Next Article in Journal
Inverse Method-Based Kinetic Modelling and Process Optimization of Reverse-Phase Chromatography for Molnupiravir Synthesis
Next Article in Special Issue
Geochemical Characteristics and Hydrocarbon Generation Potential of Coal-Measure Source Rocks in Julu Sag
Previous Article in Journal
Technical Support System for High Concurrent Power Trading Platforms Based on Microservice Load Balancing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research on Mechanism of Surfactant Improving Wettability of Coking Coal Based on Molecular Dynamics

School of Resource, Environment and Safety Engineering, Hunan University of Science and Technology, Xiangtan 411201, China
*
Author to whom correspondence should be addressed.
Processes 2024, 12(6), 1271; https://doi.org/10.3390/pr12061271
Submission received: 19 May 2024 / Revised: 14 June 2024 / Accepted: 18 June 2024 / Published: 20 June 2024

Abstract

:
Coal dust is a major safety hazard in the process of coal mining and is of great importance to ensure production safety and maintain the health of operators. In order to understand the microscopic mechanism during coal seam water injection and reveal the mechanism of surfactants in improving the wettability of coal dust, coking coal was selected as the research object. Three surfactants, SDBS, AEO-9, and CAB-35, were chosen for molecular dynamics simulation research on the wetting and adsorption properties of water/coal/surfactants. The results show that surfactant molecules can cover the hydrophobic groups on the surface of coking coal, forming a hydrophilic adsorption layer, changing the coal surface from hydrophobic to hydrophilic, and enhancing the wettability. After adding surfactants, the thickness of the adsorption layer in the z-axis direction increases, expanding the contact area between coking coal and water molecules, thereby improving the wettability. When surfactants tightly cover the surface of coking coal, their binding strength increases, forming a more stable hydrophilic layer and further improving the wettability. At the same time, surfactants promote the diffusion of water molecules and enhance the interaction between hydrophobic alkyl chains and water molecules, further enhancing the wetting effect.

1. Introduction

Coal resources play an irreplaceable role in the Chinese economy, supporting social production and driving the development of the national economy [1,2,3]. However, safety hazards in coal mining, especially coal dust safety issues, involving occupational hazards [4,5,6], explosion hazards [7,8,9], environmental pollution [10,11,12], and mechanical equipment damage [13,14,15], have become a major challenge that urgently needs to be addressed. The “Healthy China” initiative emphasizes public safety systems and the development of occupational health [16,17]. In the coal mining industry, safety production is particularly crucial. Therefore, coal dust generation should be controlled at the source. Statistics as of 2022 show that the annual number of new patients with occupational pneumoconiosis in China has decreased to 7577, but the number of deaths caused by pneumoconiosis remains at up to 9613 each year. Therefore, in the current strategy for coal dust prevention and control, more emphasis must be placed on managing the dust concentration in coal mining faces and controlling the incidence of pneumoconiosis.
The wettability of coal is an important factor in improving the efficiency of dust reduction. Coking coal itself is highly hydrophobic, so in the process of coal mining, coal dust is a serious concern that seriously affects the operation of the staff and jeopardizes their health. Chemical modification can effectively alter the surface properties of coal, enhance its performance, and provide a new approach for the efficient utilization of coal resources. Hence, many scholars have studied the wettability of coal dust. With the increasing degree of metamorphism of the coal, its wettability shows a changing trend of increasing, then decreasing, and then increasing again. Among bituminous coals, long-flame coals show the strongest wettability, while coking coals show relatively stronger hydrophobicity, i.e., their wettability is weaker [18]. Referring to previous studies, the addition of surfactants to the dust removal system can substantially improve the dust removal efficiency [19]. Based on contact angle experiments and reverse osmosis experiments to study the effect of surfactant concentration on the contact angle and dust removal efficiency of coal dust, it was found that with an increase in surfactant concentration, the wettability of coal dust was enhanced [20]. When the surfactant concentration increases to the critical micelle concentration, the wetting effect of the solution on coal dust reaches its optimum state. Among the different types of surfactants, anionic surfactants had the most significant wetting effect on coal dust. While scanning electron microscopy experiments and infrared spectroscopy experiments were used to study the physical properties of coal [21], some scholars have investigated the modification effects of surfactants and ionic liquids on the physicochemical properties of acidified coal, showing that HNO3 and SLS can improve the wettability of coal and increase its porosity [22]. SDS promoted the depressing effect of acidified coal bodies and reduced the strength of the treated coal samples [23]. Some scholars have mainly focused on the in-depth R&D and preparation of dust suppressants in order to be able to improve the wettability of coal more effectively [24,25,26]. Some scholars have used molecular simulation to reveal at depth the wetting mechanism of coal dust surface and concluded that an important factor affecting the wettability of coal surface is the ability of hydrophilic functional groups on the coal surface to form hydrogen bonds with water molecules [27,28,29].
Currently, adding surfactants to coal dust prevention and control has shown better wetting effects [30,31,32]. However, available studies mainly focus on the impact of water injection parameters and methods on the water injection effect, while research on how to enhance the water injection effect by changing coal quality characteristics is still insufficient. In this study, coking coal was selected as the research object, and we established the coking coal model based on elemental analysis, 13CNMR, and XPS analysis. In addition, complex systems, including water/coal and water/coal/surfactant, were built. The wetting adsorption performance of different systems was simulated at the molecular level, aiming to deeply analyze the wetting mechanism and reveal the microscopic mechanisms during coal seam water injection. The research results provide theoretical support for the application of surfactants in coal seam water injection.

