structure and DOS of (<b>a</b>) TMS<sub>2</sub>ZrBr<sub>6</sub> and (<b>b</b>) TMS<sub>2</sub>ZrI<sub>6</sub>.</p>
Full article ">Figure 5
<p>The calculated HSE06 DOS of (<b>a</b>) TMS<sub>2</sub>ZrBr<sub>6</sub> and (<b>b</b>) TMS<sub>2</sub>ZrI<sub>6</sub>.</p>
Full article ">Figure 6
<p>The optimized crystal structures of (<b>a</b>) MA<sub>2</sub>ZrCl<sub>6</sub>, (<b>b</b>) MA<sub>2</sub>ZrBr<sub>6</sub>, and (<b>c</b>) MA<sub>2</sub>ZrI<sub>6</sub>. The images were created using VESTA software, version 3.90.1a. Color assignment: blue for N, black for carbon, white for hydrogen, grey for Zr, green for Cl, brown for Br, and violet for I.</p>
Full article ">Figure 7
<p>The calculated DOS of (<b>a</b>) MA<sub>2</sub>ZrCl<sub>6</sub>, (<b>b</b>) MA<sub>2</sub>ZrBr<sub>6</sub>, and (<b>c</b>) MA<sub>2</sub>ZrI<sub>6</sub> using the GGA-PBE functional.</p>
Full article ">Figure 8
<p>The calculated DOS of (<b>a</b>) MA<sub>2</sub>ZrCl<sub>6</sub>, (<b>b</b>) MA<sub>2</sub>ZrBr<sub>6</sub>, and (<b>c</b>) MA<sub>2</sub>ZrI<sub>6</sub> using the HSE06 functional.</p>
Full article ">Figure 9
<p>The optimized crystal structures are shown for (<b>a</b>) FA<sub>2</sub>ZrCl<sub>6</sub>, (<b>b</b>) FA<sub>2</sub>ZrBr<sub>6</sub>, and (<b>c</b>) FA<sub>2</sub>ZrI<sub>6</sub>. The images were created using VESTA software, version 3.90.1a. Color assignment: blue for N, black for carbon, white for hydrogen, grey for Zr, green for Cl, brown for Br and violet for I.</p>
Full article ">Figure 10
<p>The calculated DOS of (<b>a</b>) FA<sub>2</sub>ZrCl<sub>6</sub>, (<b>b</b>) FA<sub>2</sub>ZrBr<sub>6</sub>, and (<b>c</b>) FA<sub>2</sub>ZrI<sub>6</sub> using the GA-PBE functional.</p>
Full article ">Figure 11
<p>The calculated DOS of (<b>a</b>) FA<sub>2</sub>ZrCl<sub>6</sub>, (<b>b</b>) FA<sub>2</sub>ZrBr<sub>6</sub>, and (<b>c</b>) FA<sub>2</sub>ZrI<sub>6</sub> using the HSE06 functional.</p>
Full article ">Figure 12
<p>Bandgap energy as a function of the molecular weight of the A-site cation (g/mol) for the A<sub>2</sub>ZrX<sub>6</sub> crystal, calculated using the HSE06 (solid lines) and GGA-PBE (dash lines) functionals. Available experimental value is also included [<a href="#B20-materials-18-00726" class="html-bibr">20</a>]. Color assignment: blue for N, grey for carbon, white for hydrogen, yellow for S.</p>
Full article ">Figure 13
<p>Bandgap energy as a function of halogen electronegativity of the A<sub>2</sub>ZrX<sub>6</sub>, calculated using the HSE06 (solid lines) and GGA-PBE (dash lines) functionals. Available experimental value is also included [<a href="#B20-materials-18-00726" class="html-bibr">20</a>].</p>
Full article ">
16 pages, 5819 KiB  
Article
Synthesis, Characterization, and Electrocatalytic Properties of PrMn0.5M0.5O3 (M = Cr, Fe, Co, Ni) Perovskites
by Besarta Cheliku Ramadani, Jeta Sela, Leon Stojanov, Sofija Popovska, Valentin Mirčeski, Miha Bukleski, Sandra Dimitrovska-Lazova, Arianit A. Reka and Slobotka Aleksovska
Materials 2025, 18(3), 717; https://doi.org/10.3390/ma18030717 - 6 Feb 2025
Abstract
In this paper, the synthesis, characterization, and investigation of electrocatalytic properties of perovskites of general formula PrMn0.5M0.5O3 (M = Cr, Fe, Co, Ni) are presented. The synthesis was conducted by the solution combustion method using glycine as a [...] Read more.
