[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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
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
remove_circle_outline
remove_circle_outline

Search Results (3,205)

Search Parameters:
Keywords = co-adsorption

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 4064 KiB  
Article
Investigation of the Photocatalytic Activity of Copper-Modified Commercial Titania (P25) in the Process of Carbon Dioxide Photoreduction
by Konrad Sebastian Sobczuk, Iwona Pełech, Daniel Sibera, Piotr Staciwa, Agnieszka Wanag, Ewa Ekiert, Joanna Kapica-Kozar, Katarzyna Ćmielewska, Ewelina Kusiak-Nejman, Antoni Waldemar Morawski and Urszula Narkiewicz
Materials 2024, 17(24), 6139; https://doi.org/10.3390/ma17246139 (registering DOI) - 15 Dec 2024
Viewed by 302
Abstract
Abstract: The photocatalytic reduction of CO2 to useful products is an area of active research because it shows a potential to be an efficient tool for mitigating climate change. This work investigated the modification of titania with copper(II) nitrate and its [...] Read more.
Abstract: The photocatalytic reduction of CO2 to useful products is an area of active research because it shows a potential to be an efficient tool for mitigating climate change. This work investigated the modification of titania with copper(II) nitrate and its impact on improving the CO2 reduction efficiency in a gas-phase batch photoreactor under UV–Vis irradiation. The investigated photocatalysts were prepared by treating P25-copper(II) nitrate suspensions (with various Cu2+ concentrations), alkalized with ammonia water, in a microwave-assisted solvothermal reactor. The titania-based photocatalysts were characterized by SEM, EDS, ICP-OES, XRD and UV-Vis/DR methods. Textural properties were measured by the low-temperature nitrogen adsorption/desorption studies at 77 K. P25 photocatalysts modified with copper(II) nitrate used in the process of carbon dioxide reduction allowed for a higher efficiency both for the photocatalytic reduction of CO2 to CH4 and for the photocatalytic water decomposition to hydrogen as compared to a reference. Similarly, modified samples showed significantly higher selectivity towards methane in the CO2 conversion process than the unmodified sample (a change from 30% for a reference sample to 82% for the P25-R-Cu-0.1 sample after the 6 h process). It was found that smaller loadings of Cu are more beneficial for increasing the photocatalytic activity of a sample. Full article
(This article belongs to the Special Issue Advances in Photocatalyst Materials and Green Chemistry)
19 pages, 2199 KiB  
Article
Assessment of the Effect of Multiple Processing of PHBV–Ground Buckwheat Hull Biocomposite on Its Functional and Mechanical Properties
by Grzegorz Janowski, Marta Wójcik, Wiesław Frącz, Łukasz Bąk and Grażyna Ryzińska
Materials 2024, 17(24), 6136; https://doi.org/10.3390/ma17246136 (registering DOI) - 15 Dec 2024
Viewed by 360
Abstract
The influence of the addition of ground buckwheat hulls on the properties of biocomposite on the basis of 3-hydroxybutyrate-co-3-hydroxyvalerate (PHBV) is presented here. The changes in the material after repeated reprocessing—up to five recycling cycles—are written in the paper. Analysis of the shrinkage, [...] Read more.
The influence of the addition of ground buckwheat hulls on the properties of biocomposite on the basis of 3-hydroxybutyrate-co-3-hydroxyvalerate (PHBV) is presented here. The changes in the material after repeated reprocessing—up to five recycling cycles—are written in the paper. Analysis of the shrinkage, water adsorption, selected mechanical properties, tensile impact strength, hardness and the microstructure of the surface layer was performed. The results show that the application of the buckwheat hulls into the biopolymer decreases the material shrinkage. It improves the material dimensional stability, as well as increases the water adsorption in the wake of the hydrophobic properties of the filler. The addition of the natural filler also leads to an increase in composite stiffness. The decrease in the tensile impact strength and the elongation at break is also noted. The reprocessing of the biocomposite initially led to a decrease in its mechanical properties, but the results stabilized after further processing cycles. This indicates the improvement of the microstructure homogeneity. The microscopic analysis shows that buckwheat hull particles were better embedded in the matrix after recycling. The increase in hardness was also noted. The PHBV–ground buckwheat hull biocomposite is characterized by stable mechanical properties and by recycling resistance, which makes it a promising material in terms of the sustainable development. Full article
Show Figures

Figure 1

Figure 1
<p>Specimen with marked places used for the determination of selected parameters ((1–9)–number of measuring point).</p>
Full article ">Figure 2
<p>The volumetric shrinkage for biocomposites used in the research.</p>
Full article ">Figure 3
<p>The water adsorption for GBH–PHBV biopolymer composites (samples 0–5x) and for the pure PHBV.</p>
Full article ">Figure 4
<p>The graphical representation of the tensile impact strength for the biocomposite originally produced (“0” sample) and after the subsequent recycling cycles (1x–5x) in comparison to the pure PHBV.</p>
Full article ">Figure 5
<p>The stress–strain curves for the pure PHBV biopolymer and for the GBH–PHBV biocomposite (samples 0–5x).</p>
Full article ">Figure 6
<p>The graphical representation of hardness in four areas for composite originally produced (0 sample) and after the subsequent recycling cycles (1x–5x).</p>
Full article ">Figure 7
<p>Photographs of representative probes for the composite originally produced (“0”) and for ones after the subsequent recycling cycles (1x–5x) in particular measurement areas.</p>
Full article ">
11 pages, 1686 KiB  
Article
Theoretical Investigation of Single-Atom Catalysts for Hydrogen Evolution Reaction Based on Two-Dimensional Tetragonal Mo3C2
by Bo Xue, Qingfeng Zeng, Shuyin Yu and Kehe Su
Materials 2024, 17(24), 6134; https://doi.org/10.3390/ma17246134 (registering DOI) - 15 Dec 2024
Viewed by 298
Abstract
Developing highly efficient and cost-competitive electrocatalysts for the hydrogen evolution reaction (HER), which can be applied to hydrogen production by water splitting, is of great significance in the future of the zero-carbon economy. Here, by means of first-principles calculations, we have scrutinized the [...] Read more.
Developing highly efficient and cost-competitive electrocatalysts for the hydrogen evolution reaction (HER), which can be applied to hydrogen production by water splitting, is of great significance in the future of the zero-carbon economy. Here, by means of first-principles calculations, we have scrutinized the HER catalytic capacity of single-atom catalysts (SACs) by embedding transition-metal atoms in the C and Mo vacancies of a tetragonal Mo3C2 slab, where the transition-metal atoms refer to Ti, V, Cr, Mn, Fe, Co, Ni and Cu. All the Mo3C2-based SACs exhibit excellent electrical conductivity, which is favorable to charge transfer during HER. An effective descriptor, Gibbs free energy difference (ΔGH*) of hydrogen adsorption, is adopted to evaluate catalytic ability. Apart from SACs with Cr, Mn and Fe located at C vacancies, all the other SACs can act as excellent catalysts for HER. Full article
(This article belongs to the Special Issue Advances in Multicomponent Catalytic Materials)
Show Figures

Figure 1

Figure 1
<p>The structure and active sites of Mo<sub>3</sub>C<sub>2</sub>-based single-atom catalysts, where transition-metal atoms (Ti, V, Cr, Mn, Fe, Co, Ni and Cu) are located at a (<b>a</b>) C-vacancy or (<b>b</b>) Mo<sub>surf</sub>-vacancy.</p>
Full article ">Figure 2
<p>Total and partial density of states of (<b>a</b>) (4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub>, (<b>b</b>) Ti@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub>, (<b>c</b>) V@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub>, (<b>d</b>) Cr@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub>, (<b>e</b>) Mn@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub>, (<b>f</b>) Fe@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub>, (<b>g</b>) Co@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub>, (<b>h</b>) Ni@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub> and (<b>i</b>) Cu@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>C</sub>. The Fermi level is set to 0 eV and marked with the dashed line.</p>
Full article ">Figure 3
<p>Total and partial density of states of (<b>a</b>) (4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub>, (<b>b</b>) Ti@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub>, (<b>c</b>) V@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub>, (<b>d</b>) Cr@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub>, (<b>e</b>) Mn@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub>, (<b>f</b>) Fe@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub>, (<b>g</b>) Co@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub>, (<b>h</b>) Ni@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub> and (<b>i</b>) Cu@(4 × 4)-Mo<sub>3</sub>C<sub>2</sub>-V<sub>surf-Mo</sub>. The Fermi level is set to 0 eV and marked with the dashed line.</p>
Full article ">
17 pages, 2748 KiB  
Article
Exploring the Influence of Biochar-Supported Nano-Iron Oxide on Phosphorus Speciation Transformation and Bacterial Community Structure in Aerobic Pig Manure Composting Processes
by Ning Yuan, Kang Wang, Mengyue Liang, Jia Zhou and Rui Yu
Microorganisms 2024, 12(12), 2593; https://doi.org/10.3390/microorganisms12122593 (registering DOI) - 14 Dec 2024
Viewed by 395
Abstract
Existing studies have demonstrated the positive effects of nano-sized iron oxide on compost maturity, yet the impact of nano-sized iron oxide on phosphorus speciation and bacterial communities during the composting process remains unclear. In this study, pig manure and straw were used as [...] Read more.
Existing studies have demonstrated the positive effects of nano-sized iron oxide on compost maturity, yet the impact of nano-sized iron oxide on phosphorus speciation and bacterial communities during the composting process remains unclear. In this study, pig manure and straw were used as raw materials, with biochar-supported nano-sized iron oxide (BC-Fe3O4NPs) as an additive and calcium peroxide (CaO2) as a co-agent, to conduct an aerobic composting experiment with pig manure. Four treatments were tested: CK (control), F1 (1% BC-Fe3O4NPs), F2 (5% BC-Fe3O4NPs), and F3 (5% BC-Fe3O4NPs + 5% CaO2). Key findings include the following. (1) BC-Fe3O4NPs increased compost temperatures, with F3 reaching 61℃; F1 showed optimal maturity (C/N ratio: 12.90). (2) BC-Fe3O4NPs promoted stable phosphorus forms; Residual-P proportions were higher in F1, F2, and F3 (25.81%, 51.16%, 51.68%) than CK (19.32%). (3) Bacterial phyla Firmicutes, Actinobacteria, and Proteobacteria dominated. BC-Fe3O4NPs altered community composition, especially on day 7. Firmicutes dominated CK, F1, and F3; Proteobacteria dominated F2. At the genus level, day 7 showed Corynebacterium (CK), Clostridum (F1, F3), and Caldibacillus (F2) as predominant. (4) Pearson correlation analysis revealed shifted correlations between phosphorus forms and bacterial phyla after BC-Fe3O4NPs addition. Firmicutes positively correlated with NaOH-OP in F1 during the thermophilic phase, facilitating phosphate release and adsorption by BC-Fe3O4NPs. The significance of correlations diminished with increasing additive concentration; in F3, all phyla positively correlated with various phosphorus forms. Full article
(This article belongs to the Section Microbial Biotechnology)
Show Figures