2. Modeling

2.1. Experimental Materials

The coal samples selected for the experiment were coking coal from Wanfeng Coal Mine in Shanxi Province, China. The industrial analysis and elemental analysis of the coking coal are shown in Table 1. The coal samples were crushed for 1 min with a crusher and then sieved with a 150-mesh standard industrial sieve before being sealed in a container for storage. Figure 1 shows the particle size distribution of coking coal dust, with the particle size mainly within 0–200 μm. The particle size distribution is similar, so they can be used as experimental samples.
Based on relevant domestic and international studies, the surfactants selected in this study include sodium dodecylbenzene sulfonate (SDBS) (Wanhua (Guangzhou) Supply Chain Service Co., Ltd., Guangzhou, China), nonylphenol ethoxylate (AEO-9) (Wanhua (Guangzhou) Supply Chain Service Co., Ltd., Guangzhou, China), and cocoamidopropyl betaine (CAB-35) (Wanhua (Guangzhou) Supply Chain Service Co., Ltd., Guangzhou, China).

2.2. Coking Coal Modeling

2.2.1. Carbon Atom Analysis

The different forms of carbon in coal and their relative content can be effectively determined by 13CNMR technology. Figure 2 shows the 13CNMR spectrum of coking coal. The 13CNMR spectrum of coal samples includes three parts: the aliphatic carbon region from 0 to 75 ppm, the aromatic carbon region from 100 to 165 ppm, and the carbonyl carbon region from 165 to 230 ppm.
For the nuclear magnetic resonance (NMR) carbon spectrum of coking coal, data processing was performed by peak fitting to determine the carbon assignment and peak area, as shown in Table 2. On this basis, the 12 main structural parameters of coking coal were calculated and determined. The specific results are listed in Table 3.
The ratio of aromatic bridging carbon to peripheral carbon (i.e., the bridging-to-peripheral ratio) XBP is an important parameter for the molecular structure of coal. XBP is calculated based on the formula in the table:
X B P = f a B f a H + f a P + f a S
From the distribution of different types of carbon atoms in the coal molecular carbon skeleton, it can be seen that in aliphatic carbon atoms, the content of methylene or methyl groups is much higher than that of methyl or methoxy groups. The faH is above 44%, indicating that the large molecular structure of coal is almost entirely composed of protonated aromatic carbon. In this study, the XBP of coking coal is 0.24.