In this paper, the synthesis, characterization, and investigation of electrocatalytic properties of perovskites of general formula PrMn0.5M0.5O3 (M = Cr, Fe, Co, Ni) are presented. The synthesis was conducted by the solution combustion method using glycine as a fuel. The perovskite with the formula PrMn0.5Fe0.5O3 was also synthesized by the sol–gel combustion method with citric acid as fuel. The obtained perovskites were investigated by X-ray powder diffraction (XRPD), scanning electron microscopy (SEM) with energy dispersive X-ray spectroscopy (EDX), infrared spectroscopy, and cyclic voltammetry. The XRPD patterns showed that the compounds are pure and isostructural within the series. The unit cell parameters of the compounds were determined within the Pnma space group, and several crystallochemical parameters were calculated and discussed. The recorded SEM images of the perovskites revealed a porous morphology, while the EDX analysis confirmed the 2:1:1 atomic percentage ratio of Pr:Mn:M. Within this investigation, the electrocatalytic properties of the obtained perovskites towards oxidation of OH ions and H2O2 oxidation in phosphate buffer were studied by cyclic voltammetry, using a paraffin-impregnated graphite electrode (PIGE) modified with microcrystals of the investigated perovskites. PrMn0.5Fe0.5O3 showed high electrocatalytic activity for OH oxidation, while both PrMn0.5Fe0.5O3 and PrMn0.5Co0.5O3 exhibited significant efficiency for H2O2 oxidation, with a distinct oxidation peak with a peak potential of 0.6 V. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray powder diffraction patterns of PrMn<sub>0.5</sub>M<sub>0.5</sub>O<sub>3</sub> (M = Cr, Fe, Co, Ni) perovskites.</p>
Full article ">Figure 2
<p>The XRPD patterns of the PrMn<sub>0.5</sub>Fe<sub>0.5</sub>O<sub>3</sub> perovskite, obtained by two different synthetic procedures.</p>
Full article ">Figure 3
<p>SEM images of PrMn<sub>0.5</sub>M<sub>0.5</sub>O<sub>3</sub> (M = Ni, Co, Cr, Fe).</p>
Full article ">Figure 4
<p>Infrared spectra of the PrMn<sub>0.5</sub>M<sub>0.5</sub>O<sub>3</sub> (M = Ni, Co, Cr, Fe) series.</p>
Full article ">Figure 5
<p>Cyclic voltammograms of a bare PIGE electrode (black dotted line) and PIGE modified with particles of PrMn<sub>0.5</sub>Ni<sub>0.5</sub>O<sub>3</sub> (black), PrMn<sub>0.5</sub>Fe<sub>0.5</sub>O<sub>3</sub> (green), PrMn<sub>0.5</sub>Cr<sub>0.5</sub>O<sub>3</sub> (blue), and PrMn<sub>0.5</sub>Co<sub>0.5</sub>O<sub>3</sub> (red) recorded in contact with a 0.5 mol/dm<sup>3</sup> KNO<sub>3</sub> used as a supporting electrolyte. These measurements correspond to a third consecutive scan. The potential was initially scanned from OCP in the positive direction. Additional conditions are specified in the Experimental Section.</p>