Figure 1

Figure 1
<p>SEM images of BC (<b>a</b>) and BC-Fe<sub>3</sub>O<sub>4</sub>NPs (<b>b</b>).</p>
Full article ">Figure 2
<p>Mapping of BC-Fe<sub>3</sub>O<sub>4</sub>NPs.</p>
Full article ">Figure 3
<p>Changes in physical and chemical properties of compost in different periods. Note: (<b>a</b>) Temperature, (<b>b</b>) pH, (<b>c</b>) Electrical Conductivity (EC), (<b>d</b>) Carbon to Nitrogen Ratio (C/N), (<b>e</b>) Germination Index and (<b>f</b>) Total Phosphorus.CK: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 0%; F1: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 1%; F2: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5%; F3: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5% + CaO<sub>2</sub> 5%.</p>
Full article ">Figure 4
<p>Changes of phosphorus components. Note: (<b>a</b>) CK: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 0%; (<b>b</b>) F1: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 1%; (<b>c</b>) F2: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5%;(<b>d</b>) F3: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5% + CaO<sub>2</sub> 5%.</p>
Full article ">Figure 5
<p>Changes of bacterial communities at phyla (<b>a</b>) and genus (<b>b</b>) levels during composting. Note: CK: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 0%; F1: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 1%; F2: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5%; F3: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5% + CaO<sub>2</sub> 5%. 0, 7, 28, and 50 represent the composting time points.</p>
Full article ">Figure 6
<p>NMDS analysis of metagenomic sequencing data for each process. Note: CK: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 0%; F1: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 1%; F2: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5%; F3: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5% + CaO<sub>2</sub> 5%. 0, 7, 28, and 50 represent the composting time points.</p>
Full article ">Figure 7
<p>Correlation between the top 10 relative abundances of bacterial phyla and phosphorus speciation at the phylum level. Note: CK: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 0%; F1: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 1%; F2: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5%; F3: BC-Fe<sub>3</sub>O<sub>4</sub>NPs 5% + CaO<sub>2</sub> 5%. IP and OP denote the inorganic and organic forms of phosphorus, respectively. H<sub>2</sub>O-IP and H<sub>2</sub>O-OP represent the phosphorus fractions extracted using H<sub>2</sub>O, while NaHCO<sub>3</sub>-IP and NaHCO<sub>3</sub>-OP signify the phosphorus fractions extracted with NaHCO<sub>3</sub>. Similarly, NaOH-IP and NaOH-OP indicate the phosphorus forms obtained through NaOH extraction, and HCl-IP and HCl-OP represent those extracted using HCl. Lastly, Residual-P refers to the phosphorus remaining in its residual form. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
16 pages, 10766 KiB  
Article
Investigating the Impact of Polymers on Clay Flocculation and Residual Oil Behaviour Using a 2.5D Model
by Xianda Sun, Yuchen Wang, Qiansong Guo, Zhaozhuo Ouyang, Chengwu Xu, Yangdong Cao, Tao Liu and Wenjun Ma
Polymers 2024, 16(24), 3494; https://doi.org/10.3390/polym16243494 (registering DOI) - 14 Dec 2024
Viewed by 284
Abstract
In the process of oilfield development, the surfactant–polymer (SP) composite system has shown significant effects in enhancing oil recovery (EOR) due to its excellent interfacial activity and viscoelastic properties. However, with the continuous increase in the volume of composite flooding injection, a decline [...] Read more.
In the process of oilfield development, the surfactant–polymer (SP) composite system has shown significant effects in enhancing oil recovery (EOR) due to its excellent interfacial activity and viscoelastic properties. However, with the continuous increase in the volume of composite flooding injection, a decline in injection–production capacity (I/P capacity) has been observed. Through the observation of frozen core slices, it was found that during the secondary composite flooding (SCF) process, a large amount of residual oil in the form of intergranular adsorption remained in the core pores. This phenomenon suggests that the displacement efficiency of the composite flooding may be affected. Research has shown that polymers undergo flocculation reactions with clay minerals (such as kaolinite, Kln) in the reservoir, leading to the formation of high-viscosity mixtures of migrating particles and crude oil (CO). These high-viscosity mixtures accumulate in local pores, making it difficult to further displace them, which causes oil trapping and negatively affects the overall displacement efficiency of secondary composite flooding (SCF). To explore this mechanism, this study used a microscopic visualization displacement model (MVDM) and microscopy techniques to observe the migration of particles during secondary composite flooding. By using kaolinite water suspension (Kln-WS) to simulate migrating particles in the reservoir, the displacement effects of the composite flooding system on the kaolinite water suspension, crude oil, and their mixtures were observed. Experimental results showed that the polymer, acting as a flocculant, promoted the flocculation of kaolinite during the displacement process, thereby increasing the viscosity of crude oil and affecting the displacement efficiency of secondary composite flooding. Full article
(This article belongs to the Section Polymer Applications)
Show Figures

