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

Article Types

Countries / Regions

Search Results (66)

Search Parameters:
Keywords = cycloalkane

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 3522 KiB  
Article
Crude Oil Biodegradation by a Biosurfactant-Producing Bacterial Consortium in High-Salinity Soil
by Weiwei Chen, Jiawei Sun, Renping Ji, Jun Min, Luyao Wang, Jiawen Zhang, Hongjin Qiao and Shiwei Cheng
J. Mar. Sci. Eng. 2024, 12(11), 2033; https://doi.org/10.3390/jmse12112033 - 10 Nov 2024
Viewed by 584
Abstract
Bioremediation is a promising strategy to remove crude oil contaminants. However, limited studies explored the potential of bacterial consortia on crude oil biodegradation in high salinity soil. In this study, four halotolerant strains (Pseudoxanthomonas sp. S1-2, Bacillus sp. S2-A, Dietzia sp. CN-3, [...] Read more.
Bioremediation is a promising strategy to remove crude oil contaminants. However, limited studies explored the potential of bacterial consortia on crude oil biodegradation in high salinity soil. In this study, four halotolerant strains (Pseudoxanthomonas sp. S1-2, Bacillus sp. S2-A, Dietzia sp. CN-3, and Acinetobacter sp. HC8-3S), with strong environmental tolerance (temperature, pH, and salinity), distinctive crude oil degradation, and beneficial biosurfactant production, were combined to construct a bacterial consortium. The inoculation of the consortium successfully degraded 97.1% of total petroleum hydrocarbons in 10 days, with notable removal of alkanes, cycloalkanes, branched alkanes, and aromatic hydrocarbons. Functional optimization showed that this consortium degraded crude oil effectively in a broad range of temperature (20–37 °C), pH (6–9), and salinity (0–100 g/L). In salt-enriched crude-oil-contaminated soil microcosms, the simultaneous treatment of bioaugmentation and biostimulation achieved the highest crude oil degradation rate of 568.6 mg/kg/d, compared to treatments involving abiotic factors, natural attenuation, biostimulation, and bioaugmentation after 60 days. Real-time PCR targeting the 16S rRNA and alkB genes showed the good adaptability and stability of this consortium. The degradation property of the constructed bacterial consortium and the engineered consortium strategy may have potential use in the bioremediation of crude oil pollution in high-salinity soil. Full article
(This article belongs to the Section Marine Environmental Science)
Show Figures

Figure 1

Figure 1
<p>Evolutionary relationship of S1-2, S2-A, CN-3, and HC8-3S with other related strains. Based on the 16S rRNA gene sequences, the phylogenetic tree was constructed using the neighbor-joining algorithm method with 1000 bootstrap trials in MEGA 7.0 software.</p>
Full article ">Figure 2
<p>Emulsification activity (<b>a</b>) and cell surface hydrophobicity (<b>b</b>) of S1-2, S2-A, CN-3, and HC8-3S cultivated in MSM with different carbon sources (tetradecane, paraffin oil, and crude oil). The letters above the columns represent significant differences (<span class="html-italic">p</span> &lt; 0.05) among various groups using the one-way ANOVA test. Error bars represent the standard deviation of three independent measurements.</p>
Full article ">Figure 3
<p>Degradation efficacy of petroleum hydrocarbons with diverse carbon numbers (<b>a</b>) and compounds (alkanes, cycloalkanes, branched alkanes, and aromatic hydrocarbons) (<b>b</b>) in crude oil by <span class="html-italic">Pseudoxanthomonas</span> sp. S1-2, <span class="html-italic">Bacillus</span> sp. S2-A, <span class="html-italic">Dietzia</span> sp. CN-3, <span class="html-italic">Acinetobacter</span> sp. HC8-3S, and the consortium. The letters above the columns represent significant differences (<span class="html-italic">p</span> &lt; 0.05) among various groups through a one-way ANOVA test. Error bars represent the standard deviation of three independent measurements.</p>
Full article ">Figure 4
<p>Crude oil degradation by the consortium under diverse temperatures (<b>a</b>), pH values (<b>b</b>), and salinities (<b>c</b>) after incubation in shakers for 10 d. The MSM (pH 8 and NaCl 40 g/L) with crude oil was the control group. Error bars represent the standard deviation of three independent measurements.</p>
Full article ">Figure 5
<p>Crude oil depletion of AC, NA, BS, BA, and BS + BA treatments in non-saline soil microcosms (<b>a</b>) and S-AC, S-NA, S-BS, S-BA, and S-BS + BA treatments in salt-enriched soil microcosms (<b>b</b>) after 15, 30, 45, and 60 d. The letters above the columns (a, b, c, d, e) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) among various groups. Error bars represent the standard deviation of three independent measurements.</p>
Full article ">Figure 6
<p>The gene copies of 16S rRNA (<b>a</b>) and <span class="html-italic">alkB</span> (<b>b</b>) in different treatments after 0, 15, 30, 45, and 60 d. Error bars represent the standard deviation of three independent measurements.</p>
Full article ">
20 pages, 3193 KiB  
Review
Production of Sustainable Liquid Fuels
by Nathan Ormond, Dina Kamel, Sergio Lima and Basudeb Saha
Energies 2024, 17(14), 3506; https://doi.org/10.3390/en17143506 - 17 Jul 2024
Viewed by 899
Abstract
As the world aims to address the UN Sustainable Development Goals (SDGs), it is becoming more urgent for heavy transportation sectors, such as shipping and aviation, to decarbonise in an economically feasible way. This review paper investigates the potential fuels of the future [...] Read more.
As the world aims to address the UN Sustainable Development Goals (SDGs), it is becoming more urgent for heavy transportation sectors, such as shipping and aviation, to decarbonise in an economically feasible way. This review paper investigates the potential fuels of the future and their capability to mitigate the carbon footprint when other technologies fail to do so. This review looks at the technologies available today, including, primarily, transesterification, hydrocracking, and selective deoxygenation. It also investigates the potential of fish waste from the salmon industry as a fuel blend stock. From this, various kinetic models are investigated to find a suitable base for simulating the production and economics of biodiesel (i.e., fatty acid alkyl esters) and renewable diesel production from fish waste. Whilst most waste-oil-derived biofuels are traditionally produced using transesterification, hydrotreating looks to be a promising method to produce drop-in biofuels, which can be blended with conventional petroleum fuels without any volume percentage limitation. Using hydrotreatment, it is possible to produce renewable diesel in a few steps, and the final liquid product mixture includes paraffins, i.e., linear, branched, and cyclo-alkanes, with fuel properties in compliance with international fuel standards. There is a wide range of theoretical models based on the hydrodeoxygenation of fatty acids as well as a clear economic analysis that a model could be based on. Full article
(This article belongs to the Section A4: Bio-Energy)
Show Figures

Figure 1

Figure 1
<p>Global CO<sub>2</sub> emissions by sector and transportation subsector [<a href="#B1-energies-17-03506" class="html-bibr">1</a>].</p>
Full article ">Figure 2
<p>Transesterification reaction.</p>
Full article ">Figure 3
<p>Cyclic regeneration of hydroxide ion and triglyceride to diglyceride mechanism.</p>
Full article ">Figure 4
<p>Reversible reactions in the production of biodiesel.</p>
Full article ">Figure 5
<p>Pathways for producing sustainable liquid fuels from animal fats and vegetable oils.</p>
Full article ">Figure 6
<p>Production of biodiesel (FAMEs) through transesterification.</p>
Full article ">Figure 7
<p>Pathway of action for hydrocracking of triglycerides.</p>
Full article ">Figure 8
<p>Schematic for the Neste Oil NExBTL plant.</p>
Full article ">Figure 9
<p>Reactions involved in the deoxygenation of triglycerides.</p>
Full article ">Figure 10
<p>Reaction pathway for HDO, adapted from [<a href="#B32-energies-17-03506" class="html-bibr">32</a>].</p>
Full article ">Figure 11
<p>Reaction pathway for decarbonylation of stearate, adapted from [<a href="#B32-energies-17-03506" class="html-bibr">32</a>].</p>
Full article ">Figure 12
<p>Reaction scheme for HDO of stearic acid.</p>
Full article ">
19 pages, 2390 KiB  
Article
Examination of the Influence of Alternative Fuels on Particulate Matter Properties Emitted from a Non-Proprietary Combustor
by Liam D. Smith, Joseph Harper, Eliot Durand, Andrew Crayford, Mark Johnson, Hugh Coe and Paul I. Williams
Atmosphere 2024, 15(3), 308; https://doi.org/10.3390/atmos15030308 - 29 Feb 2024
Viewed by 1226
Abstract
The aviation sector, like most other sectors, is moving towards becoming net zero. In the medium to long term, this will mean an increase in the use of sustainable aviation fuels. Research exists on the impact of fuel composition on non-volatile particulate matter [...] Read more.
The aviation sector, like most other sectors, is moving towards becoming net zero. In the medium to long term, this will mean an increase in the use of sustainable aviation fuels. Research exists on the impact of fuel composition on non-volatile particulate matter (nvPM) emissions. However, there is more sparsity when considering the impact on volatile particulate matter (vPM) emissions. Here, nine different fuels were tested using an open-source design combustor rig. An aerosol mass spectrometer (AMS) was used to examine the mass-loading and composition of vPM, with a simple linear regression algorithm used to compare relative mass spectrum similarity. The diaromatic, cycloalkane and aromatic contents of the fuels were observed to correlate with the measured total number concentration and nvPM mass concentrations, resulting in an inverse correlation with increasing hydrogen content. The impacts of fuel properties on other physical properties within the combustion process and how they might impact the particulate matter (PM) are considered for future research. Unlike previous studies, fuel had a very limited impact on the organic aerosol’s composition at the combustor exit measurement location. Using a novel combination of Positive Matrix Factorization (PMF) and high-resolution AMS analysis, new insight has been provided into the organic composition. Both the alkane organic aerosol (AlkOA) and quenched organic aerosol (QOA) factors contained CnH2n+1, CnH2n−1 and CnH2n ion series, implying alkanes and alkenes in both, and approximately 12% oxygenated species in the QOA factor. These results highlight the emerging differences in the vPM compositional data observed between combustor rigs and full engines. Full article
(This article belongs to the Section Air Pollution Control)
Show Figures