Full article ">Figure 6
<p>Cyclic voltammograms of a bare PIGE electrode (black dotted line) and PIGE modified with particles of PrMn<sub>0.5</sub>Ni<sub>0.5</sub>O<sub>3</sub> (black), PrMn<sub>0.5</sub>Fe<sub>0.5</sub>O<sub>3</sub> (green), PrMn<sub>0.5</sub>Cr<sub>0.5</sub>O<sub>3</sub> (blue), and PrMn<sub>0.5</sub>Co<sub>0.5</sub>O<sub>3</sub> (red) recorded in contact with a 0.5 mol/dm<sup>3</sup> KNO<sub>3</sub> and 0.1 mol/dm<sup>3</sup> KOH. These measurements correspond to the first (<b>A</b>) and third (<b>B</b>) consecutive cycles. The potential was initially scanned from OCP in the positive direction. Additional conditions are specified in the Experimental Section.</p>
Full article ">Figure 7
<p>Cyclic voltammograms of a bare PIGE electrode (black dotted line) and of PIGE modified with particles of PrMn<sub>0.5</sub>Ni<sub>0.5</sub>O<sub>3</sub> (black), PrMn<sub>0.5</sub>Fe<sub>0.5</sub>O<sub>3</sub> (green), PrMn<sub>0.5</sub>Cr<sub>0.5</sub>O<sub>3</sub> (blue), and PrMn<sub>0.5</sub>Co<sub>0.5</sub>O<sub>3</sub> (red) recorded in contact with a 0.1 mol/dm<sup>3</sup> phosphate buffer. These measurements correspond to the first (<b>A</b>) and third (<b>B</b>) consecutive cycles. The potential was initially scanned from OCP in the positive direction. Additional conditions are specified in the Experimental Section.</p>
Full article ">Figure 8
<p>Cyclic voltammograms of a bare PIGE electrode (black dotted line) and of PIGE modified with particles of PrMn<sub>0.5</sub>Ni<sub>0.5</sub>O<sub>3</sub> (black), PrMn<sub>0.5</sub>Fe<sub>0.5</sub>O<sub>3</sub> (green), PrMn<sub>0.5</sub>Cr<sub>0.5</sub>O<sub>3</sub> (blue), and PrMn<sub>0.5</sub>Co<sub>0.5</sub>O<sub>3</sub> (red) recorded in contact with a 0.1 mol/dm<sup>3</sup> phosphate buffer and 10<sup>−2</sup> mol/dm<sup>3</sup> H<sub>2</sub>O<sub>2</sub>. These measurements correspond to the first (<b>A</b>) and third (<b>B</b>) consecutive cycles. The potential was initially scanned from OCP in the positive direction. Additional conditions are specified in the Experimental Section.</p>
Full article ">
12 pages, 3506 KiB  
Article
Photoluminescence and Stability of Dion–Jacobson Tin-Based Halide Perovskites with Different Spacer Cation Chain Length
by Muhammad Umair Ali, Wen Ting Sun, Aleksandr A. Sergeev, Atta Ur Rehman, Kam Sing Wong, Aleksandra B. Djurišić and Jasminka Popović
Molecules 2025, 30(3), 703; https://doi.org/10.3390/molecules30030703 - 5 Feb 2025
Abstract
Two-dimensional tin halide perovskites are of significant interest for light emitting applications. Here, we investigate the effect of organic cation A on the stability of different Dion–Jacobson tin-based halide perovskites. The ASnBr4 materials using diammonium cation A with shorter alkyl chains are [...] Read more.