Figure 1

Figure 1
<p>Inter-particle adsorption of residual oil diagram.</p>
Full article ">Figure 2
<p>Inter-particle adsorption of residual oil in (<b>a</b>) CT 3D image and (<b>b</b>) fluorescence photograph.</p>
Full article ">Figure 3
<p>Flocculation diagram. (<b>a</b>) Kaolinite water suspension. (<b>b</b>) Formation of flocs. (<b>c</b>) Precipitation.</p>
Full article ">Figure 4
<p>Experimental apparatus diagram. 1 Micro-injection pump. 2 Injection syringe. 3 Microscope. 4 Valve. 5 Micro-visualization model holder. 6 Micro-visualization model. 7 Heating plate. 8 Parallel light source. 9 Liquid recovery collection container. 10 Image acquisition computer.</p>
Full article ">Figure 5
<p>Cast thin sheet. (<b>a</b>) Laser confocal scanning injection thin sheet; (<b>b</b>) vectorized image after vectorization by software(Imagine v1.7.1).</p>
Full article ">Figure 6
<p>Microscopic displacement etching model.</p>
Full article ">Figure 7
<p>Kaolinite aqueous suspension (30% concentration), transmitted light photograph, temperature 46 °C.</p>
Full article ">Figure 8
<p>The oil displacement effect images from Experiment 1 are as follows. (<b>a</b>) Waterflooding (10 PV) transmitted light photograph. (<b>b</b>) Binary composite flooding (10 PV) transmitted light photograph. (<b>c</b>) Waterflooding (20 PV) transmitted light photograph. (<b>d</b>) Binary composite flooding (20 PV) transmitted light photograph.</p>
Full article ">Figure 9
<p>The oil displacement effect images from Experiment 2 are as follows. (<b>a</b>) Binary composite flooding (10 PV) displacement effect. (<b>b</b>) Binary composite flooding (20 PV) displacement effect. (<b>c</b>) Binary composite flooding (80 PV) displacement effect.</p>
Full article ">Figure 10
<p>The oil displacement effect images from Experiment 3 are as follows. (<b>a</b>) Binary composite flooding (10 PV) displacement effect. (<b>b</b>) Binary composite flooding (20 PV) displacement effect. (<b>c</b>) Binary composite flooding (80 PV) displacement effect.</p>
Full article ">Figure 11
<p>The distribution of the mixed solution is shown as follows: (<b>a</b>) transmitted light photograph and (<b>b</b>) orthogonal light photograph.</p>
Full article ">Figure 12
<p>The schematic of residual oil distribution types is as follows. 1 Pore surface film-type. 2 Grain adsorption-type. 3 Corner-type. 4 Throat-type. 5 Cluster-type. 6 Intergranular adsorption-type.</p>
Full article ">Figure 13
<p>The residual oil distribution types are as follows. (<b>a</b>) Cluster-type. (<b>b</b>) Throat-type. (<b>c</b>) Pore surface film-type. (<b>d</b>) Corner-type. (The arrow points to the location where the residual oil is located).</p>
Full article ">
21 pages, 4985 KiB  
Article
DSSCs Sensitized with Phenothiazine Derivatives Containing 1H-Tetrazole-5-acrylic Acid as an Anchoring Unit
by Muhammad Faisal Amin, Paweł Gnida, Jan Grzegorz Małecki, Sonia Kotowicz and Ewa Schab-Balcerzak
Materials 2024, 17(24), 6116; https://doi.org/10.3390/ma17246116 (registering DOI) - 14 Dec 2024
Viewed by 246
Abstract
Phenothiazine-based photosensitizers bear the intrinsic potential to substitute various expensive organometallic dyes owing to the strong electron-donating nature of the former. If coupled with a strong acceptor unit and the length of N-alkyl chain is appropriately chosen, they can easily produce high efficiency [...] Read more.
Phenothiazine-based photosensitizers bear the intrinsic potential to substitute various expensive organometallic dyes owing to the strong electron-donating nature of the former. If coupled with a strong acceptor unit and the length of N-alkyl chain is appropriately chosen, they can easily produce high efficiency levels in dye-sensitized solar cells. Here, three novel D-A dyes containing 1H-tetrazole-5-acrylic acid as an acceptor were synthesized by varying the N-alkyl chain length at its phenothiazine core and were exploited in dye-sensitized solar cells. Differential scanning calorimetry showed that the synthesized phenothiazine derivatives exhibited behavior characteristic of molecular glasses, with glass transition and melting temperatures in the range of 42–91 and 165–198 °C, respectively. Based on cyclic and differential pulse voltammetry measurements, it was evident that their lowest unoccupied molecular orbital (LUMO) (−3.01–−3.14 eV) and highest occupied molecular orbital (HOMO) (−5.28–−5.33 eV) values were fitted to the TiO2 conduction band and the redox energy of I/I3 in electrolyte, respectively. The experimental results were supported by density functional theory, which was also utilized for estimation of the adsorption energy of the dyes on the TiO2 and its size. Finally, the compounds were tested in dye-sensitized solar cells, which were characterized based on current–voltage measurements. Additionally, for the compound giving the best photovoltaic response, the efficiency of the DSSCs was optimized by a photoanode modification involving the use of cosensitization and coadsorption approaches and the introduction of a blocking layer. Subsequently, two types of tandem dye-sensitized solar cells were constructed, which resulted in an increase in photovoltaic efficiency to 6.37%, as compared to DSSCs before modifications, with a power conversion value of 2.50%. Full article
(This article belongs to the Special Issue Advances in Solar Cell Materials and Structures—Second Edition)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H NMR spectra of (<b>a</b>) <b>PETA</b>, (<b>b</b>) <b>PBTA</b>, and (<b>c</b>) <b>POTA</b>.</p>
Full article ">Figure 1 Cont.
<p><sup>1</sup>H NMR spectra of (<b>a</b>) <b>PETA</b>, (<b>b</b>) <b>PBTA</b>, and (<b>c</b>) <b>POTA</b>.</p>
Full article ">Figure 2
<p>(<b>a</b>) DSC thermogram of <b>PETA</b>. (<b>b</b>) Thermal investigation data of compounds starting from phenothiazine.</p>
Full article ">Figure 3
<p>The voltammograms of the (<b>a</b>) reduction and oxidation process measured in the cyclic voltammetry method and (<b>b</b>) voltammograms of the oxidation process measured in the differential pulse voltammetry method (GC, 0.1 mol/dm<sup>3</sup> Bu<sub>4</sub>NPF<sub>6</sub> in DMF, 100 mV/s; the dashed lines mean reduction, and the solid lines mean oxidation).</p>
Full article ">Figure 4
<p>Molecular electrostatic potential surfaces on the molecules of the dyes (scale range −7.03 × 10<sup>−2</sup> (red) to 7.03 × 10<sup>−2</sup> (blue) neural and −0.19 a.u. (red) to 0.19 a.u (blue) anionic form).</p>
Full article ">Figure 5
<p>Adsorption of the dyes on Ti<sub>30</sub>O<sub>66</sub>H<sub>12</sub> cluster calculated in acetonitrile solutions (values calculated in the gas phase are given in brackets).</p>
Full article ">Figure 5 Cont.
<p>Adsorption of the dyes on Ti<sub>30</sub>O<sub>66</sub>H<sub>12</sub> cluster calculated in acetonitrile solutions (values calculated in the gas phase are given in brackets).</p>
Full article ">Figure 6
<p>UV–Vis absorption spectra of the dyes (<b>a</b>) in solution form (c = 2 × 10<sup>−5</sup> mol dm<sup>−3</sup>), (<b>b</b>) adsorbed on TiO<sub>2</sub> surface, and (<b>c</b>) PL spectra of the dyes in solution form.</p>
Full article ">Figure 7
<p>Block diagram of ongoing research on DSSCs.</p>
Full article ">Figure 8
<p>(<b>a</b>) J–V curves for DSSCs sensitized with PTZ dyes and N719 with and without BL. (<b>b</b>) Schematic energy level diagram of dyes under vacuum in terms of eV.</p>
Full article ">Figure 9
<p>J–V characteristics of tandem DSSCs with (<b>a</b>) FTO/BL/TiO<sub>2</sub>@<b>POTA</b> photoanode in top cell, (<b>b</b>) FTO/BL/TiO<sub>2</sub>@<b>POTA</b> photoanode in bottom cell.</p>
Full article ">Scheme 1
<p>Scheme of the designed dyes synthesis. (<b>i</b>) Acetone, TBAI, reflux 24 h. (<b>ii</b>) DMF, POCl<sub>3</sub>, 1,2-dichloroethane, reflux 24 h. (<b>iii</b>) Diethylamine, 1H-tetrazole-5-acetic acid, CH<sub>3</sub>CN, reflux 24 h.</p>
Full article ">
12 pages, 6094 KiB  
Article
A Fluorine-Functionalized Tb(III)–Organic Framework for Ba2+ Detection
by Yang Zhang, Hua Tan, Jiaping Zhu, Linhai Duan, Yuchi Ding, Fenglan Liang, Yongshi Li, Xinteng Peng, Ruomei Jiang, Jiaxin Yu, Jianjiong Fan, Yuhang Chen, Rimeng Chen and Deyun Ma
Molecules 2024, 29(24), 5903; https://doi.org/10.3390/molecules29245903 (registering DOI) - 13 Dec 2024
Viewed by 339
Abstract
The development of lanthanide–organic frameworks (Ln-MOFs) using for luminescence sensing and selective gas adsorption applications is of great significance from an energy and environmental perspective. This study reports the solvothermal synthesis of a fluorine-functionalized 3D microporous Tb-MOF with a face-centered cubic (fcu [...] Read more.
The development of lanthanide–organic frameworks (Ln-MOFs) using for luminescence sensing and selective gas adsorption applications is of great significance from an energy and environmental perspective. This study reports the solvothermal synthesis of a fluorine-functionalized 3D microporous Tb-MOF with a face-centered cubic (fcu) topology constructed from hexanuclear clusters (Tb6O30) bridged by fdpdc ligands, formulated as {[Tb6(fdpdc)6(μ3-OH)8(H2O)6]·4DMF}n (1), (fdpdc = 3-fluorobiphenyl-4,4′-dicarboxylate). Complex 1 displays a 3D framework with the channel of 7.2 × 7.2 Å2 (measured between opposite atoms) perpendicular to the a-axis. With respect to Ba2+ cation, the framework of activated 1 (1a) exhibits high selectivity and reversibility in luminescence sensing function, with an LOD of 4.34665 mM. According to the results of simulations, compared to other small gas molecules (CO2, N2, H2, CO, and CH4), activated 1 (1a) shows a high adsorption selectivity for C2H2 at 298 K. Full article