2.2.2. Elemental Structural Analysis

In the complex structure of coal, various elements such as carbon, nitrogen, oxygen, and sulfur exist in the form of functional groups. Figure 3 shows the XPS spectrum after peak fitting. According to the results of peak fitting, the predominant form of carbon elements in the coking coal molecular structure is C-C/C-H, supplemented by C-O. Nitrogen mainly exists in pyridine-type nitrogen, supplemented by nitrogen oxides. Oxygen mainly exits in the form of C-O, supplemented by C=O. Sulfur mainly exists in the form of thiophene-type sulfur, supplemented by sulfur oxides.

2.2.3. Comprehensive Analysis

A comprehensive analysis was carried out on the aromatic structure, aliphatic carbon structure, and heteroatom structure of coal in combination with the above analyses.
(1)
Aromatic structure
Aromatic macromolecules are the main structures forming the backbone of coal molecules. The higher the proportion of aromatic macromolecules to cycloalkanes, the higher the degree of condensation. Table 4 shows the forms of aromatic structures in coking coal molecules. For aromatic structures, the XBP of benzene, naphthalene, anthracene or phenanthrene, dibenzofuran, and dibenzothiophene is 0, 0.25, 0.4, 0.5, and 0.57, respectively. The XBP of coking coal is 0.24, which is between benzene and naphthalene and closer to naphthalene, indicating that there is a structure with two bridging carbons, with benzene and naphthalene being the main components of the original coal. The types and quantities of aromatic structures are listed in Table 4. Actual coal molecular compositions can form aromatic structures with an XBP of 0.25 when connected to benzene rings, i.e., aromatic structures with the same bridging carbon perimeter ratio as naphthalene.
The mass fraction of C in coking coal is 89.38%. The NMR data analysis reveals that the ratio of aromatic bridging carbon to peripheral carbon in coking coal XBP is 0.24, indicating that the ratio of bridging carbons to peripheral carbons in aromatic compounds with degrees of condensation 1 and 2 is 0 and 0.25, respectively. From this, it can be inferred that the molecular model of coking coal is primarily composed of naphthalene, followed by benzene rings. The bridging-to-peripheral ratio is obtained through the combination calculation of different aromatic structural units. Moreover, the types and quantities of aromatic structural units are determined. At this point, the total number of aromatic ring carbons in the model is 246. Additionally, based on the results of 13CNMR, the proportion of aromatic carbons in coking coal reaches 65.98%, estimating a total of 373 carbon elements in the coal molecule.
(2)
Aliphatic carbon structure
Ethyl side chains, methyl groups, and cycloalkanes are the primary forms of aliphatic structures in coal. The coal quality is positively correlated with the carbon–hydrogen ratio; the higher the coal quality, the higher the carbon–hydrogen ratio, while the quantities of cycloalkanes and aliphatic side chains decrease accordingly. Aliphatic structures tend to exist in the form of cycloalkanes. With a carbon mass fraction of 89.38%, the average number of atoms in alkyl side chains ranges from 1 to 2, indicating that alkyl side chains in the coal structure should not be too long, with short chains being predominant. For coking coal, falH = 21.29% and fal* = 7.93%, indicating that the relative percentage of methylene, methine, and quaternary carbons is greater than the relative percentage of methyl groups. Since the aromaticity of the coal is 65.98% and the number of aromatic atoms is 246, the number of aliphatic carbon atoms is tentatively fixed at 129.
(3)
Heteroatom structures
The heteroatom structures in coal are mainly composed of nitrogen and sulfur atoms. Nitrogen atoms exist in the form of pyrrole or pyridine, while sulfur atoms exist in the form of thiols, thioethers, or aromatic compounds. Table 5 shows the forms of aromatic structures in coking coal molecules. The number of carbon atoms in coking coal molecules is 373. The coal sample contains a certain amount of nitrogen, while the sulfur content is generally very low. Based on the atomic ratios of each element to carbon, the ratio of nitrogen atoms to oxygen atoms to sulfur atoms in coking coal is 8:11:2.
The specific forms and distribution of aromatic carbon, aliphatic carbon, and heteroatoms in coal are determined by calculation and analysis. Based on this and combined with the results of elemental analysis, an initial structural model of the coking coal macromolecule was built with the Chemdraw 19.0. The molecular formula of coking coal is determined to be C352H320N8O11S2, as shown in Figure 4.