Figure 1

Figure 1
<p>Hydrogen content versus alkane, mono-aromatic and cycloalkane content of the nine fuels examined.</p>
Full article ">Figure 2
<p>Schematic of experimental setup. Note: nvPM number instrumentation and catalytic stripper described in text are not shown here for clarity as the results from these are not presented.</p>
Full article ">Figure 3
<p>Comparison of different instrument measurements (<b>A</b>)—number concentrations versus EC mass concentration, (<b>B</b>)—number concentrations versus TC and EC mass concentrations, (<b>C</b>)—EC mass concentrations) from several different aerosol instruments. Error bars show one standard deviation. The 1:1 line is shown in all graphs.</p>
Full article ">Figure 4
<p>Particle number concentrations (<b>A</b>) as a function of fuel hydrogen content, averaged across a given fuel flow rate, whilst other variables were changing (for all other variables held constant, see <a href="#app1-atmosphere-15-00308" class="html-app">Figure S5</a>). In (<b>B</b>), relative differences in hydrogen content and number concentrations are presented, the latter of which is averaged across all conditions (T1–T3 and T6–T7; excluded test points are as such due to SMPS failure for J-LA’s TP8 and J-HA’s lack of TP4–5). Values in 4b are compared against those of J-REF. X-axis error bars in 4a show one standard deviation for fuel’s repeatability of their compositional analysis. Y error bars are not shown but are SMPS measurement uncertainty (±10%). Regression values are results of linear regression with the natural log of displayed y values.</p>
Full article ">Figure 5
<p>nvPM mass concentrations (<b>A</b>) as a function of fuel hydrogen content, averaged across a given fuel flow rate, whilst other variables are changing (for all other variables held constant, see <a href="#app1-atmosphere-15-00308" class="html-app">Figure S5</a>). In (<b>B</b>), relative differences of hydrogen content and nvPM mass concentrations are presented, the latter of which is averaged across all conditions (T1–T3 and T6–T7). Both values are compared against those of J-REF. X-axis error bars in 5a show one standard deviation for fuel’s repeatability of their compositional analysis. Regression values are results of linear regression with the natural log of displayed y values.</p>
Full article ">Figure 6
<p>(<b>a</b>): Boxplot of ordinary least squares linear regression performed on organic mass concentration produced at each <span class="html-italic">m/z</span> (up to <span class="html-italic">m/z</span> 150, excluding <span class="html-italic">m/z</span> 28) by a fuel at a given condition, compared against the equivalent <span class="html-italic">m/z</span> of the organic vPM emitted by another fuel at the same operating condition, above the threshold of 0.5 µg/m<sup>3</sup>. Boxes show interquartile ranges, with median line at centre. Error bars show top and bottom quartiles of the dataset. (<b>b</b>): The same as (<b>a</b>) except <span class="html-italic">m/z</span> 12, 18, 28 and 44 were removed from comparisons.</p>
Full article ">Figure 7
<p>PMF factors and their mass concentrations as a function of MSS EC mass concentration. AlkOA (<b>left</b>) and QOA (<b>right</b>). Shaded by the hydrogen content of the fuel in the given example.</p>
Full article ">Figure 8
<p>PMF factors, categorised by their AMS ions detected via high-resolution analysis.</p>
Full article ">
16 pages, 12440 KiB  
Article
Structural Characterization and Molecular Model Construction of Lignite: A Case of Xianfeng Coal
by Ying Shi, Yanming Zhu, Shangbin Chen, Yang Wang and Yu Song
Energies 2024, 17(5), 1049; https://doi.org/10.3390/en17051049 - 23 Feb 2024
Cited by 3 | Viewed by 1203
Abstract
The object of the study is lignite. Analytical testing techniques, such as elemental analysis, 13C nuclear magnetic resonance (13C NMR) spectroscopy, Fourier transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), and high-resolution transmission electron microscopy (HRTEM), were used to acquire [...] Read more.
The object of the study is lignite. Analytical testing techniques, such as elemental analysis, 13C nuclear magnetic resonance (13C NMR) spectroscopy, Fourier transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), and high-resolution transmission electron microscopy (HRTEM), were used to acquire information on the structural parameters of lignite. The aromaticity of Xianfeng lignite is 43.57%, and the aromatic carbon structure is mainly naphthalene and anthracene/phenanthrene. The aliphatic carbon structure is dominated by cycloalkanes, alkyl side chains, and hydrogenated aromatics. Oxygen is mainly present in ether oxygen, carboxyl, and carbonyl groups. Nitrogen is mainly in the form of pyrrole nitrogen and quaternary nitrogen. Sulfur is mainly thiophene sulfur. According to the analysis results, the molecular structure model of XF lignite was constructed. The molecular formula is C184H172O39N6S2. The 2D structure was converted to a 3D structure using computer simulation software and optimized. The optimized model has a remarkable stereoconfiguration, and the aromatic lamellae are irregularly arranged in space. The aromatic rings were mainly connected by methylene, hypomethylene, methoxy, and aliphatic rings. In addition, the simulated 13C NMR spectra are in good agreement with the experimental spectra. This shows the rationality of the 3D chemical structure model. Full article
(This article belongs to the Section H3: Fossil)
Show Figures

Figure 1

Figure 1
<p>Flow chart for 3D coal macromolecular structure modeling.</p>
Full article ">Figure 2
<p><sup>13</sup>C NMR peak fitting spectra of coal sample.</p>
Full article ">Figure 3
<p>FTIR spectrum fitting curve of coal samples. (<b>a</b>) 900–700 cm<sup>−1</sup>, (<b>b</b>) 1200–1000 cm<sup>−1</sup>, (<b>c</b>) 3000–2800 cm<sup>−1</sup>, (<b>d</b>) 3700–3400 cm<sup>−1</sup>.</p>
Full article ">Figure 4
<p>XPS wide-scan spectra of coal sample.</p>
Full article ">Figure 5
<p>Curve-fitted XPS C1s scan (<b>a</b>), O1s scan (<b>b</b>), N1s scan (<b>c</b>), and S2p scan (<b>d</b>) of coal sample.</p>
Full article ">Figure 6
<p>Peak fitting of Raman spectrum.</p>
Full article ">Figure 7
<p>Steps of HRTEM aromatic stripe image analysis of XF coal.</p>
Full article ">Figure 8
<p>Angle distribution of lattice fringes of coal samples.</p>
Full article ">Figure 9
<p>XRD spectra of coal sample.</p>
Full article ">Figure 10
<p>Partition-fitting XRD spectrum of coal sample.</p>
Full article ">Figure 11
<p>The 2D model of the XF coal molecular structure.</p>
Full article ">Figure 12
<p>Comparison of <sup>13</sup>C NMR spectra between experimental and model.</p>
Full article ">Figure 13
<p>The 3D model of the XF coal molecular structure. (Carbon atoms are gray, hydrogen atoms are white, oxygen atoms are red, nitrogen atoms are blue, and sulfur atoms are yellow).</p>
Full article ">
17 pages, 7624 KiB  
Article
Chromatographic Analysis of the Chemical Composition of Exhaust Gas Samples from Urban Two-Wheeled Vehicles
by Natalia Szymlet, Łukasz Rymaniak and Beata Kurc
Energies 2024, 17(3), 709; https://doi.org/10.3390/en17030709 - 1 Feb 2024
Cited by 1 | Viewed by 1093
Abstract
The subject of the article was the chemical analysis of gasoline and exhaust gas samples taken from an urban two-wheeled vehicle. The main aim of the work was to identify chemical compounds emitted by a group of urban two-wheeled vehicles depending on the [...] Read more.
The subject of the article was the chemical analysis of gasoline and exhaust gas samples taken from an urban two-wheeled vehicle. The main aim of the work was to identify chemical compounds emitted by a group of urban two-wheeled vehicles depending on the engine’s operating parameters. First, engine operating parameters and driving parameters of three urban two-wheeled vehicles were measured in real operating conditions. Based on the averaged results, engine operating points were determined for exhaust gas samples that were collected into Tedlar bags. The exhaust gas composition of individual chemical substances obtained in the chromatographic separation process were subjected to a detailed analysis relating the engine operating point with their emission rate, with each individual component being assessed in terms of its impact on human health. The obtained qualitative analysis results indicated the presence of alkenes, alkanes, aliphatic aldehydes, and aromatic and cyclic hydrocarbons (cycloalkanes) in the tested samples. The experiments provided a variety of conclusions relating to the operating parameters of a two-wheeler engine. Qualitative assessment of exhaust samples showed that a two-wheeled vehicle was characterized by the most varying composition of BTX aromatic hydrocarbons derivatives, which are particularly dangerous to human health and life. Therefore, the authors suggest that in the future, approval procedures regarding toxic emissions should be extended to include chromatographic tests. The presented results are an extension of previous studies on toxic emissions from urban two-wheeled vehicles in real operating conditions that were published in other journals. Full article
(This article belongs to the Special Issue Emission Control Technology in Internal Combustion Engines)
Show Figures