(This article belongs to the Special Issue Inorganic Chemistry in Asia)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Single-capped tetragonal prism of Tb(III) in <b>1</b>, blue line, configuration bar; red bullet, oxygen atoms; green circle, Tb. The coordination environment of Zn(II) ions in <b>1</b>. (<b>b</b>) Hexanuclear Tb<sub>6</sub>O<sub>30</sub> unit of <b>1</b>. (<b>c</b>) View of the 3D framework packing perpendicular to the <span class="html-italic">a</span>-axis. (<b>d</b>) The <b>fcu</b> topology network of <b>1</b>.</p>
Full article ">Figure 2
<p>PXRD patterns of <b>1</b>.</p>
Full article ">Figure 3
<p>(<b>a</b>) The fluorescence excitation and emission of <b>1</b> and H<sub>2</sub>fdpdc ligand in solid state. (<b>b</b>) The cation-dependent fluorescence emission of <b>1</b>. (<b>c</b>) The luminescence intensity at 545 nm with the presence of different cations. (<b>d</b>) The luminescent emission spectra of <b>1</b> in Ba<sup>2+</sup> solutions with different concentrations.</p>
Full article ">Figure 4
<p>(<b>a</b>) The correlation between relative intensity and the addition of Ba<sup>2+</sup> ion at different concentrations over time at 545 nm. (<b>b</b>) Calculated LOD of <b>1</b> in Ba<sup>2+</sup> solutions.</p>
Full article ">Figure 5
<p>(<b>a</b>) Luminescence intensity of <b>1</b> at 545 nm dispersed in water with the addition of different metal ions. The concentration of interferential metal ions and Ba<sup>2+</sup> ion are both 1 mM. (<b>b</b>) Luminescence intensity (545 nm) of <b>1</b> during five recycling.</p>
Full article ">Figure 6
<p>(<b>a</b>) N<sub>2</sub> sorption isotherm of <b>1a</b> at 77 K (the inset shows the pore size distribution). (<b>b</b>) Room temperature adsorption isotherms of C<sub>2</sub>H<sub>2</sub>, CO<sub>2</sub>, N<sub>2</sub>, H<sub>2</sub>, CO, and CH<sub>4</sub> based on the simulation. (<b>c</b>) Isosteric heat of adsorption of C<sub>2</sub>H<sub>2</sub> for <b>1</b> calculated based on molecular simulation isotherms at 298 K. (<b>d</b>) Selectivities for C<sub>2</sub>H<sub>2</sub> in the equimolar C<sub>2</sub>H<sub>2</sub>/CO<sub>2</sub>, C<sub>2</sub>H<sub>2</sub>/N<sub>2</sub>, C<sub>2</sub>H<sub>2</sub>/H<sub>2</sub>, C<sub>2</sub>H<sub>2</sub>/CO, and C<sub>2</sub>H<sub>2</sub>/CH<sub>4</sub> at 298 K from the simulation.</p>
Full article ">
11 pages, 4041 KiB  
Article
Enhanced Performance of Ce Doping VW/Ti Catalysts for Synergistic Catalytic Removal of NOx and Chlorobenzene
by Na Zhu, Lingyu Yu, Pengpeng Xu and Yang Deng
Catalysts 2024, 14(12), 919; https://doi.org/10.3390/catal14120919 - 12 Dec 2024
Viewed by 438
Abstract
Nitrogen oxides (NOx) and chlorobenzene (CB) released during waste incineration and iron ore sintering pose significant threats to both the atmosphere and human health, necessitating effective control measures. Vanadium-based catalysts are commonly employed for the simultaneous control of NOx and [...] Read more.
Nitrogen oxides (NOx) and chlorobenzene (CB) released during waste incineration and iron ore sintering pose significant threats to both the atmosphere and human health, necessitating effective control measures. Vanadium-based catalysts are commonly employed for the simultaneous control of NOx and CB; however, their catalytic performance requires further enhancement. In this study, the NH3-SCR activity and CB catalytic oxidation (CBCO) activity were significantly enhanced by doping the V10W/Ti catalyst with Ce. During the multi-pollutant control (MPC) reaction, the optimized 15CeV10W/Ti catalyst demonstrated NOx conversion approaching 100% and N2 selectivity exceeding 95% at temperatures between 210 and 450 °C. Additionally, it achieved CB conversion nearing 100% and CO2 selectivity above 80% at temperatures above 350 °C. These results were markedly superior to those of the conventional commercial 1%V2O5–10%WO3/TiO2 catalyst. Characterization studies indicated that the 15CeV10W/Ti catalyst possessed improved redox performance and more acidic sites. In the MPC reaction, the declined CBCO activity, compared to the CB separate oxidation, can be attributed primarily to the competitive adsorption of NH3 with CB. Conversely, the observed decrease in NOx conversion at lower temperatures was primarily due to the suppression of the oxidation of NO to NO2 by CB. Full article
(This article belongs to the Section Catalytic Materials)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of 15CeV10W/Ti-X catalysts.</p>
Full article ">Figure 2
<p>H<sub>2</sub>-TPR profiles (<b>a</b>) and H<sub>2</sub> consumption (<b>b</b>) of 15CeV10W/Ti-X catalysts.</p>
Full article ">Figure 3
<p>XPS spectra of (<b>a</b>) V 2p, (<b>b</b>) O 1s, and (<b>c</b>) Ce 3d for 15CeV10W/Ti-X catalysts.</p>
Full article ">Figure 4
<p>NH<sub>3</sub>-TPD profiles of 15CeV10W/Ti-X catalyst.</p>
Full article ">Figure 5
<p>NH<sub>3</sub>-SCR activity of 15CeV10W/Ti-X catalysts: (<b>a</b>) NO<span class="html-italic"><sub>x</sub></span> conversion; (<b>b</b>) N<sub>2</sub> selectivity.</p>
Full article ">Figure 6
<p>V10W/Ti and 15CeV10W/Ti catalysts: (<b>a</b>) NO<span class="html-italic"><sub>x</sub></span> conversion and (<b>c</b>) N<sub>2</sub> selectivity in the presence/absence of 50 ppm CB, (<b>b</b>) CB conversion, and (<b>d</b>) CO<sub>2</sub> selectivity in the presence/absence of SCR reactants.</p>
Full article ">Figure 7
<p>Desorption of (<b>a</b>) NO and (<b>b</b>) NO<sub>2</sub> from NO + O<sub>2</sub>-TPD and NO + O<sub>2</sub> + CB-TPD over V10W/Ti and 15CeV10W/Ti catalysts.</p>
Full article ">Figure 8
<p>TPSR curves of (<b>a</b>) V10W/Ti and (<b>b</b>) 15CeV10W/Ti catalysts.</p>
Full article ">
22 pages, 8801 KiB  
Review
Modifications and Applications of Metal-Organic-Framework-Based Materials for Photocatalysis
by Weimin Ma, Liang Yu, Pei Kang, Zhiyun Chu and Yingxuan Li
Molecules 2024, 29(24), 5834; https://doi.org/10.3390/molecules29245834 - 11 Dec 2024
Viewed by 307
Abstract
Metal–organic frameworks (MOFs) represent a category of crystalline materials formed by the combination of metal ions or clusters with organic linkers, which have emerged as a prominent research focus in the field of photocatalysis. Owing to their distinctive characteristics, including structural diversity and [...] Read more.
Metal–organic frameworks (MOFs) represent a category of crystalline materials formed by the combination of metal ions or clusters with organic linkers, which have emerged as a prominent research focus in the field of photocatalysis. Owing to their distinctive characteristics, including structural diversity and configurations, significant porosity, and an extensive specific surface area, they provide a flexible foundation for various potential applications in photocatalysis. In recent years, researchers have tackled many issues in the MOF-based photocatalytic yield. However, limited light adsorption regions, lack of active sites and active species, and insufficient efficiency of photogenerated charge carrier separation substantially hinder the photocatalytic performance. In this review, we summarized the strategies to improve the photocatalytic performance and recent developments achieved in MOF and MOF-based photocatalysis, including water splitting, CO2 conversion, photocatalytic degradation of pollutants, and photocatalytic nitrogen fixation into ammonia. In conclusion, the existing challenges and prospective advancements in MOF-based photocatalysis are also discussed. Full article
(This article belongs to the Special Issue Design and Application of Periodic Frameworks)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of MOFs structure, regulation strategies, and their catalytic applications.</p>
Full article ">Figure 2
<p>(<b>a</b>) Model structure of MOF-253-Pt. (<b>b</b>) UV–vis spectra of MOF-253, MOF-253-Pt, and Pt(bpydc)Cl<sub>2</sub> and the corresponding quantum efficiencies of hydrogen evolution for MOF-253-Pt at different wavelengths. The inset shows the colors of the samples. Adapted with permission from RSC [<a href="#B50-molecules-29-05834" class="html-bibr">50</a>]. (<b>c</b>) Structure of the UiO-66 framework. (<b>d</b>) Experimental linker spectra and photographs of the pure linkers. Reproduced with permission from ACS [<a href="#B65-molecules-29-05834" class="html-bibr">65</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Coordination environment and (<b>b</b>) free energy diagram of CO<sub>2</sub>RR for MOF–Ni, MOF–Co, and MOF–Cu catalysts. (<b>c</b>) Free energy diagram of HER. (<b>d</b>) Yield of CO and H<sub>2</sub> produced by photocatalytic reduction of CO<sub>2</sub> by MOF–Cu, MOF–Co, and MOF–Ni. * = the active species. Adapted with permission from ACS [<a href="#B66-molecules-29-05834" class="html-bibr">66</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Framework structure of UiO–66. (<b>b</b>) The band alignment of UiO–66(Ce)–X, with the structure of the BDC–X linkers displayed in the in the box at lower right. Adapted with permission from ACS [<a href="#B69-molecules-29-05834" class="html-bibr">69</a>]. (<b>c</b>) Schematic diagram of photocatalytic H<sub>2</sub> generation before and after contact with MIL–125–NH<sub>2</sub> and metal oxides. (<b>d</b>) SPV spectra curves of MIL–125–NH<sub>2</sub>, MoO<sub>3</sub>/MIL–125–NH<sub>2</sub>, and V<sub>2</sub>O<sub>5</sub>/MIL–125–NH<sub>2</sub>. (<b>e</b>) Photocatalytic H<sub>2</sub> generation rates for Pt–loaded MoO<sub>3</sub>/MIL–125–NH<sub>2</sub> and V<sub>2</sub>O<sub>5</sub>/MIL–125–NH<sub>2</sub> catalysts at different positions. (<b>f</b>) Schematic diagram of photocatalytic H<sub>2</sub> production before and after combination. Adapted with permission from WILEY [<a href="#B58-molecules-29-05834" class="html-bibr">58</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) A schematic representation depicts the synthesis of Pt@UiO-66-NH<sub>2</sub> and Pt/UiO-66-NH<sub>2</sub>, highlighting the photocatalytic hydrogen production process occurring over Pt@UiO-66-NH<sub>2</sub>. (<b>b</b>) TA spectra of UiO-66-NH<sub>2</sub> (excitation at 400 nm) with TA signal given