Two-dimensional tin halide perovskites are of significant interest for light emitting applications. Here, we investigate the effect of organic cation A on the stability of different Dion–Jacobson tin-based halide perovskites. The ASnBr4 materials using diammonium cation A with shorter alkyl chains are found to exhibit improved stability, exhibiting dramatic stability difference between the most stable HDASnBr4, where HDA denotes 1,6-hexanediammonium, and two materials with 8- and 10-carbon alkyl chain ammonium cations. The HDASnBr4 powders were thermally stable at 100 °C in an argon environment but exhibited decreasing photoluminescence with time in ambient air at 100 °C. The sample degradation at 100 °C is accelerated compared to room temperature, but it proceeds along similar pathways, namely phase transformation followed by perovskite decomposition. Light emission from HDASnBr4 thin films could be further enhanced by methanol vapor treatment, and warm white emission with Commission Internationale de l’Eclairage (CIE) coordinates (0.37, 0.34) could be obtained by combining HDASnBr4 with a blue-emitting polymer film, while direct mixing of blue phosphor and HDASnBr4 powder yields white emission with CIE coordinates of (0.34, 0.32). Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Photoluminescence of ASnBr<sub>4</sub> as a function of time in ambient air at 100 °C for (<b>b</b>) A = HDA, (<b>c</b>) A = ODA, and (<b>d</b>) A = DDA.</p>
Full article ">Figure 2
<p>Le Bail refinement of (<b>a</b>) HDA-based perovskite, (<b>b</b>) ODA-based perovskite, (<b>c</b>) DDA-based perovskite. Experimental diffraction patterns are shown as black lines while the calculated patterns for are represented by blue lines. Diffraction lines corresponding to the 2D phase are marked with green vertical lines while the reflections of the 1D phase are marked with dark red vertical bars. (<b>d</b>) Structures of the 2D HDASnBr<sub>4</sub> phase and 1D HDA<sub>3</sub>SnBr<sub>8</sub>·(H<sub>2</sub>O) phase, Sn-octahedra are given in teal, bromides are shown as orange balls while carbons, nitrogens, oxygens, hydrogens are gray, light blue, red, and white balls, respectively; (<b>e</b>) enlarged part of the XRD patterns showing the most prominent diffraction lines for as-prepared HDA, ODA, and DDA samples. Diffraction lines corresponding to the 2D phase are colored in green while the reflections of the 1D phase are colored in dark red.</p>
Full article ">Figure 3
<p>XRD patterns of thermally treated samples with different spacers: (<b>a</b>) ODA, (<b>b</b>) DDA, (<b>c</b>) HDA; unidentified phase is denoted by *, (<b>d</b>) enlarged part of XRD pattern for HDA sample, (<b>e</b>) Le Bail refinement of diffraction lines at ~8.9°2θ. The diffraction line corresponding to the 2D phase is colored in green while the reflection of the 1D phase is colored in dark red.</p>
Full article ">Figure 4
<p>Thermogravimetric analysis plot of ASnBr<sub>4</sub> perovskites for different spacers A (HDA, ODA, and DDA).</p>
Full article ">Figure 5
<p>(<b>a</b>) PL spectra of HDASnBr<sub>4</sub> film with different blue-emitting polymer formulations. TFB10 and TFB20 denote 10 and 20 mg of TFB dissolved in 1 mL chlorobenzene, and BP@TFB indicates 10 mg of BP dissolved in TFB solution. (<b>b</b>) CIE coordinates chart for HDASnBr<sub>4</sub>+ BP@TFB corresponding to (0.37, 0.34). (<b>c</b>) PL spectrum of HDASnBr<sub>4</sub> BP mixture with a ratio 1:0.50 (inset shows the corresponding photo under UV illumination) and (<b>d</b>) corresponding CIE coordinates (0.34, 0.32).</p>
Full article ">
10 pages, 2740 KiB  
Communication
Yttrium Doping of Perovskite Oxide La2Ti2O7 Nanosheets for Enhanced Proton Conduction and Gas Sensing Under HighHumidity Levels
by Jian Wang, Caicai Sun, Jusheng Bao, Zhiwei Yang, Jian Zhang and Xiao Huang
Sensors 2025, 25(3), 901; https://doi.org/10.3390/s25030901 - 2 Feb 2025
Abstract
Water molecules from the environment or human breath are one of the main factors affecting the accuracy, efficiency, and long-term stability of electronic gas sensors. In this contribution, yttrium (Y)-doped La2Ti2O7 (LTO) nanosheets were synthesized by a hydrothermal [...] Read more.