Figure 1

Figure 1
<p>Characteristics of the operating time density of three urban two-wheeled vehicles in real operating conditions.</p>
Full article ">Figure 2
<p>Measurement station for collecting exhaust gas samples.</p>
Full article ">Figure 3
<p>Scheme of sample collection on the Tenax surface.</p>
Full article ">Figure 4
<p>Schematic of the ATD–GC–MS measurement technique.</p>
Full article ">Figure 5
<p>View of the research equipment.</p>
Full article ">Figure 6
<p>Temperature program of the oven and column.</p>
Full article ">Figure 7
<p>Chromatograms of exhaust gas samples from an urban two-wheeled vehicle at various operating points of the combustion engine.</p>
Full article ">Figure 8
<p>Mass spectra of eluted substances for retention times (<b>a</b>) 1.66 min (<b>b</b>) 1.67 min.</p>
Full article ">
28 pages, 12063 KiB  
Article
Improving the Antioxidant Activity, Yield, and Hydrocarbon Content of Bio-Oil from the Pyrolysis of Açaí Seeds by Chemical Activation: Effect of Temperature and Molarity
by Flávio Pinheiro Valois, Kelly Christina Alves Bezerra, Fernanda Paula da Costa Assunção, Lucas Pinto Bernar, Simone Patrícia Aranha da Paz, Marcelo Costa Santos, Waldeci Paraguassu Feio, Renan Marcelo Pereira Silva, Neyson Martins Mendonça, Douglas Alberto Rocha de Castro, Sergio Duvoisin Jr., Antônio Rafael Quadros Gomes, Victor Ricardo Costa Sousa, Marta Chagas Monteiro and Nélio Teixeira Machado
Catalysts 2024, 14(1), 44; https://doi.org/10.3390/catal14010044 - 9 Jan 2024
Viewed by 2006
Abstract
Biomass-derived products are a promising way to substitute the necessity for petroleum-derived products, since lignocellulosic material is widely available in our atmosphere and contributes to the reduction of greenhouse gases (GHGs), due to zero net emissions of CO2. This study explores [...] Read more.
Biomass-derived products are a promising way to substitute the necessity for petroleum-derived products, since lignocellulosic material is widely available in our atmosphere and contributes to the reduction of greenhouse gases (GHGs), due to zero net emissions of CO2. This study explores the impact of temperature and molarity on the pyrolysis of açaí seeds (Euterpe oleracea, Mart.) activated with KOH and subsequently on the yield of bio-oil, hydrocarbon content of bio-oil, antioxidant activity of bio-oil, and chemical composition of the aqueous phase. The experiments were carried out at 350, 400, and 450 °C and 1.0 atmosphere, with 2.0 M KOH, and at 450 °C and 1.0 atmosphere, with 0.5 M, 1.0 M, and 2.0 M KOH, at laboratory scale. The composition of bio-oils and the aqueous phase were determined by GC-MS, while the acid value, a physicochemical property of fundamental importance in biofuels, was determined by AOCS methods. The antioxidant activity of bio-oils was determined by the TEAC method. The solid phase (biochar) was characterized by X-ray diffraction (XRD). The diffractograms identified the presence of Kalicinite (KHCO3) in biochar, and those higher temperatures favor the formation peaks of Kalicinite (KHCO3). The pyrolysis of açaí seeds activated with KOH show bio-oil yields from 3.19 to 6.79 (wt.%), aqueous phase yields between 20.34 and 25.57 (wt.%), solid phase yields (coke) between 33.40 and 43.37 (wt.%), and gas yields from 31.85 to 34.45 (wt.%). The yield of bio-oil shows a smooth exponential increase with temperature. The acidity of bio-oil varied between 12.3 and 257.6 mg KOH/g, decreasing exponentially with temperature, while that of the aqueous phase varied between 17.9 and 118.9 mg KOH/g, showing an exponential decay behavior with temperature and demonstrating that higher temperatures favor not only the yield of bio-oil but also bio-oils with lower acidity. For the experiments with KOH activation, the GC-MS of bio-oil identified the presence of hydrocarbons (alkanes, alkenes, cycloalkanes, cycloalkenes, and aromatics) and oxygenates (carboxylic acids, phenols, ketones, and esters). The concentration of hydrocarbons varied between 10.19 and 25.71 (area.%), increasing with temperature, while that of oxygenates varied between 52.69 and 72.15 (area.%), decreasing with temperature. For the experiments with constant temperature, the concentrations of hydrocarbons in bio-oil increased exponentially with molarity, while those of oxygenates decreased exponentially, showing that higher molarities favor the formation of hydrocarbons in bio-oil. The antioxidant activity of bio-oils decreases with increasing temperature, as the content of phenolic compounds decreases, and it decreases with increasing KOH molarity, as higher molarities favor the formation of hydrocarbons. Finally, it can be concluded that chemical activation of açaí seeds with KOH favors not only the yield of bio-oil but also the content of hydrocarbons. The study of process variables is of utmost importance in order to clearly assess reaction mechanisms, economic viability, and design goals that could be derived from chemically activated biomass pyrolysis processes. The study of the antioxidant properties of pyrolysis oils provides insight into new products derived from biomass pyrolysis. Full article
(This article belongs to the Section Biomass Catalysis)
Show Figures