in mOD (OD: optical density). (<b>c</b>) Time-resolved PL decay profiles and (<b>d</b>) TA kinetics for UiO-66-NH<sub>2</sub>, Pt@UiO-66-NH<sub>2</sub>, and Pt/UiO-66-NH<sub>2</sub>, respectively. Reproduced with permission from WILEY [<a href="#B51-molecules-29-05834" class="html-bibr">51</a>]. (<b>e</b>) The schematic synthesis of MTV-Ti-MOF/COF. (<b>f</b>) The photocatalytic H<sub>2</sub> evolution activities. (<b>g</b>) The photocatalytic H<sub>2</sub> evolution mechanism. Reproduced with permission from WILEY [<a href="#B52-molecules-29-05834" class="html-bibr">52</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Immobilization of C<sub>60</sub> within the pores of NU–901. (<b>b</b>) The internal electric field intensity for NU–901 and C<sub>60</sub>@NU–901. (<b>c</b>) Photocatalytic hydrogen evolution rates for C<sub>3</sub>N<sub>4</sub>, NU–901, and C<sub>60</sub>@NU–901. (<b>d</b>) The photoluminescence spectra. Reproduced with permission from WILEY [<a href="#B53-molecules-29-05834" class="html-bibr">53</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Pore size distribution and morphology of MIL–53(Fe)–2OH and MOF–74–Fe catalysts. (<b>b</b>) Schematic synthesis of MIL–53(Fe)–2OH and MOF–74–Fe catalysts and the analysis Fe–3d, OH–s,p and C–s,p. (<b>c</b>) Overpotential histogram of MIL–53(Fe)–2OH, MOF–74–Fe, and IrO<sub>2</sub> catalysts before and after reaction at different current densities. (<b>d</b>) Electrokinetic current density under optimal kinetic fitting of OER. (<b>e</b>) Coverage of O*, OH*, and OOH* intermediates of MIL–53(Fe)–2OH catalyst before and after stability testing at different potentials. Reproduced with permission from WILEY [<a href="#B71-molecules-29-05834" class="html-bibr">71</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) The schematic illustration for the preparation of Ni MOLs. (<b>b</b>) CO<sub>2</sub> photoreduction performance over Ni MOLs and Co MOLs in pure CO<sub>2</sub> and diluted CO<sub>2</sub> (10%). (<b>c</b>) The proposed mechanism delineates the conversion of CO<sub>2</sub> to CO via Ni MOLs using [Ru(bpy)<sub>3</sub>]<sup>2+</sup> and TEOA under visible light. Reproduced with permission from WILEY [<a href="#B72-molecules-29-05834" class="html-bibr">72</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) Schematic illustration of the morphology and facet over MIL–125–NH<sub>2</sub>(Ti). (<b>b</b>) SEM images of as–synthesized MIL–125–NH<sub>2</sub>(Ti) with different shape. Scale bar: 500 nm. (<b>c</b>) HOMO–LUMO gap of the as–synthesized MIL–125–NH<sub>2</sub>(Ti). (<b>d</b>) The yield of CO and CH<sub>4</sub> products and (<b>e</b>) effective photo–electrons after irradiation for 5 h over the as–synthesized MIL–125–NH<sub>2</sub>(Ti). Adopted with permission from Elsevier [<a href="#B73-molecules-29-05834" class="html-bibr">73</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) Three-dimensional frame structure diagram of Ni–MOF and the corresponding trinuclear [Ni<sub>3</sub>(COO)<sub>6</sub>] node. (<b>b</b>) CO<sub>2</sub>, N<sub>2</sub>, and O<sub>2</sub> adsorption curves of Pt/Ni–MOF at 25 °C, insert: CO<sub>2</sub> absorption at a pressure of 0.4 mbar (the pressure pertinent to direct air capture). (<b>c</b>) Thermal-photocatalytic yields of CO and CH<sub>4</sub> on the pristine Pt/Ni–MOF in H<sub>2</sub>: CO<sub>2</sub> (with a 1:7 ratio) and on Pt/Ni–MOF with captured ambient CO<sub>2</sub> in H<sub>2</sub>: Ar (with a 1:7 ratio), respectively. (<b>d</b>) AQYs of Pt/Ni–MOF for CO<sub>2</sub> reduction under different wavelengths. (<b>e</b>) Schematic diagram of the proposed pathway for Pt/Ni–MOF photocatalytic conversion of CO<sub>2</sub>. Reproduced with permission from WILEY [<a href="#B54-molecules-29-05834" class="html-bibr">54</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) A schematic illustration demonstrates the effective reduction of CO<sub>2</sub> to CH<sub>4</sub> using an IR–light–driven UiO–66/Co<sub>9</sub>S<sub>8</sub> composite catalyst, leveraging the advantages of both porous UiO–66 and metallic Co<sub>9</sub>S<sub>8</sub>. (<b>b</b>) CH<sub>4</sub> evolution rates of CO<sub>2</sub> photoreduction over different catalysts. (<b>c</b>) Stability assessments of UiO–66/Co<sub>9</sub>S<sub>8</sub> for photocatalytic activity under full–spectrum light irradiation. (<b>d</b>) A possible reaction pathway for photocatalytic CO<sub>2</sub> methanation, * = active site. Reproduced with permission from WILEY [<a href="#B55-molecules-29-05834" class="html-bibr">55</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) Diagrammatic representation of the synthesis process for In<sub>2</sub>O<sub>3</sub>@Pd and h-In<sub>2</sub>O<sub>3</sub>/Pd catalysts (above) as well as c-In<sub>2</sub>O<sub>3</sub>/Pd and r-In<sub>2</sub>O<sub>3</sub>/Pd catalysts (below). (<b>b</b>) The conversion of CO<sub>2</sub> and the distribution of products across four different Pd/In<sub>2</sub>O<sub>3</sub> catalysts. (<b>c</b>) Time on stream (TOS) for the four catalysts in CO<sub>2</sub> hydrogenation under standard conditions. (<b>d</b>) Schematic illustration of CO<sub>2</sub> hydrogenation to methanol. Reproduced with permission from ACS [<a href="#B59-molecules-29-05834" class="html-bibr">59</a>]. (<b>e</b>) Schematic illustration of the synthetic process of hollow-structured In<sub>2</sub>O<sub>3</sub>@ZrO<sub>2</sub> heterostructure. (<b>f</b>) Simplified model showing the main reaction mechanism on In<sub>2</sub>O<sub>3</sub>@ZrO<sub>2</sub> catalyst. Reproduced with permission from WILEY [<a href="#B60-molecules-29-05834" class="html-bibr">60</a>].</p>
Full article ">Figure 13
<p>(<b>a</b>) Schematic diagram of the preparation of the nitrogen-doped carbon layer-coated cobalt nanoparticles catalyst (Co@NC). (<b>b</b>) Nitrogen species and content distribution of catalytic pyrolysis at different temperatures. (<b>c</b>) UV–Vis–NIR absorption spectra for the synthesized Co@NC catalysts. (<b>d</b>) Monitoring of photothermal temperatures for Co@NC-700, Co/C, and uncoated Co nanoparticles under a light intensity of 3.0 W cm<sup>−2</sup>. (<b>e</b>) Photothermal CO<sub>2</sub> hydrogenation stability evaluation for the Co@NC-700 with Xe irradiation at 3.0 W cm<sup>−2</sup>. (<b>f</b>) Optimized configuration for CO adsorption on Co<sub>9</sub>@graphene, graphene, and the Co (111) surface. Reproduced with permission from WILEY [<a href="#B61-molecules-29-05834" class="html-bibr">61</a>].</p>
Full article ">Figure 14
<p>(<b>a</b>) Schematic presentation of the preparation route of MxBy heterojunctions. (<b>b</b>) Temporal analysis of photo–Fenton degradation of ENR across various reaction systems. (<b>c</b>) Time–resolved PL decay spectra. (<b>d</b>) Possible ENR catalytic degradation mechanism. Adopted with permission from Elsevier [<a href="#B56-molecules-29-05834" class="html-bibr">56</a>].</p>
Full article ">Figure 15
<p>(<b>a</b>) Schematic diagram of TCL degradation by MXFAA/MIL(Fe)-POR photo–Fenton. (<b>b</b>) Photo–Fenton TCL degradation performance. Adopted with permission from Elsevier [<a href="#B81-molecules-29-05834" class="html-bibr">81</a>]. (<b>c</b>) Schematic diagram of the synthesis route of NH<sub>2</sub>–UiO–66@DAT–HOF photocatalysts. (<b>d</b>) Efficiency of TC photodegradation in the presence of the U@H2 hybrid. (<b>e</b>) Energy band structure of the U@H2 heterojunction. Reproduced with permission from WILEY [<a href="#B57-molecules-29-05834" class="html-bibr">57</a>].</p>
Full article ">Figure 16
<p>(<b>a</b>) The synthesis of MIL–125(Ti) involves a solvothermal method followed by thermal calcination. T1 and T2 denote the different calcination temperature. MIL–125(Ti)–250 with oxygen vacancies and an increase in BET surface area. (<b>b</b>) The adsorption of an N<sub>2</sub> molecule on both free and defect modes, along with the associated binding energy. (<b>c</b>) Density of states of free and defect structure. (<b>d</b>) The associative pathway encompasses both “end–on” (vertical) and “side–on” (parallel) configurations for N<sub>2</sub> adsorption. Reproduced with permission from WILEY [<a href="#B86-molecules-29-05834" class="html-bibr">86</a>]. (<b>e</b>) Schematic illustration for the construction of ZnO@NC-Ni2 with dinuclear Ni2 sites based on ZIF–8. (<b>f</b>) Schematic diagram of possible pathways for N<sub>2</sub> reduction to NH<sub>3</sub>. Adapted with permission from ACS [<a href="#B62-molecules-29-05834" class="html-bibr">62</a>].</p>
Full article ">
11 pages, 2441 KiB  
Article
High-Pressure Gas Adsorption on Covalent Organic Framework CTF-1
by Gregory S. Deyko, Valery N. Zakharov, Lev M. Glukhov, Dmitry O. Charkin, Dmitry Yu. Kultin, Vladimir V. Chernyshev, Leonid A. Aslanov and Leonid M. Kustov
Crystals 2024, 14(12), 1066; https://doi.org/10.3390/cryst14121066 - 10 Dec 2024
Viewed by 297
Abstract
Triazine-based covalent organic framework CTF-1 was synthesized via polymerization of 1,4-dicyanobenzene in the presence of zinc chloride. Two different methods of the post-synthesis treatment of the obtained material were compared. It was demonstrated that ultrasonication effectively removes impurities from CTF-1. Adsorption of hydrocarbon [...] Read more.
Triazine-based covalent organic framework CTF-1 was synthesized via polymerization of 1,4-dicyanobenzene in the presence of zinc chloride. Two different methods of the post-synthesis treatment of the obtained material were compared. It was demonstrated that ultrasonication effectively removes impurities from CTF-1. Adsorption of hydrocarbon gases (methane and ethane) and carbon dioxide was measured at 298 K in a wide pressure range for the first time. Ideal selectivity and IAST values for methane/ethane and methane/CO2 pairs were calculated from the obtained isotherms. Full article
(This article belongs to the Section Macromolecular Crystals)
Show Figures