Water molecules from the environment or human breath are one of the main factors affecting the accuracy, efficiency, and long-term stability of electronic gas sensors. In this contribution, yttrium (Y)-doped La2Ti2O7 (LTO) nanosheets were synthesized by a hydrothermal reaction, demonstrating improved proton conductivity compared to their non-doped counterparts. The response of Y-doped LTO with the optimal doping concentration to 100 ppm NO2 at 43% relative humidity (RH) was −21%, which is four times higher than that of bare La2Ti2O7. As the humidity level increased to 75%, the response of Y-doped LTO further increased to −64%. Unlike the gas doping effect observed in previous studies of semiconducting metal oxides, the sensing mechanism of Y-doped LTO nanosheets is based on the enhanced dissociation of H2O in the presence of target NO2 molecules, leading to the generation of more protons for ion conduction. This also resulted in a greater resistance drop and thus a larger sensing response at elevated humidity levels. Our work demonstrates that proton-conductive oxide materials are promising gas-sensing materials under humid conditions. Full article
(This article belongs to the Special Issue Advanced Sensors in Atomic Level)
Show Figures

Figure 1

Figure 1
<p>Transmission electron microscope images of (<b>a</b>) LTO and (<b>b</b>) 0.7% Y-doped LTO. (<b>c</b>) High resolution transmission electron microscope image and (<b>d</b>) energy-dispersive X-ray mapping of 0.7% Y-doped LTO nanosheets. (<b>e</b>) X-ray diffraction patterns of Y-doped LTO nanosheets compared to non-doped LTO.</p>
Full article ">Figure 2
<p>High-resolution (<b>a</b>) La 3d, (<b>b</b>) Ti 2p, (<b>c</b>) Y 3d, and (<b>d</b>) O 1s X-ray photoelectron spectroscopy spectra of 0.7% Y-doped LTO nanosheets.</p>
Full article ">Figure 3
<p>(<b>a</b>) Response–recovery curves of 0.7% Y-doped LTO nanosheets to NO<sub>2</sub> of increasing concentrations under 43% RH. (<b>b</b>) Comparison of responses of Y-doped LTO (doping levels of 0.3%, 0.7%, and 1.2%) toward NO<sub>2</sub> of different concentrations at 43% RH. (<b>c</b>) Response–recovery curves of 0.7% Y-doped LTO nanosheets to NO<sub>2</sub> of increasing concentrations under 75% RH. (<b>d</b>) Comparison of responses of 0.7% Y-doped LTO nanosheets toward NO<sub>2</sub> of different concentrations under 43% RH, 60% RH, and 75% RH. (<b>e</b>) Cyclic stability test of 0.7% Y-doped LTO nanosheets in response to 5 ppm NO<sub>2</sub> gas under 75% RH. (<b>f</b>) Sensing responses of 0.7% Y-doped LTO nanosheets towards NO<sub>2</sub> and other gasses at a concentration of 10 ppm under 75% RH.</p>
Full article ">Figure 4
<p>(<b>a</b>) Nyquist plots of 0.7% Y-doped LTO in D<sub>2</sub>O and H<sub>2</sub>O atmospheres. (<b>b</b>) Estimation of the proton conduction activation energy of 0.7% Y-doped LTO under 43% and 75% relative humidity. (<b>c</b>) Schematic illustration of the NO<sub>2</sub>-sensing mechanism of YLTO.</p>
Full article ">Figure 5
<p>Nyquist plots of LTO and YLTO nanosheets with different Y doping concentrations under (<b>a</b>) 43% relative humidity, (<b>b</b>) 60% relative humidity, and (<b>c</b>) 75% relative humidity. (<b>d</b>) Comparison of conductivity of various Y-doped LTO nanosheets under 43% relative humidity, 60% relative humidity, and 75% relative humidity conditions.</p>
Full article ">
12 pages, 1302 KiB  
Article
Theoretical Analysis of Power Conversion Efficiency of Lead-Free Double-Perovskite Cs2TiBr6 Solar Cells with Different Hole Transport Layers
by Vivek Bhojak and Praveen Kumar Jain
Eng 2025, 6(2), 28; https://doi.org/10.3390/eng6020028 - 1 Feb 2025
Abstract
In recent years, there has been significant investigation into the high efficiency of perovskite solar cells. These cells have the capacity to attain efficiencies above 14%. As the perovskite materials that include lead pose a substantial environmental risk, components that are free from [...] Read more.