Figure 1

Figure 1
<p>Yield of reaction products (bio-oil, H<sub>2</sub>O, biochar, gas) by pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart) in the temperature range of 350–450 °C.</p>
Full article ">Figure 2
<p>Concentration of acyclic saturated/unsaturated hydrocarbons (alkanes + alkenes) and heterocyclic hydrocarbons (cycloalkanes + cycloalkenes + aromatics) in bio-oil by pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart), in the temperature range of 350–450 °C.</p>
Full article ">Figure 3
<p>Concentration of oxygenates (phenols, ketones, and esters) in bio-oil by pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart), in the temperature range of 350–450 °C.</p>
Full article ">Figure 4
<p>Acidity of bio-oil obtained by pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart), in the temperature range of 350–450 °C.</p>
Full article ">Figure 5
<p>Acidity of aqueous phase obtained by pyrolysis of açaí seeds.</p>
Full article ">Figure 6
<p>Total antioxidant capacity in the bio-oil produced by the pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart) with 2 M KOH solution, in the temperature range of 350–450 °C.</p>
Full article ">Figure 7
<p>Yield of reaction products (bio-oil, H<sub>2</sub>O, biochar, gas) by pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart), at 450 °C, 1.0 atmosphere, and activated with 0.5 M, 1.0 M, and 2.0 M KOH, at laboratory scale.</p>
Full article ">Figure 8
<p>Concentrations of hydrocarbons and oxygenates in bio-oil obtained by pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart) at 450 °C, 1.0 atmosphere, and using different KOH concentrations (0.5–2.0 M).</p>
Full article ">Figure 9
<p>Acidity of bio-oil obtained by pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart) at 450 °C, 1.0 atmosphere, and using different KOH concentrations (0.5–2.0 M).</p>
Full article ">Figure 10
<p>Total antioxidant capacity in the bio-oil produced by the pyrolysis of açaí seeds (<span class="html-italic">Euterpe oleracea</span>, Mart) with different molarities (0.5–2 M KOH solution), at a temperature of 450 °C.</p>
Full article ">Figure 11
<p>XRD of biochar produced by pyrolysis of açaí seeds at 350 °C (<b>a</b>), 400 °C (<b>b</b>), and 450 °C (<b>c</b>), 1.0 atmosphere, and activated with 2.0 M KOH, at laboratory scale.</p>
Full article ">Figure 12
<p>Mechanism of lignin pyrolysis in the KOH-activated pyrolysis of açaí seeds.</p>
Full article ">Figure 13
<p>Process flow schema of bio-oil production by pyrolysis of açaí seeds at 350, 400, and 450 °C, 1.0 atm, 2.0 M KOH, and 450 °C, 1.0 atm, 0.5 M, 1.0 M, and 2.0 M KOH, using a fixed bed reactor, at laboratory scale.</p>
Full article ">Figure 14
<p>Biomass waste in the form of açaí seeds in Belém-Pará.</p>
Full article ">Figure 15
<p>Açaí seeds pre-treatment: dried açaí seeds (<b>a</b>); knife cutting mill (<b>b</b>); mechanical sieve shaker (<b>c</b>); dried, ground, and sieved açaí seeds (<b>d</b>).</p>
Full article ">Figure 16
<p>Chemical activation of dried, ground, and sieved açaí seeds with 2.0 M KOH solution: açaí seed fine powders mixed with 0.5 M, 1.0 M, and 2.0 M KOH solution (<b>a</b>); washing/filtration of açaí pasty cake (<b>b</b>); KOH-activated açaí fine powders seeds (<b>c</b>).</p>
Full article ">Figure 17
<p>Schematic diagram of a laboratory-scale borosilicate glass reactor.</p>
Full article ">Figure 18
<p>Laboratory-scale pyrolysis reactor.</p>
Full article ">
20 pages, 7383 KiB  
Article
Pyrolysis of Cyclohexane and 1-Hexene at High Temperatures and Pressures—A Photoionization Mass Spectrometry Study
by Robert S. Tranter, Colin Banyon, Ryan E. Hawtof and Keunsoo Kim
Energies 2023, 16(24), 7929; https://doi.org/10.3390/en16247929 - 6 Dec 2023
Viewed by 1178
Abstract
Cycloalkanes are important components of a wide range of fuels. However, there are few experimental data at simultaneously high temperatures and pressures similar to those found in practical systems. Such data are necessary for developing and testing chemical kinetic models. In this study, [...] Read more.
Cycloalkanes are important components of a wide range of fuels. However, there are few experimental data at simultaneously high temperatures and pressures similar to those found in practical systems. Such data are necessary for developing and testing chemical kinetic models. In this study, data relevant to cycloalkane pyrolysis were obtained from high repetition rate shock tube experiments coupled with synchrotron-based photoionization mass spectrometry diagnostics. The pyrolysis of cyclohexane was studied over 1270–1550 K and ~9 bar, while the more reactive primary decomposition product, 1-hexene, was studied at 1160–1470 K and ~5 bar. Insights into the decomposition of the parent molecules, the formation of primary products and the production of aromatic species were gained. Simulations were performed with models for cyclohexane and 1-hexene that were based on literature models. The results indicate that over several hundred microseconds reaction time at high pressures and temperatures the pyrolysis of cyclohexane is largely dominated by reactions initiated by cyclohexyl radicals. Furthermore, good agreement between the simulations and the experiments were observed for cyclohexane and 1-hexene with a modified version of the cyclohexane model. Conversely, the 1-hexene model did not reproduce the experimental observations. Full article
(This article belongs to the Special Issue Advances in Fuels and Combustion)
Show Figures

Figure 1

Figure 1
<p>Mass spectra from pyrolysis of cyclohexane; 1400 K, 8 bar, 10.75 eV, no MgF<sub>2</sub> filter.</p>
Full article ">Figure 2
<p>Mass spectra from pyrolysis of cyclohexane; 300 µs reaction time, 10.75 eV, no MgF<sub>2</sub> filter.</p>
Full article ">Figure 3
<p>Effective concentrations of 1,3-butadiene from the pyrolysis of cyclohexane at various temperatures, ~8 bar, 10.75 eV, no MgF<sub>2</sub>. t = 0 corresponds to formation of the reflected shock wave.</p>
Full article ">Figure 4
<p>Effective concentrations of key species from the dissociation of cyclohexane at 8 bar. No MgF<sub>2</sub> filter. m/z 40 = allene + propyne. m/z 41 = allyl + <sup>13</sup>C from m/z 40. t = 0 corresponds to formation of the reflected shock wave.</p>
Full article ">Figure 5
<p>Mass-selected photoionization spectra corresponding to <a href="#energies-16-07929-f004" class="html-fig">Figure 4</a>. Vertical lines indicate approximate locations of ionization thresholds (see text for exact values).</p>
Full article ">Figure 6
<p>Effective concentration of the allyl radical produced from dissociation of cyclohexane. PE = 10.0 eV, with the MgF<sub>2</sub> filter. m/z 41 is corrected for the <sup>13</sup>C isotopologue of m/z 40. The red line is a best fit to the m/z 41 signal as a visual aid. t = 0 corresponds to formation of the reflected shock wave.</p>
Full article ">Figure 7
<p>Photoionization spectra for m/z 84 from the dissociation of cyclohexane at two temperatures.</p>
Full article ">Figure 8
<p>Mass spectra from the dissociation of 1-hexene; 1160 K, 4.4 bar, 10.70 eV.</p>
Full article ">Figure 9
<p>Mass spectra from the dissociation of 1-hexene at various temperatures and 4.4–6.4 bar; 300 µs reaction time, 10.70 eV.</p>
Full article ">Figure 10
<p>Effective concentrations of 1-hexene and the main products; 1160 K and 4.4 bar. t = 0 corresponds to formation of the reflected shock wave.</p>