Figure 1

Figure 1
<p>XRD powder patterns of CTF-1 (blue) and CTF-1us (red).</p>
Full article ">Figure 2
<p>SEM images of obtained CTF-1 (<b>a</b>) and CTF-1us (<b>b</b>) samples, magnification is 100 times.</p>
Full article ">Figure 3
<p>Adsorption isotherms of CO<sub>2</sub>, CH<sub>4</sub>, and C<sub>2</sub>H<sub>6</sub> on the obtained samples at 298 K. Hollow symbols—CTF-1us sample after ultrasound treatment; solid symbols—CTF-1 sample without ultrasound treatment.</p>
Full article ">Figure 4
<p>IAST selectivity values for C<sub>2</sub>H<sub>6</sub>/CH<sub>4</sub> (1:9 mol) and CO<sub>2</sub>/CH<sub>4</sub> (1:9 mol) gas pairs at 298 K.</p>
Full article ">Figure 5
<p>Ideal selectivity values for C<sub>2</sub>H<sub>6</sub>/CH<sub>4</sub> and CO<sub>2</sub>/CH<sub>4</sub> gas pairs for CTF-1 samples at 298 K.</p>
Full article ">
28 pages, 8560 KiB  
Article
Methyl Mercaptan Removal from Methane Using Metal-Oxides and Aluminosilicate Materials
by Gerson Martinez-Zuniga, Samuel Antwi, Percival Soni-Castro, Olatunji Olayiwola, Maksym Chuprin, William E. Holmes, Prashanth Buchireddy, Daniel Gang, Emmanuel Revellame, Mark E. Zappi and Rafael Hernandez
Catalysts 2024, 14(12), 907; https://doi.org/10.3390/catal14120907 - 10 Dec 2024
Viewed by 536
Abstract
Methyl mercaptan is a sulfur-based chemical found as a co-product in produced natural gas and it causes corrosion in pipelines, storage tanks, catalysts, and solid adsorption beds. To improve the quality of methane produced, researchers have studied the use of metal oxides and [...] Read more.
Methyl mercaptan is a sulfur-based chemical found as a co-product in produced natural gas and it causes corrosion in pipelines, storage tanks, catalysts, and solid adsorption beds. To improve the quality of methane produced, researchers have studied the use of metal oxides and aluminum silicates as catalysts for removing mercaptan. However, there are restrictive limitations on the efficiency of metal oxides or aluminum silicates as adsorbents for this application. Therefore, this study investigated the performance of these materials in a fixed-bed reactor with simulated natural gas streams under various operating conditions. The testing procedure includes a detailed assessment of the adsorbent/catalysts by several techniques, such as Braeuer–Emmett–Teller (BET), Scanning Electron Microscope (SEM), Energy-Dispersive X-ray Spectrometry (EDS), and X-ray Photoelectron Spectroscopy. The results revealed that metal oxides such as copper, manganese, and zinc performed well in methyl mercaptan elimination. The addition of manganese, copper, and zinc oxides to the aluminum silicate surface resulted in a sulfur capacity of 1226 mg S/g of catalyst. These findings provide critical insights for the development of catalysts that combine metal oxides to increase adsorption while reducing the production of byproducts like dimethyl sulfide (DMS) and dimethyl disulfide (DMDS) during methyl mercaptan removal. Full article
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of the Select HP catalyst and CTG-ESC-011 catalyst.</p>
Full article ">Figure 2
<p>FTIR spectra for (<b>a</b>) Halloysite Pure, (<b>b</b>) Halloysite MinO, and (<b>c</b>) Halloysite Mixed.</p>
Full article ">Figure 3
<p>XPS spectra for (<b>a</b>) Select HP, (<b>b</b>) CTG-ESC-011, and (<b>c</b>) Fe 2p.</p>
Full article ">Figure 4
<p>XPS spectra for: (<b>a</b>) Halloysite MinO, (<b>b</b>) Halloysite Pure, and (<b>c</b>) Halloysite Mixed.</p>
Full article ">Figure 5
<p>Breakthrough curves for Select HP catalyst.</p>
Full article ">Figure 6
<p>Breakthrough curves for CTG-ESC-011 catalyst.</p>
Full article ">Figure 7
<p>Breakthrough curves for Halloysite Pure catalyst.</p>
Full article ">Figure 8
<p>Breakthrough curves for Halloysite MinO catalyst.</p>
Full article ">Figure 9
<p>Breakthrough curves for Halloysite Mixed catalyst.</p>
Full article ">Figure 10
<p>(<b>a</b>) Best sulfur capacities for each commercial catalyst and (<b>b</b>) Breakthrough curves for the best results of each catalyst.</p>
Full article ">Figure 11
<p>(<b>a</b>) Cu—one factor model graph for methyl mercaptan removal, and (<b>b</b>) Mg—one factor model graph for methyl mercaptan removal.</p>
Full article ">Figure 12
<p>A 3D response graph for sulfur capacity in methyl mercaptan removal versus concentrations of copper and magnesium.</p>
Full article ">Figure 13
<p>Breakthrough curves for UL-Best catalyst vs. Select HP at 200 psi, 25 °C, 36 mL/min, and 200 ppm CH<sub>3</sub>SH.</p>
Full article ">Figure 14
<p>Breakthrough curves for UL-Best catalyst vs. Select HP at different temperatures.</p>
Full article ">Figure 15
<p>SEM images and EDS spectra for (<b>a</b>,<b>c</b>) catalyst before use and (<b>b</b>,<b>d</b>) catalyst after use.</p>
Full article ">Figure 15 Cont.
<p>SEM images and EDS spectra for (<b>a</b>,<b>c</b>) catalyst before use and (<b>b</b>,<b>d</b>) catalyst after use.</p>
Full article ">Figure 16
<p>Before and After XPS spectra of (<b>a</b>) Cu 2p, (<b>b</b>) Zn 2p, (<b>c</b>) Mn 2p, and (<b>d</b>) S 2p.</p>
Full article ">Figure 17
<p>FT-IR spectrum of Raw Select HP and Spent Select HP.</p>
Full article ">Figure 18
<p>Total Ion chromatogram for Select HP catalyst at 75 °C, 200 PSI, 36 mL/min.</p>
Full article ">Figure 19
<p>Possible reaction mechanism for methyl mercaptan removal.</p>
Full article ">Figure 20
<p>Experimental setup for the methyl mercaptan removal.</p>
Full article ">Figure 21
<p>Graphical representation of methyl mercaptan removal.</p>
Full article ">
14 pages, 2868 KiB  
Article
Chemodiversity and Molecular Mechanism Between Per-/Polyfluoroalkyl Substance Complexation Behavior of Humic Substances in Landfill Leachate
by Jia Li, Haoqun Sha, Rongchuan Ye, Peipei Zhang, Shuhe Chen, Ganghui Zhu and Wenbing Tan
Water 2024, 16(23), 3527; https://doi.org/10.3390/w16233527 - 7 Dec 2024
Viewed by 627
Abstract
Landfill leachate contains a range of organic and inorganic pollutants, including per-/polyfluoroalkyl substances (PFASs), which can infiltrate into surrounding soil and groundwater through leaching processes, and can pose a threat to human health via food chains and drinking water processes. Thus, the transport [...] Read more.
Landfill leachate contains a range of organic and inorganic pollutants, including per-/polyfluoroalkyl substances (PFASs), which can infiltrate into surrounding soil and groundwater through leaching processes, and can pose a threat to human health via food chains and drinking water processes. Thus, the transport of PFASs in landfill leachate is a research hotspot in environmental science. This study investigates the complexation and adsorption mechanisms between humic substances and PFASs in landfill leachate at the molecular level. Experimental results demonstrate that the binding constant logKsv of humic substances with PFASs correlates positively with specific ultraviolet absorbance (SUVA254), absorbance ratio (A250/A365), humification index (HIX), and fluorescence index (FI), while it exhibits a negative correlation with the biological index (BIX). These findings indicate that high aromaticity is a prerequisite for molecular interactions between humic substances and PFASs, with polar functional groups further facilitating the interaction. Molecular-level analysis revealed that humic substances undergo complexation and adsorption with PFASs through hydrophobic interactions, van der Waals forces, hydrogen bonding, ionic bonding, and covalent bonding, by functional groups such as hydroxyl, aliphatic C-H bonds, aromatic C=C double bonds, amides, quinones, and ketones. Future efforts should focus on enhanced co-regulation and mitigation strategies addressing the combined pollution of PFASs and humic substances in landfill leachate. Full article
(This article belongs to the Section Water Quality and Contamination)
Show Figures