In recent years, there has been significant investigation into the high efficiency of perovskite solar cells. These cells have the capacity to attain efficiencies above 14%. As the perovskite materials that include lead pose a substantial environmental risk, components that are free from lead are used during the process of solar cell development. In this work, we use a lead-free double-perovskite material, namely Cs2TiBr6, as the main absorbing layer in perovskite solar cells to enhance power conversion efficiency (PCE). This work is centered on the development of solar cell structures with materials such as an ETL (electron transport layer) and an HTL (hole transport layer) to enhance the PCE. In this theoretical work, we perform simulations and analysis on double-perovskite Cs2TiBr6 to assess its efficacy as an absorber material in various HTLs like Cu2O and CuI, with a fixed ETL of C60 using SCAPS (Solar Cell Capacitance Simulator, SCAPS 3.3.10) Software. This is a one-dimensional solar cell simulation program. In this work, the thickness of the double-perovskite material is also varied between 0.2 and 2.0 µm, and its efficiency is observed. The effect of temperature variation on efficiency in the range of 300 K to 350 K is observed. The effect of defect density on efficiency is also observed in the range of 1 × 1011 to 1 × 1016. In this theoretical work, perovskite solar cells, including their absorbing layer, demonstrate outstanding ETLs and HTLs, respectively. As a result, the cells’ achieved PCE is improved. This work demonstrates the effectiveness of this lead-free double-perovskite structure that absorbs light in perovskite solar cells. Full article
Show Figures

Figure 1

Figure 1
<p>Structure diagram of designed solar cell: (<b>a</b>) C60/Cs<sub>2</sub>TiBr<sub>6</sub>/Cu<sub>2</sub>O; (<b>b</b>) C60/Cs<sub>2</sub>TiBr<sub>6</sub>/CuI; (<b>c</b>) energy level diagram of proposed structure C60/Cs<sub>2</sub>TiBr<sub>6</sub>/Cu<sub>2</sub>O; (<b>d</b>) energy level diagram of proposed structure C60/Cs<sub>2</sub>TiBr<sub>6</sub>/CuI.</p>
Full article ">Figure 2
<p>Methodology for the optimization of the parameters.</p>
Full article ">Figure 3
<p>Result of C60/Cs<sub>2</sub>TiBr<sub>6</sub>/Cu<sub>2</sub>O structure with thickness variation in main perovskite layer; (<b>a</b>) current density (Jsc) obtained at different voltages; (<b>b</b>) open-circuit voltage (Voc); (<b>c</b>) short-circuit current density (Jsc); (<b>d</b>) quantum efficiency; (<b>e</b>) Jsc with respect to different Voc (<b>f</b>) effect on PCE with temperature variation.</p>
Full article ">Figure 4
<p>Result of C60/Cs<sub>2</sub>TiBr<sub>6</sub>/CuI structure with thickness variation in main perovskite layer: (<b>a</b>) current density (Jsc) obtained at different voltages; (<b>b</b>) open-circuit voltage (Voc); (<b>c</b>) short-circuit current density (Jsc); (<b>d</b>) quantum efficiency; (<b>e</b>) Jsc with respect to different Voc; (<b>f</b>) Effect on PCE with temperature variation.</p>
Full article ">Figure 5
<p>Effect of defect density variation on PCE of structure (<b>a</b>): C60/Cs<sub>2</sub>TiBr<sub>6</sub>/Cu<sub>2</sub>O; (<b>b</b>): C60/Cs<sub>2</sub>TiBr<sub>6</sub>/CuI at thickness of 1 µm.</p>
Full article ">
Back to TopTop