Full article ">Figure 11
<p>Effective concentration of the allyl radical produced from dissociation of 1-hexene; 1160 K and 4.4 bar, PE = 10.0 eV. m/z 41 is corrected for the <sup>13</sup>C isotopologue of m/z 40. The red line is a best fit to the m/z 41 signal as a visual aid. t = 0 corresponds to formation of the reflected shock wave.</p>
Full article ">Figure 12
<p>Effective concentration plots at 10.0 eV for several main products from the dissociation of 1-hexene; 1160 K and 4.4 bar. t = 0 corresponds to formation of the reflected shock wave.</p>
Full article ">Figure 13
<p>Simulation results for pyrolysis of 1-hexene with the MCHb model; 1160 K and 5 bar.</p>
Full article ">Figure 14
<p>Simulation results for pyrolysis of 1-hexene with the Fan model; 1160 K and 5bar.</p>
Full article ">Figure 15
<p>Simulation results with the MCH model for the pyrolysis of cyclohexane; 1400 K and 10 bar.</p>
Full article ">Figure 16
<p>Photoionization spectra from dissociation of cyclohexane; 1400 K, 8 bar. The species are involved in the formation of benzene by dehydrogenation of cyclohexane.</p>
Full article ">Figure 17
<p>Species involved in the conversion of cyclohexane to benzene. T = 1270 K, P = 7 bar, no MgF<sub>2</sub> filter.</p>
Full article ">
13 pages, 9296 KiB  
Article
Mesoporous-Layered Double Oxide/MCM-41 Composite with Enhanced Catalytic Performance for Cyclopentanone Aldol Condensation
by Jinfan Yang, Ning Shang, Jiachen Wang and Huimin Liu
Molecules 2023, 28(23), 7920; https://doi.org/10.3390/molecules28237920 - 3 Dec 2023
Viewed by 1405
Abstract
Layered double oxides are widely employed in catalyzing the aldol condensation for producing biofuels, but its selectivity and stability need to be further improved. Herein, a novel MCM-41-supported Mg–Al-layered double oxide (LDO/MCM-41) was prepared via the in situ integration of a sol–gel process [...] Read more.
Layered double oxides are widely employed in catalyzing the aldol condensation for producing biofuels, but its selectivity and stability need to be further improved. Herein, a novel MCM-41-supported Mg–Al-layered double oxide (LDO/MCM-41) was prepared via the in situ integration of a sol–gel process and coprecipitation, followed by calcination. This composite was first employed to catalyze the self-condensation of cyclopentanone for producing high-density cycloalkane precursors. LDO/MCM-41 possessed large specific surface area, uniform pore size distribution, abundant medium basic sites and Bronsted acid sites. Compared with the bulk LDO, LDO/MCM-41 exhibited a higher selectivity for C10 and C15 oxygenates at 150 °C (93.4% vs. 84.6%). The selectivity for C15 was especially enhanced on LDO/MCM-41, which was three times greater than that on LDO. The stability test showed that naked LDO with stronger basic strength had a rapid initial activity, while it suffered an obvious deactivation due to its poor carbon balance. LDO/MCM-41 with lower basic strength had an enhanced stability even with a lower initial activity. Under the optimum conditions (50% LDO loading, 170 °C, 7 h), the cyclopentanone conversion on LDO/MCM-41 reached 77.8%, with a 60% yield of C10 and 15.2% yield of C15. Full article
(This article belongs to the Special Issue Porous Materials as Catalysts and Sorbents)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) XRD patterns of samples before calcination. (<b>b</b>) XRD patterns of LDO and LDO/MCM-41. (<b>c</b>) Low-angle XRD patterns of MCM-41 and LDO/MCM-41.</p>
Full article ">Figure 2
<p>(<b>a</b>) Nitrogen adsorption−desorption isotherms and (<b>b</b>) pore size distribution curve of different samples.</p>
Full article ">Figure 3
<p>SEM images of (<b>a</b>) MCM-41, (<b>b</b>) LDO, (<b>c</b>,<b>d</b>) LDO/MCM-41, and (<b>e</b>) SEM–EDS analysis of the specified red box area.</p>
Full article ">Figure 4
<p>(<b>a</b>) CO<sub>2</sub>-TPD, (<b>b</b>) NH<sub>3</sub>-TPD, (<b>c</b>) Py-FTIR of different samples.</p>
Full article ">Figure 5
<p>(<b>a</b>) Comparison of conversion (white bar), yield of C10 yield (grey bar), C15 (black bar), and oligomer (patterned bar) on LDO and LDO/MCM-41. (<b>b</b>) Selectivity of C10 (red bar), C15 (blue bar), and oligomer (purple bar). Reaction conditions: 0.2 g catalyst, 2.0 g CPO, 160 °C, 5 h. (<b>c</b>) The reaction route of CPO self- condensation.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) Effect of reaction temperature. Reaction conditions: 5 h, 1.0 g CPO, 0.1 g LDO/MCM-41. (<b>c</b>,<b>d</b>) Effect of reaction time. Reaction conditions: 170 °C, 1.0 g CPO, 0.1 g LDO/MCM-41. (<b>e</b>,<b>f</b>) Effect of LDO loading. Reaction conditions: 170 °C, 7 h, 1.0 g CPO, 0.1 g LDO/MCM-41. Yield of C10 (grey bar) and C15 (black bar), selectivity of C10 (red bar) and C15 (blue bar).</p>
Full article ">
16 pages, 8890 KiB  
Article
Study on the Aging Behavior of Asphalt Binder Exposed to Different Environmental Factors
by Shanglin Song, Linbing Wang, Chunping Fu, Meng Guo, Xiaoqiang Jiang, Meichen Liang and Luchun Yan
Appl. Sci. 2023, 13(23), 12651; https://doi.org/10.3390/app132312651 - 24 Nov 2023
Cited by 2 | Viewed by 1218
Abstract
Accelerated aging methods commonly used in laboratories struggle to replicate the outdoor aging process of asphalt binder. The aim of this study is to elucidate the impact of different environmental factors on the aging of asphalt binder and recreate the exposure process of [...] Read more.
Accelerated aging methods commonly used in laboratories struggle to replicate the outdoor aging process of asphalt binder. The aim of this study is to elucidate the impact of different environmental factors on the aging of asphalt binder and recreate the exposure process of asphalt binder. To achieve the study’s objectives, the asphalt binder was subjected to various environmental conditions through different aging modes. Three exposure modes (all environmental factors, the effects of light, temperature, oxygen, the effects of temperature, oxygen, and others) were established to assess the impact of various environmental factors on asphalt binder aging behavior. This mode was labeled O+UV-aging, earning it the name O-aging. The aging behaviors were assessed across multiple dimensions, considering apparent morphology, rheological properties, and chemical composition. The study’s findings highlight that factors such as ultraviolet radiation are primarily responsible for the formation of micro-cracks and increased surface roughness in aged asphalt binder. Ultraviolet radiation significantly affected the aging of asphalt binder during outdoor exposure. SBS modifiers increased the risk of fatigue cracking in the virgin asphalt binder but enhanced its aging resistance. After All-aging, the G-R parameter increase of virgin asphalt binder was 2.6 times that of SBS-modified asphalt binder. Throughout the exposure process, the broken molecular chains and the original molecular chains in the asphalt binder underwent polymerization reactions, resulting in longer carbon chains and cycloalkane aromatization. It was discovered that exposure showed less effect on the characteristic functional groups of SBS-modified binder than on virgin binder. After All-aging, the carbonyl index of SBS-modified asphalt binder was 56.4% higher than that of virgin asphalt binder. Full article
Show Figures