Figure 1

Figure 1
<p>Chemical properties of humic substances revealed by SUVA254 (<b>a</b>), SUVA280 (<b>b</b>), A250/A365 (<b>c</b>), HIX (<b>d</b>), HIXg (<b>e</b>), BIX (<b>f</b>), TOC (<b>g</b>) and FI (<b>h</b>). The significance levels are as follows: *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Fluorescence shapes of two components (C1 and C2) derived from PARAFAC-modeling (<b>a</b>,<b>b</b>), significant difference analysis of C1 and C2 (<b>c</b>,<b>d</b>), and Fmax and relative contents of C1 and C2 (<b>e</b>,<b>f</b>). The significance levels are as follows: *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Characteristics of humic substances and PFASs: (<b>a</b>) the test group of HA and PFBS; (<b>b</b>) the test group of HA and PFBA; (<b>c</b>) the group of FA and PFBS; and (<b>d</b>) the test group of FA and PFBA. Sample information corresponding to numbers 1–22 is shown in <a href="#app1-water-16-03527" class="html-app">Table S2</a>.</p>
Full article ">Figure 4
<p>Correlation plot of each component after complexation of humic substances and PFASs ((<b>a</b>) for Circos, (<b>b</b>) for Mantel). The significance levels are as follows: **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4 Cont.
<p>Correlation plot of each component after complexation of humic substances and PFASs ((<b>a</b>) for Circos, (<b>b</b>) for Mantel). The significance levels are as follows: **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Redundancy analysis diagram of the chemical characteristics of humic substances and the complexation parameters of humic substances with PFASs.</p>
Full article ">Figure 6
<p>Quenching curves of humic substances interacting with PFBS (<b>a</b>), and with PFBA (<b>b</b>).</p>
Full article ">
16 pages, 11395 KiB  
Article
Hydrated Calcium Silicate Erosion in Sulfate Environments a Molecular Dynamics Simulation Study
by Mengjie You, Xiaosan Yin, Yuzhou Sun, Hairong Wu, Jimin Li and Xiangming Zhou
Materials 2024, 17(23), 6005; https://doi.org/10.3390/ma17236005 - 7 Dec 2024
Viewed by 592
Abstract
To investigate the micro-mechanism of the erosion of hydrated calcium silicate (C-S-H gel) in a sulfate environment, a solid–liquid molecular dynamics model of C-S-H gel/sodium sulfate was developed. This model employs molecular dynamics methods to simulate the transport processes between C-S-H gel and [...] Read more.
To investigate the micro-mechanism of the erosion of hydrated calcium silicate (C-S-H gel) in a sulfate environment, a solid–liquid molecular dynamics model of C-S-H gel/sodium sulfate was developed. This model employs molecular dynamics methods to simulate the transport processes between C-S-H gel and corrosive ions at concentrations of 5%, 8%, and 10% sodium sulfate (Na2SO4), aiming to elucidate the interaction mechanism between sulfate and C-S-H gel. The micro-morphology of the eroded samples was also investigated using scanning electron microscopy (SEM). The findings indicate that the adsorption capacity of C-S-H for ions significantly increases with higher concentrations of Na2SO4 solution. Notably, the presence of sulfate ions facilitates the decalcification reaction of C-S-H, leading to the formation of swollen gypsum and AFt (ettringite). This process results not only in the hydrolysis of the C-S-H gel but also in an increase in the diffusion coefficients of Na+ and Ca2+, thereby exacerbating the erosion. Additionally, the pore surfaces of the C-S-H structure exhibited strong adsorption of Na+, and as the concentration of Na2SO4 solution increased, Na+ was more stably adsorbed onto the C-S-H pore surfaces via Na-Os bonds. The root-mean-square displacement curves of water molecules were significantly higher than those of SO42, Na+ and Ca2+, which indicated that SO42 could co-penetrate and migrate with water molecules faster compared with other ions in the solution containing SO42, resulting in stronger corrosion and hydrolysis effects on the C-S-H structure. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Initial structure of Tobermorite and optimized crystal models.</p>
Full article ">Figure 2
<p>Convergence curve of the C-S-H model.</p>
Full article ">Figure 3
<p>Energy profile after dynamics optimization.</p>
Full article ">Figure 4
<p>C-S-H/Na<sub>2</sub>SO<sub>4</sub> solid–liquid modeling process.</p>
Full article ">Figure 5
<p>Solid–liquid ionic model.</p>
Full article ">Figure 6
<p>SEM images of concrete before erosion at different magnifications.</p>
Full article ">Figure 7
<p>Scanning electron microscope images of 5% sodium sulphate eroded concrete for 90 days at different magnifications.</p>
Full article ">Figure 8
<p>Scanning electron microscope images of concrete after 90 days of 8% sodium sulphate erosion at different magnifications.</p>
Full article ">Figure 9
<p>Calcium and Sulfate Ion Distribution at Initial (0 ps) and (50 ps) Time Points. (<b>a</b>) Concentration of 5%; (<b>b</b>) Concentration of 8%; (<b>c</b>) Concentration of 10%.</p>
Full article ">Figure 10
<p>Relative concentrations at the C-S-H/sodium sulfate interface. (<b>a</b>) Concentration of 5%; (<b>b</b>) Concentration of 8%; (<b>c</b>) Concentration of 10%.</p>
Full article ">Figure 11
<p>Distribution of sodium and silicon ions at initial (0 ps) and (50 ps) time points. (<b>a</b>) Concentration of 5%; (<b>b</b>) Concentration of 8%; (<b>c</b>) Concentration of 10%.</p>
Full article ">Figure 12
<p>Radial distribution function and average coordination number at different concentrations. (<b>a</b>) Concentration of 5%; (<b>b</b>) Concentration of 8%; (<b>c</b>) Concentration of 10%.</p>
Full article ">Figure 13
<p>Real-time screenshot at pore surface (green for Ca<sup>2+</sup>, yellow for sulphur atoms, red for O, purple for Na<sup>+</sup>). (<b>a</b>) Concentration of 5%; (<b>b</b>) Concentration of 8%; (<b>c</b>) Concentration of 10%.</p>
Full article ">Figure 14
<p>MSD at different concentrations. (<b>a</b>) Concentration of 5%; (<b>b</b>) Concentration of 8%; (<b>c</b>) Concentration of 10%.</p>
Full article ">
22 pages, 3355 KiB  
Article
Structural Characteristics and Adsorption of Phosphorus by Pineapple Leaf Biochar at Different Pyrolysis Temperatures
by Shuhui Song, Siru Liu, Yanan Liu, Weiqi Shi and Haiyang Ma
Agronomy 2024, 14(12), 2923; https://doi.org/10.3390/agronomy14122923 - 6 Dec 2024
Viewed by 343
Abstract
Biochar is a potential material for making slow-releasing phosphorus (P) fertilizers for the sake of increasing soil P-use efficiency. The adsorption of phosphorus by pineapple leaf biochar (PB) prepared at different pyrolysis temperatures and its mechanism remain unclear. In order to study the [...] Read more.
Biochar is a potential material for making slow-releasing phosphorus (P) fertilizers for the sake of increasing soil P-use efficiency. The adsorption of phosphorus by pineapple leaf biochar (PB) prepared at different pyrolysis temperatures and its mechanism remain unclear. In order to study the effect of preparation temperature on the structural characteristics of biochar from pineapple leaves and the adsorption of phosphorus by biochar, pineapple leaves were used as raw materials to prepare biochar by restricting oxygen supply at 300 °C, 500 °C, and 700 °C. The structural characteristics and adsorption of phosphorus by pineapple leaf biochar at different temperatures (PB300, PB500, and PB700) were analyzed. The results showed the following: (1) The pore structure of biochar pyrolysis at 300 °C (PB300) did not significantly change, while the surface structure of biochar pyrolysis at 700 °C (PB700) significantly changed, the specific surface area (SBET) increased by 26.91~37.10 times that observed in PB300 and PB500, and the pore wall became thinner. (2) The number of functional groups (C=O) in PB700 decreased, and the relative content of C-H/-CHO in PB500 and PB700 increased by 4.38 times that observed in PB300. (3) The adsorption of phosphorus by biochar was a multi-molecular layer chemisorption, accompanied by single-molecular-layer physical adsorption and intramolecular diffusion. For PB300, both the physical and chemical processes of the adsorption of PO43− by biochar were weakened, and the chemical process was dominated by cationic (Ca2+, Mg2+, Fe3+, and Al3+) adsorption at 500 °C. For PB700, the physical adsorption dominated by pore size structure was the main process, and the physicochemical adsorption at 700 °C was significantly stronger than that observed at 300 °C and 500 °C. These results indicate that biochar prepared at 500 °C can save energy in the preparation process and has excellent physical and chemical structure, which can be used as the basic material for further modification and preparation of biochar phosphate fertilizer. Full article
(This article belongs to the Section Soil and Plant Nutrition)
Show Figures