2.3. System Modeling

In this study, the Forcite module in Materials Studio 8.0 software was used, with the COMPASS II force field based on quantum mechanics selected. Layers of coal molecules and water molecules were constructed with the Amorphous Cell module and were merged into a system with a vacuum layer density of 20 Å. Four wetting adsorption system models were built. Figure 5 shows System B and simulates the reaction process between coking coal and SDBS solution. System A represents the water/coal system, System C is the water/coal/AEO-9 system, and System D is the water/coal/CAB-35 system. The Forcite module was chosen to geometrically optimize the different systems constructed, ensuring the molecular energy reaches its minimum and the structure becomes more stable. Dynamic tasks were executed after optimization. A nose thermostat was used in a 298K NVT system for molecular dynamics calculations for 300 ps with a time step of 1.0 fs and a cutoff set to 12.5 Å. The CPU time required for a system simulation was about 4.5 h.

3. Study on Wetting Adsorption Performance of Water/Coal/Surfactant

3.1. Spatial Distribution Characteristics

The states after the simulation of different systems are shown in Figure 6. In the initial configuration before simulation, the two molecular layers do not contact each other, and the surfactant molecules are randomly distributed in the aqueous solution. After simulation, there are significant changes in the positions of the two molecular layers. In Systems B, C, and D, there is a noticeable interaction between the surfactant solution layer and the coal surface. This interaction causes the solution layer to move towards the coal surface, finally reaching a new equilibrium state. The surfactant adsorbs onto the coal surface and aligns in a specific manner: the hydrophilic head group faces the water phase, while the hydrophobic tail chain points towards the coal molecules. This effectively covers the hydrophobic groups on the coal sample surface, forming an oriented adsorption layer with the hydrophilic end facing the water phase, leading to a significant change in the surface properties of the coal sample from hydrophobic to hydrophilic.

3.2. Relative Concentration Distribution Characteristics

In order to quantitatively describe the spatial distribution of this adsorption behavior and to delve into the adsorption characteristics of the water/coal and water/coal/surfactant systems, relative concentration distribution curves along the z-axis were plotted based on simulation data, as shown in Figure 7.
It can be observed from the figure that the thickness of the adsorption layer at the water/coal interface in System A is only 16.04 Å, indicating a certain repulsion between the aqueous solution and coking coal due to poor wettability. In System B, the thickness of the adsorption layer of the surfactant SDBS in the z-axis direction is the largest, at 20.34 Å; in System C, the thickness of the adsorption layer of the surfactant AEO-9 on water/coal is 17.57 Å; in System D, the thickness of the adsorption layer of the surfactant CAB-35 on water/coal is 16.13 Å. A greater thickness of the adsorption layer in the z-axis direction indicates a larger contact area between coal and water molecules, implying better wettability. The results of molecular dynamics simulations reflect the enhanced ability of the SDBS solution to promote water molecule movement and indicate its advantage in improving the wettability of the coking coal surface.

3.3. Analysis of Non-Bonding Interaction Energy

The energy in the adsorption system mainly comes from non-bonding interaction energy (Eint), where van der Waals energy (Evan) and electrostatic energy (Eele) are the dominant factors during the interaction and adsorption processes. Non-bonding interaction energy between molecules is generally negative, and the magnitude of its absolute value reflects the strength of the interaction. Table 6 shows the interaction energies for the four systems. It can be observed that Ein for the four systems is −9416.006, −9906.651, −9857.330 kcal/mol, and −9652.390 kcal/mol, all of which are negative. The absolute values of Ein follow the order: System B > System C > System D > System A. In all four systems, the process of water molecules adsorbing on the coal surface is spontaneous. When surfactant molecules cover the coal surface, their binding ability significantly increases, leading to an enhancement in the wettability on the surface of coking coal. Particularly, the effect of the SDBS solution on improving the wettability of the surface of coking coal is notably significant. The analysis of Evan and Eele within Ein showed that in the water/coal/surfactant system, Eele contributes more, indicating that the surfactant molecules’ head groups primarily adsorb to water molecules through electrostatic interactions.