Figure 1

Figure 1
<p>Outdoor exposure test.</p>
Full article ">Figure 2
<p>Outdoor exposure test modes. (<b>a</b>) Mode 1; (<b>b</b>) Mode 2; (<b>c</b>) Mode 3.</p>
Full article ">Figure 3
<p>The apparent morphology of the O-aged samples. (<b>a</b>) Virgin asphalt binder; (<b>b</b>) SBS modified asphalt binder.</p>
Full article ">Figure 4
<p>The apparent morphology of the All-aged samples. (<b>a</b>) Virgin asphalt binder; (<b>b</b>) SBS modified asphalt binder.</p>
Full article ">Figure 5
<p>The apparent morphology of the O+UV-aged samples. (<b>a</b>) Virgin asphalt binder; (<b>b</b>) SBS modified asphalt binder.</p>
Full article ">Figure 6
<p>The apparent morphology of the O-aged samples. (<b>a</b>) Virgin asphalt binder; (<b>b</b>) SBS modified asphalt binder.</p>
Full article ">Figure 7
<p>Complex modulus master curve of virgin asphalt binder.</p>
Full article ">Figure 8
<p>Complex modulus master curve of SBS-modified asphalt binder.</p>
Full article ">Figure 9
<p>G-R parameter of asphalt binder.</p>
Full article ">Figure 10
<p>Positions of different hydrogen in nuclear magnetic resonance hydrogen spectra.</p>
Full article ">Figure 11
<p>The <sup>1</sup>H-NMR spectra of virgin asphalt binder. (<b>a</b>) Unaged; (<b>b</b>) TFOT aging; (<b>c</b>) O-aging; (<b>d</b>) O+UV-aging, (<b>e</b>) All-aging; (<b>f</b>) PAV-aging.</p>
Full article ">Figure 12
<p>The <sup>1</sup>H-NMR spectra of SBS-modified asphalt binder. (<b>a</b>) Unaged; (<b>b</b>) TFOT aging; (<b>c</b>) O-aging; (<b>d</b>) O+UV-aging; (<b>e</b>) All-aging; (<b>f</b>) PAV-aging.</p>
Full article ">Figure 12 Cont.
<p>The <sup>1</sup>H-NMR spectra of SBS-modified asphalt binder. (<b>a</b>) Unaged; (<b>b</b>) TFOT aging; (<b>c</b>) O-aging; (<b>d</b>) O+UV-aging; (<b>e</b>) All-aging; (<b>f</b>) PAV-aging.</p>
Full article ">Figure 13
<p>FTIR of virgin asphalt binder.</p>
Full article ">Figure 14
<p>FTIR of SBS modified asphalt binder.</p>
Full article ">
13 pages, 8449 KiB  
Article
Origin of Optoelectronic Contradictions in 3,4-Cycloalkyl[c]-chalcogenophenes: A Computational Study
by Ganesh Masilamani, Gamidi Rama Krishna, Sashi Debnath and Anjan Bedi
Polymers 2023, 15(21), 4240; https://doi.org/10.3390/polym15214240 - 27 Oct 2023
Cited by 2 | Viewed by 1170
Abstract
The planar morphology of the backbone significantly contributes to the subtle optoelectronic features of π-conjugated polymers. On the other hand, the atomistic tuning of an otherwise identical π-backbone could also impact optoelectronic properties systematically. In this manuscript, we compare a series of 3,4-cycloalkylchalcogenophenes [...] Read more.
The planar morphology of the backbone significantly contributes to the subtle optoelectronic features of π-conjugated polymers. On the other hand, the atomistic tuning of an otherwise identical π-backbone could also impact optoelectronic properties systematically. In this manuscript, we compare a series of 3,4-cycloalkylchalcogenophenes by tuning them atomistically using group-16 elements. Additionally, the effect of systematically extending these building blocks in the form of oligomers and polymers is studied. The size of the 3,4-substitution affected the morphology of the oligomers. In addition, the heteroatoms contributed to a further alteration in their geometry and resultant optoelectronic properties. The chalcogenophenes, containing smaller 3,4-cycloalkanes, resulted in lower bandgap oligomers or polymers compared to those with larger 3,4-cycloalkanes. Natural bonding orbital (NBO) calculations were performed to understand the disparity alongside the contour maps of frontier molecular orbitals (FMO). Full article
(This article belongs to the Special Issue Computational and Experimental Approaches in Polymeric Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of <b>CPX</b>, <b>PCPX</b>, <b>CHX</b> and <b>PCHX</b>.</p>
Full article ">Figure 2
<p>(<b>a</b>) Solid-state structure (<b>b</b>) Packing and (<b>c</b>) Optimized geometry of <b>(5,5-bis(methoxymethyl)-5,6-dihydro-4H-cyclopenta[<span class="html-italic">c</span>]thiophene-1,3-diyl)bis(trimethylsilane)</b> (CCDC 2249392). Hydrogen atoms were removed from subfigure (<b>a</b>,<b>c</b>) for clarity. The blue trace in subfigure (<b>b</b>) is the van der Waals force (O⋅⋅⋅H = 2.08 Å). Please see <a href="#app1-polymers-15-04240" class="html-app">supporting information</a> (<a href="#app1-polymers-15-04240" class="html-app">Figures S1 and S2</a>) for a detailed structure. The ORTEP plot of the molecule was drawn at a 50% probability level.</p>
Full article ">Figure 3
<p>Computationally (TDDFT-B3LYP-SDD) obtained gas-phase UV–vis spectra of (<b>a</b>) <b><span class="html-italic">n</span>CHO</b> and (<b>b</b>) <b><span class="html-italic">n</span>CPO</b>. Transitions at a shorter wavelength were omitted for data clarity.</p>
Full article ">Figure 4
<p>Computationally (TDDFT-B3LYP-SDD) obtained gas-phase UV–vis spectra of (<b>a</b>) <b><span class="html-italic">n</span>CHS</b> and (<b>b</b>) <b><span class="html-italic">n</span>CPS</b>.</p>
Full article ">Figure 5
<p>Computationally (TDDFT-B3LYP-SDD) obtained gas-phase UV–vis spectra of (<b>a</b>) <b><span class="html-italic">n</span>CHSe</b> and (<b>b</b>) <b><span class="html-italic">n</span>CPSe</b>. Transitions at a shorter wavelength were omitted only for selenophene and <b>CHSe</b> for data clarity.</p>
Full article ">Figure 6
<p>Computationally (TDDFT-B3LYP-SDD) obtained gas-phase UV–vis spectra of (<b>a</b>) <b><span class="html-italic">n</span>CHTe</b> and (<b>b</b>) <b><span class="html-italic">n</span>CPTe</b>. Transitions at a shorter wavelength were omitted only for tellurophene and <b>CHTe</b> for data clarity.</p>
Full article ">Figure 7
<p>Variation in λ<sub>max</sub> of (<b>a</b>) oligofuran, <b><span class="html-italic">n</span>CPO</b> and <b><span class="html-italic">n</span>CHO</b>, (<b>b</b>) oligothiophene, <b><span class="html-italic">n</span>CPS</b> and <b><span class="html-italic">n</span>CHS</b>, (<b>c</b>) oligoselenophene, <b><span class="html-italic">n</span>CPSe</b> and <b><span class="html-italic">n</span>CHSe</b>, and (<b>d</b>) oligotellurophene, <b><span class="html-italic">n</span>CPTe</b> and <b><span class="html-italic">n</span>CHTe</b> with the increase oligomers’ length. The black trace is parent chalcogenophenes (furan, thiophene, selenophene and tellurophene), the blue trace is <b><span class="html-italic">n</span>CPX</b>s, and red trace is <b><span class="html-italic">n</span>CHX</b>s.</p>
Full article ">Figure 8
<p>(<b>a</b>) Highest occupied molecular orbital (HOMO) and (<b>b</b>) lowest unoccupied molecular orbital (LUMO) contour maps of <b>2CPS</b>. (<b>c</b>) HOMO and (<b>d</b>) LUMO contour maps of <b>2CHS</b>. The isovalue of 0.04 was considered while generating FMOs. Optimized geometry of (<b>e</b>) <b>2CPS</b> and (<b>f</b>) <b>2CHS</b>. Hydrogen atoms were omitted for clarity.</p>
Full article ">Figure 9
<p>Computationally (DFT-B3LYP-SDD) obtained HOMO–LUMO gap of (<b>a</b>) <b><span class="html-italic">n</span>CHX</b>s and (<b>b</b>) <b><span class="html-italic">n</span>CPX</b>s.</p>
Full article ">
27 pages, 13065 KiB  
Article
Efficient and Selective Oxygenation of Cycloalkanes and Alkyl Aromatics with Oxygen through Synergistic Catalysis of Bimetallic Active Centers in Two-Dimensional Metal-Organic Frameworks Based on Metalloporphyrins
by Xin-Yan Zhou, Bo Fu, Wen-Dong Jin, Xiong Wang, Ke-Ke Wang, Mei Wang, Yuan-Bin She and Hai-Min Shen
Biomimetics 2023, 8(3), 325; https://doi.org/10.3390/biomimetics8030325 - 21 Jul 2023
Cited by 1 | Viewed by 1560
Abstract
Confined catalytic realms and synergistic catalysis sites were constructed using bimetallic active centers in two-dimensional metal-organic frameworks (MOFs) to achieve highly selective oxygenation of cycloalkanes and alkyl aromatics with oxygen towards partly oxygenated products. Every necessary characterization was carried out for all the [...] Read more.
Confined catalytic realms and synergistic catalysis sites were constructed using bimetallic active centers in two-dimensional metal-organic frameworks (MOFs) to achieve highly selective oxygenation of cycloalkanes and alkyl aromatics with oxygen towards partly oxygenated products. Every necessary characterization was carried out for all the two-dimensional MOFs. The selective oxygenation of cycloalkanes and alkyl aromatics with oxygen was accomplished with exceptional catalytic performance using two-dimensional MOF Co-TCPPNi as a catalyst. Employing Co-TCPPNi as a catalyst, both the conversion and selectivity were improved for all the hydrocarbons investigated. Less disordered autoxidation at mild conditions, inhibited free-radical diffusion by confined catalytic realms, and synergistic C–H bond oxygenation catalyzed by second metal center Ni employing oxygenation intermediate R–OOH as oxidant were the factors for the satisfying result of Co-TCPPNi as a catalyst. When homogeneous metalloporphyrin T(4-COOCH3)PPCo was replaced by Co-TCPPNi, the conversion in cyclohexane oxygenation was enhanced from 4.4% to 5.6%, and the selectivity of partly oxygenated products increased from 85.4% to 92.9%. The synergistic catalytic mechanisms were studied using EPR research, and a catalysis model was obtained for the oxygenation of C–H bonds with O2. This research offered a novel and essential reference for both the efficient and selective oxygenation of C–H bonds and other key chemical reactions involving free radicals. Full article
(This article belongs to the Section Biomimetics of Materials and Structures)
Show Figures