Figure 1

Figure 1
<p>Derivative thermogravimetric analysis curves of PAL biochar (PB).</p>
Full article ">Figure 2
<p>SEM images of surface pore size structure of biochar ((<b>a</b>) PB300; (<b>b</b>) PB500; (<b>c</b>) PB700).</p>
Full article ">Figure 3
<p>Nitrogen physisorb isotherm (<b>a</b>–<b>c</b>) and pore size distribution (<b>d</b>) of biochar.</p>
Full article ">Figure 4
<p>Adsorption capacity of phosphorus by biochar materials (<span class="html-italic">n</span> = 3, different lowercase letters indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05) among treatments.).</p>
Full article ">Figure 5
<p>Fitting curves of the kinetic model of adsorption phosphorus by three biochar types ((<b>a</b>) Pseudo-first-order, (<b>b</b>) Pseudo-second-order, (<b>c</b>) Intra-particle diffusion).</p>
Full article ">Figure 6
<p>Fitting curves of phosphorus adsorption isotherm models for three biochar types ((<b>a</b>) Freundlich, (<b>b</b>) Langmuir, (<b>c</b>) Temkin).</p>
Full article ">Figure 7
<p>Principal component analysis of the functional groups, elemental content, and surface pore structure.</p>
Full article ">Figure 8
<p>Heat map analysis of the correlation between the factors.</p>
Full article ">Figure 9
<p>Adsorption process and mechanism of phosphorus by PB prepared at different temperatures (Note: The main form of phosphorus in solution with pH = 3 is H<sub>2</sub>PO<sub>4</sub><sup>−</sup> in this study).</p>
Full article ">
23 pages, 8705 KiB  
Article
Multiscale Qualitative–Quantitative Characterization of the Pore Structure in Coal-Bearing Reservoirs of the Yan’an Formation in the Longdong Area, Ordos Basin
by Rong Wang, Baohong Shi, Tao Wang, Jiahao Lin, Bo Li, Sitong Fan and Jiahui Liu
Processes 2024, 12(12), 2787; https://doi.org/10.3390/pr12122787 - 6 Dec 2024
Viewed by 389
Abstract
Accurate characterization of coal reservoir micro- and nanopores is crucial in evaluating coalbed methane storage and gas production capacity. In this work, 12 coal-bearing rock samples from the Jurassic Yan’an Formation, Longdong area, Ordos Basin were taken as research objects, and micro- and [...] Read more.
Accurate characterization of coal reservoir micro- and nanopores is crucial in evaluating coalbed methane storage and gas production capacity. In this work, 12 coal-bearing rock samples from the Jurassic Yan’an Formation, Longdong area, Ordos Basin were taken as research objects, and micro- and nanopore structures were characterized via scanning electron microscopy, high-pressure mercury pressure, low-temperature N2 adsorption and low-pressure CO2 adsorption experiments. The main factors controlling coal pore structure development and the influence of pore development on the gas content were studied by combining the reflectivity of specular samples from the research area, the pore microscopic composition and the pore gas content determined through industrial analyses and isothermal absorption experiments. The results show that the coal strata of the Yan’an coal mine are a very important gas source, and that the coal strata of the Yan’an Formation in the study area exhibit remarkable organic and clay mineral pore development accompanied by clear microfractures and clay mineral interlayer joints, which together optimize the coal gas storage conditions and form efficient microseepage pathways for gas. Coalstone, carbonaceous mudstone and mudstone show differential distributions in pore volume and specific surface area. The general trend is that coal rock is the best, carbonaceous mudstone is the second best, and mudstone is the weakest. The coal samples’ microporous properties are positively correlated with the coal sample composition for the specular group, whereas there is no clear correlation for the inert group. An increase in the moisture content of the air-dried matrix promotes adsorption pore development, leading to increases in the microporous volume and specific surface area. CH4 adsorption in coal rock increases with increasing pressure, and the average maximum adsorption is approximately 8.13 m3/t. The limit of the amount of methane adsorbed by the coal samples, VL, is positively correlated with the pore volume and specific surface area, indicating that the larger the pore volume is, the greater the amount of gas that can be adsorbed by the coal samples, and the larger the specific surface area is, the greater the amount of methane that can be adsorbed by the coal samples. The PL value, pore volume and specific surface area are not correlated, indicating that there is no direct mathematical relationship between them. Full article
Show Figures

Figure 1

Figure 1
<p>Comprehensive histogram of the main Jurassic coal-bearing zones and strata in the Ordos Basin.</p>
Full article ">Figure 2
<p>Pore development characteristics of coal-bearing rock systems of the Yan’an Formation (<b>a</b>) J40-1, two cracks crossing, development of organic pores and minerals; (<b>b</b>) J40-1, development of organic pores, cracks filled with minerals; (<b>c</b>) J40-1, development of organic pores, some organic pores filled with pyrite and clay minerals; (<b>d</b>) J40-6, development of organic matter and a crack, with pyrite being visible; (<b>e</b>) J40-6, development of clay mineral pores; (<b>f</b>) J40-6, development of clay mineral pores, clay mineral interlayer joints and pyrite in clay minerals; (<b>g</b>) J40-3, development of organic pores, with pyrite being visible; (<b>h</b>) J40-3, development of a fissure, with organic pores being visible; (<b>i</b>) J40-3, development of clay mineral pores and clay mineral interlayer joints.</p>
Full article ">Figure 3
<p>Mercury intrusion and withdrawal curves for samples from coal-bearing rock systems.</p>
Full article ">Figure 4
<p>Pore volume and pore specific surface area distribution curves based on high-pressure mercury compression tests.</p>
Full article ">Figure 5
<p>Low-temperature N<sub>2</sub> adsorption–desorption isotherms.</p>
Full article ">Figure 6
<p>Pore volume and pore specific surface area distribution curves obtained via the low-temperature N<sub>2</sub> adsorption method.</p>
Full article ">Figure 7
<p>Low-pressure CO<sub>2</sub> adsorption isotherms.</p>
Full article ">Figure 8
<p>Pore volume and pore specific surface area distribution curves obtained based on low-pressure CO<sub>2</sub> adsorption data.</p>
Full article ">Figure 9
<p>Isothermal adsorption curves for the coal rock samples.</p>
Full article ">Figure 10
<p>Characterization of the multiscale pore volume and specific surface area distributions based on the three combined methods.</p>
Full article ">Figure 11
<p>Distributions of the pore volume and specific surface area as percentages of the full pore size.</p>
Full article ">Figure 12
<p>Relationships of the pore volume and specific surface area with the microcomponent composition at different scales in coal rock samples. (<b>a</b>) Relationship between the vitrinite content and pore volume; (<b>b</b>) Relationship between the vitrinite content and specific surface area; (<b>c</b>) Relationship between the inertinite content and pore volume; (<b>d</b>) Relationship between the inertinite content and specific surface area.</p>
Full article ">Figure 12 Cont.
<p>Relationships of the pore volume and specific surface area with the microcomponent composition at different scales in coal rock samples. (<b>a</b>) Relationship between the vitrinite content and pore volume; (<b>b</b>) Relationship between the vitrinite content and specific surface area; (<b>c</b>) Relationship between the inertinite content and pore volume; (<b>d</b>) Relationship between the inertinite content and specific surface area.</p>
Full article ">Figure 13
<p>Pore volume and specific surface area of coal rock samples at different scales in relation to the proximate analysis results. (<b>a</b>) Relationship between the <span class="html-italic">Mad</span> content and pore volume; (<b>b</b>) Relationship between the <span class="html-italic">Mad</span> content and specific surface area; (<b>c</b>) Relationship between the <span class="html-italic">Aad</span> content and pore volume; (<b>d</b>) Relationship between the <span class="html-italic">Aad</span> content and specific surface area.</p>
Full article ">Figure 13 Cont.
<p>Pore volume and specific surface area of coal rock samples at different scales in relation to the proximate analysis results. (<b>a</b>) Relationship between the <span class="html-italic">Mad</span> content and pore volume; (<b>b</b>) Relationship between the <span class="html-italic">Mad</span> content and specific surface area; (<b>c</b>) Relationship between the <span class="html-italic">Aad</span> content and pore volume; (<b>d</b>) Relationship between the <span class="html-italic">Aad</span> content and specific surface area.</p>
Full article ">Figure 14
<p>Relationships of the pore volume and specific surface area with the <span class="html-italic">V<sub>L</sub></span> and <span class="html-italic">P<sub>L</sub></span> of the coal rock samples. (<b>a</b>) Relationship between the pore volume and Langmuir volume <span class="html-italic">V<sub>L</sub></span>; (<b>b</b>) Relationship between the specific surface area and Langmuir volume <span class="html-italic">V<sub>L</sub></span>; (<b>c</b>) Relationship between the pore volume and Langmuir pressure <span class="html-italic">P<sub>L</sub></span>; (<b>d</b>) Relationship between the specific surface area and Langmuir pressure <span class="html-italic">P<sub>L</sub></span>.</p>
Full article ">Figure 14 Cont.
<p>Relationships of the pore volume and specific surface area with the <span class="html-italic">V<sub>L</sub></span> and <span class="html-italic">P<sub>L</sub></span> of the coal rock samples. (<b>a</b>) Relationship between the pore volume and Langmuir volume <span class="html-italic">V<sub>L</sub></span>; (<b>b</b>) Relationship between the specific surface area and Langmuir volume <span class="html-italic">V<sub>L</sub></span>; (<b>c</b>) Relationship between the pore volume and Langmuir pressure <span class="html-italic">P<sub>L</sub></span>; (<b>d</b>) Relationship between the specific surface area and Langmuir pressure <span class="html-italic">P<sub>L</sub></span>.</p>
Full article ">
Back to TopTop