Figure 1

Figure 1
<p>FT-IR absorption spectra of prepared (<b>a</b>) Co-TCPPFe, Co-TCPPMn, Co-TCPPNi, TCPPCo and (<b>b</b>) Mn-TCPPNi, Ni-TCPPCo, Mn-TCPPCo, TCPPNi.</p>
Full article ">Figure 2
<p>(<b>a</b>) XPS survey spectra and (<b>b</b>,<b>c</b>) high-resolution XPS spectra of Co-TCPPNi and (<b>d</b>,<b>e</b>) Co-TCPPMn.</p>
Full article ">Figure 3
<p>XRD patterns of prepared two-dimensional MOFs (<b>a</b>) Co-TCPPFe, Co-TCPPMn, Co-TCPPNi and (<b>b</b>) Mn-TCPPNi, Ni-TCPPCo, Mn-TCPPCo.</p>
Full article ">Figure 4
<p>SEM pictures of the (<b>a</b>,<b>d</b>) Co-TCPPNi, (<b>b</b>,<b>e</b>) Co-TCPPFe and (<b>c</b>,<b>f</b>) Co-TCPPMn.</p>
Full article ">Figure 5
<p>TEM pictures of (<b>a</b>,<b>b</b>) Co-TCPPNi, (<b>d</b>,<b>e</b>) Co-TCPPFe, (<b>g</b>,<b>h</b>) Co-TCPPMn, and (<b>c</b>,<b>f</b>,<b>i</b>) EDS elemental mapping pictures.</p>
Full article ">Figure 6
<p>(<b>a</b>) Nitrogen adsorption–desorption isotherms and (<b>b</b>) pore width distributions of representative two-dimensional MOF Co-TCPPNi.</p>
Full article ">Figure 7
<p>Thermogravimetric curves of prepared two-dimensional MOFs in an air environment.</p>
Full article ">Figure 8
<p>Synergistic cycloalkane oxygenation catalyzed by Co-TCPPNi with oxygen.</p>
Full article ">Figure 9
<p>Pseudo-second-order fits for cyclohexane oxygenation with oxygen. Autoxidation (<b>a</b>,<b>b</b>), and in the presence of Co-TCPPNi (<b>c</b>,<b>d</b>), Co-TCPPFe (<b>e</b>,<b>f</b>), and Co-TCPPMn (<b>g</b>,<b>h</b>).</p>
Full article ">Figure 9 Cont.
<p>Pseudo-second-order fits for cyclohexane oxygenation with oxygen. Autoxidation (<b>a</b>,<b>b</b>), and in the presence of Co-TCPPNi (<b>c</b>,<b>d</b>), Co-TCPPFe (<b>e</b>,<b>f</b>), and Co-TCPPMn (<b>g</b>,<b>h</b>).</p>
Full article ">Figure 10
<p>EPR spectra of DMPO spinning adducts in simulation and experimental, and DMPO spinning adducts.</p>
Full article ">Scheme 1
<p>Schematic preparation of two-dimensional MOFs with double metal active sites.</p>
Full article ">Scheme 2
<p>Proposed mechanism for the C–H bond oxygenation with oxygen catalyzed by Co-TCPPNi employing cyclohexane as the model substrate.</p>
Full article ">
11 pages, 1542 KiB  
Article
Ketones in Low-Temperature Oxidation Products of Crude Oil
by Shuai Ma, Yunyun Li, Rigu Su, Jianxun Wu, Lingyuan Xie, Junshi Tang, Xusheng Wang, Jingjun Pan, Yuanfeng Wang, Quan Shi, Guangzhi Liao and Chunming Xu
Processes 2023, 11(6), 1664; https://doi.org/10.3390/pr11061664 - 30 May 2023
Cited by 4 | Viewed by 2120
Abstract
Ketone compounds are oxidation products of crude oil in the in-situ combustion (ISC) process. Revealing the molecular composition of ketones can provide theoretical guidance for understanding the oxidation process of crude oil and valuable clues for studying the combustion state of crude oil [...] Read more.
Ketone compounds are oxidation products of crude oil in the in-situ combustion (ISC) process. Revealing the molecular composition of ketones can provide theoretical guidance for understanding the oxidation process of crude oil and valuable clues for studying the combustion state of crude oil in the reservoir. In this study, low-temperature oxidation (LTO) processes were simulated in thermal oxidation experiments to obtain thermally oxidized oil at different temperatures (170 °C, 220 °C, 270 °C, and 320 °C). A combination of chemical derivatization and positive-ion electrospray (ESI) Fourier transform ion cyclotron resonance mass spectrometry (FT-ICR MS) was used to analyze the molecular composition of different kinds of ketones (fatty ketones, naphthenic ketones, and aromatic ketones) in the oxidized oils at different temperatures. The results showed that the concentration of aliphatic ketones and aliphatic cyclic ketones in the product oils decreased with the increase in temperature, while aromatic ketones increased with the increase in temperature. At the same oxidation temperature, the content of ketones follows this order: fatty ketones < cycloalkanes < aromatic ketones. The concentrations of ketones reached their maximum value at 170 °C and decreased at high temperatures due to over-oxidation. It was also found that nitrogen-containing compounds are more easily oxidized to ketone compounds than their hydrocarbon counterparts in the LTO process. Full article
Show Figures

Figure 1

Figure 1
<p>Sketch map of the separation of ketones and reaction schemes of ketones with Girard-T reagent.</p>
Full article ">Figure 2
<p>+ESI FT-ICR MS mass spectra (<b>left</b>) and compound-type abundance histograms (<b>right</b>) of the Gerard-T reagent derivatization products of crude oil and its oxidation products at 170 °C, 220 °C, 270 °C, and 320 °C. The legend in various colors corresponds to different DBE values, which are shown more clearly in <a href="#processes-11-01664-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 3
<p>Ion intensities of N3O1 and N4O1 class species in the mass spectra. These compound classes correspond to ketones and nitrogen-containing ketones in oils.</p>
Full article ">Figure 4
<p>Ion relative abundance plots of DBE versus carbon number for N3O1 and N4O1 class species at different temperatures. Top: N3O1, corresponding to ketones; bottom: N4O1, corresponding to nitrogen-containing ketones.</p>
Full article ">Figure 5
<p>Relative abundance of different ketones. Aliphatic ketones, alicyclic ketones, and aromatic ketones are classified by DBE values of 1, 2–4, and 5+, respectively.</p>
Full article ">
20 pages, 1692 KiB  
Article
Group Contribution Revisited: The Enthalpy of Formation of Organic Compounds with “Chemical Accuracy” Part IV
by Robert J. Meier and Paul R. Rablen
Thermo 2023, 3(2), 289-308; https://doi.org/10.3390/thermo3020018 - 26 May 2023
Cited by 4 | Viewed by 1802
Abstract
Group contribution (GC) methods to predict thermochemical properties are eminently important to process design. Following earlier work which presented a GC model in which, for the first time, chemical accuracy (1 kcal/mol or 4 kJ/mol) was accomplished, we here discuss classes of molecules [...] Read more.
Group contribution (GC) methods to predict thermochemical properties are eminently important to process design. Following earlier work which presented a GC model in which, for the first time, chemical accuracy (1 kcal/mol or 4 kJ/mol) was accomplished, we here discuss classes of molecules for which the traditional GC approach does not hold, i.e., many results are beyond chemical accuracy. We report new ring-strain-related parameters which enable us to evaluate the heat of formation of alkyl-substituted cycloalkanes. In addition, the definition of the appropriate group size is important to obtain reliable and accurate data for systems in which the electron density varies continuously but slowly between related species. For this and in the case of ring strain, G4 quantum calculations are shown to be able to provide reliable heats of formation which provide the quantitative data which we can use, in the case of absence of experimental data, to establish group and nearest-neighbour interaction parameters to extend the range of applicability of the GC method whilst retaining chemical accuracy. We also found that the strong van der Waals that overlap in highly congested branched alkanes can be qualitatively investigated by applying DFT quantum calculations, which can provide an indication of the GC approach being inappropriate. Full article
(This article belongs to the Special Issue Feature Papers of Thermo in 2023)
Show Figures

Scheme 1

Scheme 1
<p>Illustration of the corrections accounting for the methyl–methyl interactions we have introduced (taken from [<xref ref-type="bibr" rid="B8-thermo-03-00018">8</xref>]).</p>
Full article ">Scheme 2
<p>Chemical Structures for 3,3,4,4-tetraethylhexane (<bold>left</bold>) and ‘2,2,5,5-tetraethylhexane’, (formally correct chemical name 2,6-di methyl-2,6-diethyldecane (<bold>right</bold>).</p>
Full article ">Scheme 3
<p>Structure of p-ethyl-t-butylbenzene and 4-tert-butyltoluene.</p>
Full article ">Scheme 4
<p>Chemical structures for 2,2,4,4-tetramethyl-3-pentanol (<bold>left</bold>) and ‘1,1,5,5-tetramethyl-3-pentanol’ (<bold>right</bold>); the latter name is formally incorrect but is used to indicate it is the structure with the methyl groups shifted by one carbon atom on each end.</p>
Full article ">Scheme 5
<p>Chemical structures for 2,2,4,4-tetramethyl-iPr-3-pentanol (<bold>left</bold>) and ‘1,1,5,5-tetramethyl-iPr-3-pentanol’ (<bold>right</bold>), the latter name is formally incorrect but to indicate it is the structure with the methyl groups shifted by one carbon atom on each end.</p>
Full article ">
29 pages, 5960 KiB  
Article
Effect of Cyclohexane on the Combustion Characteristics of Multi-Component Gasoline Surrogate Fuels
by Shunlu Rao, Zhaolei Zheng and Chao Yang
Molecules 2023, 28(11), 4273; https://doi.org/10.3390/molecules28114273 - 23 May 2023
Cited by 3 | Viewed by 2547
Abstract
It has been discovered that there is a dynamic coupling between cycloalkanes and aromatics, which affects the number and types of radicals, thereby controlling the ignition and combustion of fuels. Therefore, it is necessary to analyze the effects of cyclohexane production in multicomponent [...] Read more.
It has been discovered that there is a dynamic coupling between cycloalkanes and aromatics, which affects the number and types of radicals, thereby controlling the ignition and combustion of fuels. Therefore, it is necessary to analyze the effects of cyclohexane production in multicomponent gasoline surrogate fuels containing cyclohexane. In this study, a five-component gasoline surrogate fuel kinetic model containing cyclohexane was first verified. Then, the effect of cyclohexane addition on the ignition and combustion performance of the surrogate fuel was analyzed. This study shows that the five-component model exhibits good predictive performance for some real gasoline. Meanwhile, the addition of cyclohexane decreases the ignition-delay time of the fuel in the low and high temperature bands, which is caused by the early oxidation and decomposition of cyclohexane molecules, generating more OH radicals; in the medium temperature band, the isomerization and decomposition reactions of cyclohexane oxide cC6H12O2 dominate the temperature sensitivity of the ignition delay, affecting the small molecule reactions that promote the generation of reactive radicals such as OH, thus inhibiting the negative temperature coefficient behavior of the surrogate fuel. The laminar flame speed of the surrogate fuels increased with the increase in the proportion of cyclohexane. This is due to the fact that the laminar flame speed of cyclohexane is higher than that of chain and aromatic hydrocarbons, and the addition of cyclohexane dilutes the ratio of chain and aromatic hydrocarbons in the mixture. In addition, engine simulation studies have shown that at higher engine speeds, the five-component surrogate fuel containing cyclohexane requires lower intake-gas temperatures to achieve positive ignition and are closer to the in-cylinder ignition of real gasoline. Full article
(This article belongs to the Section Green Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Simulation of ignition-delay time of FACE G gasoline shock tube with three surrogate mixtures.</p>
Full article ">Figure 2
<p>Three surrogate mixtures simulate the laminar flame speed of Exxon 708629-60 gasoline.</p>
Full article ">Figure 3
<p>HCCI engine experiments and simulated pressure curves of gasoline and its three surrogate mixtures under different compression ratios.</p>
Full article ">Figure 4
<p>Ignition-delay time of fuels 1–5 under different pressures, equivalence ratio of 0.5, and temperature range of 600–1250 K.</p>
Full article ">Figure 5
<p>Ignition-delay time of fuels 1 and 5 in the pressure range of 1–10 MPa at different temperatures (equivalence ratio of 0.5).</p>
Full article ">Figure 6
<p>Temperature sensitivity at 690 K.</p>
Full article ">Figure 7
<p>OH mole fraction and corresponding ignition-delay time of five surrogate fuels at low temperature 690 K.</p>
Full article ">Figure 8
<p>Temperature sensitivity at 800 K.</p>
Full article ">Figure 9
<p>OH mole fraction and corresponding ignition-delay time of five surrogate fuels at medium temperature 800 K.</p>
Full article ">Figure 10
<p>Temperature sensitivity at 1200 K.</p>
Full article ">Figure 11
<p>OH mole fraction and corresponding ignition-delay time of five surrogate fuels at high temperature 1200 K.</p>
Full article ">Figure 12
<p>Laminar flame speed of fuels 1–5 in the range of equivalence ratio 0.4–1.5 at different temperatures and pressures.</p>
Full article ">Figure 13
<p>Laminar flame speed of fuels 1–5 in different temperature and pressure ranges when equivalence ratio is 1.</p>
Full article ">Figure 14
<p>Sensitivity of laminar flame speed under different temperatures and pressures.</p>
Full article ">Figure 15
<p>H, OH concentrations and laminar flame speed of fuels 1 and 5 at temperature 298 K, pressure 0.1 MPa, and equivalence ratio 0.5–1.5.</p>
Full article ">Figure 16
<p>H, OH concentrations and laminar flame speed of fuels 1 and 5 at temperature 500 K, pressure 2 MPa, and equivalence ratio 0.5–1.5.</p>
Full article ">Figure 17
<p>Required intake temperatures at an intake pressure of 0.1MPa to phase combustion at TDC in an HCCI engine.</p>
Full article ">Figure 18
<p>Calculated temperature, CO mole fraction, and each component conversion for fuel 1 and fuel 5 at 600 rpm engine speed.</p>
Full article ">Figure 18 Cont.
<p>Calculated temperature, CO mole fraction, and each component conversion for fuel 1 and fuel 5 at 600 rpm engine speed.</p>
Full article ">Figure 19
<p>Calculated temperature, CO mole fraction, and each component conversion for fuel 1 and fuel 5 at 1200 rpm engine speed.</p>
Full article ">Figure 20
<p>Partial intermediate product mole fractions calculated for fuel 1 and fuel 5 at 1200 rpm.</p>
Full article ">Figure 21
<p>Calculated C<sub>4</sub>H<sub>6</sub> and C<sub>6</sub>H<sub>6</sub> reaction rates for fuel 1 and fuel 5 at 1200 rpm.</p>
Full article ">
1284 KiB  
Proceeding Paper
The Process of Isolation, Using Crystallization of Cis- and Trans-Isomers, of Perfluorodecalines from an Industrial Mixture of Electrochemical Fluorination of Napthaline
by Aleksey V. Kisel, Andrei V. Polkovnichenko, Egor V. Lupachev, Nikolai N. Kuritsyn, Sergey Y. Kvashnin and Nikolai N. Kulov
Eng. Proc. 2023, 37(1), 85; https://doi.org/10.3390/ECP2023-14640 - 17 May 2023
Cited by 3 | Viewed by 738
Abstract
The process of crystallization separation of an industrial mixture of perfluorinated cycloalkanes was considered. The content of target products, cis- and trans-isomers of perfluorodecalin (PFD), in all initial fractions of the investigated samples of the reaction mixture was at least 70 wt.%. Based [...] Read more.
The process of crystallization separation of an industrial mixture of perfluorinated cycloalkanes was considered. The content of target products, cis- and trans-isomers of perfluorodecalin (PFD), in all initial fractions of the investigated samples of the reaction mixture was at least 70 wt.%. Based on the experimental data, the dependences of the crystallization (partition) coefficients between the solid and mother liquor (Ds/l), the enrichment factor (ηs), and the separation factor (Fs) on the ratio of trans-PFD to cis-PFD, the ratio of mother liquor to solid, and crystallization temperature in the range −50–15 °C were obtained. It was shown that the values Ds/l and ηs depended significantly on the concentration of trans-PFD in the initial solution, and that the value of Fs decreased as the process temperature rose. Full article
Show Figures

Figure 1

Figure 1
<p>The dependence of the partition coefficient between the solid and mother liquor <math display="inline"><semantics> <mrow> <msubsup> <mi>D</mi> <mi>i</mi> <mrow> <mi>s</mi> <mo>/</mo> <mi>l</mi> </mrow> </msubsup> </mrow> </semantics></math> on the ratio of trans-PFD/cis-PFD in the initial fraction (<span class="html-italic">f<sub>trans</sub></span>/<span class="html-italic">f<sub>cis</sub></span>). Red—trans-PFD; blue—cis-PFD; black—∑PFD.</p>
Full article ">Figure 2
<p>The dependence of the partition coefficient between the solid and mother liquor (<math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mrow> <mi>s</mi> <mo>/</mo> <mi>l</mi> </mrow> </msup> </mrow> </semantics></math>): (<b>a</b>) on the ratio of trans-PFD/cis-PFD in the initial fraction (<span class="html-italic">f<sub>i</sub></span>/<span class="html-italic">f<sub>j</sub></span>) for components of the mixture at <span class="html-italic">T<sub>sep</sub></span> = −44 °C; (<b>b</b>) on the ratio of BCH/∑PFD in the initial fraction (<span class="html-italic">f<sub>i</sub></span>/<span class="html-italic">f<sub>j</sub></span>) for ∑PFD at different nomenclature and quantity of unidentified impurities at <span class="html-italic">T<sub>sep</sub></span> = −44 °C; (<b>c</b>) on the concentration of BCH in the initial fraction (<span class="html-italic">f</span><sub>BCH</sub>) for ΣPFD at different <span class="html-italic">T<sub>sep</sub></span>. Dots: red—trans-PFD; blue—cis-PFD; black—∑PFD; green—BCH.</p>
Full article ">Figure 3
<p>The dependence of the enrichment factor of PFD into solid <math display="inline"><semantics> <mrow> <msubsup> <mi>η</mi> <mrow> <mo>∑</mo> <mi>P</mi> <mi>F</mi> <mi>D</mi> </mrow> <mi>s</mi> </msubsup> </mrow> </semantics></math> on the ratio of trans-PFD/cis-PFD in the initial fraction (ftrans/fcis) at different ratios of mother liquor to solid (L/S).</p>
Full article ">Figure 4
<p>The dependence of the separation factor of cis-PFD <span class="html-italic">F<sup>s</sup></span> on the ratio <span class="html-italic">f<sub>BCH</sub></span>/<span class="html-italic">f<sub>PFD</sub></span> at different <span class="html-italic">T<sub>sep</sub></span> values.</p>
Full article ">
Back to TopTop