3.4. Diffusion Analysis

The effect of surfactants on the aggregation behavior of water molecules was revealed by an in-depth analysis of mean square displacement (MSD) and its corresponding diffusion coefficient D; the mechanism of the effect of different surfactants on the wettability of coal was further clarified. The specific calculation formulas are as follows:
M S D = 1 N i = 1 N [ r i ( t ) r i ( 0 ) ]   2
D = 1 6 N lim t d d t i = 1 N [ r i ( t ) r i ( 0 ) ]   2
where MSD is the mean square displacement; N is the number of diffusing molecules; r(t) and r(0) are the position vectors of the molecule at time t and t = 0, respectively; and D is the diffusion coefficient.
From the MSD curves of water molecules shown in Figure 8, it can be seen that the slope of the MSD curve in systems with surfactants is significantly greater than in systems without surfactants. This is attributed to surfactants enhancing the interaction between water and coal surfaces, promoting the vigorous movement of water molecules. The system with water/coal has the smallest MSD slope; adding surfactants increases the slope; and in the water/coal/surfactant SDBS system, the slope increase is the highest. This is due to the different hydrophobic alkyl chains in various surfactants exerting varying forces on water molecules, influencing their effect on coal wettability. The results indicate that surfactants have a promoting effect on the diffusion of water molecules. The diffusion coefficients are shown in Table 7. As the diffusion coefficient of water molecules increases, the interaction between hydrophobic alkyl chains and water molecules is also significantly strengthened. This enhanced effect further promotes the wetting effect between coal and water, thereby contributing to the improvement in coal’s wettability.

4. Conclusions

This study explores the mechanism by which surfactants improve the wettability of coking coal based on molecular dynamics and reveals the intrinsic mechanism of water injection into coal seams at a molecular level. The simulation results demonstrate that surfactant molecules effectively cover the hydrophobic groups on the surface of coking coal and form an oriented adsorption layer with hydrophilic ends facing the water phase, significantly enhancing the wettability on the coal surface. After adding surfactants, the thickness of the adsorption layer in the z-axis direction significantly increases, indicating an expanded contact area between coal and water molecules and an enhanced wettability. When surfactant molecules tightly cover the coal surface, their binding ability is significantly enhanced, which not only helps the formation of a more stable hydrophilic layer but also improves the wettability on the surface of the coking coal. Furthermore, surfactants have a promoting effect on the diffusion of water molecules. As the diffusion coefficient of water molecules increases, the interaction between hydrophobic alkyl chains and water molecules is also significantly strengthened. Such strengthened interaction further promotes the wetting effect between coal and water, helping improve the wettability of coal.

Author Contributions

R.L.: methodology, validation, investigation, visualization, and review. S.L.: conceptualization and writing—review and editing. Y.L.: data curation and writing—original draft. Y.Z.: visualization and review. W.L.: visualization and review. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Publicly available datasets were analyzed in this study. These data can be found here: Part of the datasets are available in previous publications and were reworked. Databases associated with the exposure/vulnerability will be available under request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kori, K.B.; Agrawal, H.D. Dust Monitoring Systems and Health Hazards in Coal Mining A Review. J. Trend Sci. Res. Dev. 2021, 5, 172–177. [Google Scholar]
  2. Zhao, S.F.; Chen, X.; Zhao, X. Analysis of current status and development direction of coal mining technology in China. AIP Conf. Proc. 2019, 2066, 020017. [Google Scholar] [CrossRef]
  3. Zhang, C.; Wang, P.; Wang, E.; Chen, D.; Li, C. Characteristics of coal resources in China and statistical analysis and preventive measures for coal mine accidents. Int. J. Coal Sci. Technol. 2023, 10, 22. [Google Scholar] [CrossRef]
  4. Zhang, J.F.; Yang, F.F.; Yadong Xie, Y.D.; Zhang, J.J.; Miao, Z.Q.; Ma, D.Q. Study on Dust Control of Coal Seam Shallow Hole Dynamic Pressure Water Injection Based on Prevention and Treatment of Coal Miner’s Pneumoconiosis. Basic Clin. Pharmacol. Toxicol. 2020, 126, 274. [Google Scholar]
  5. Ma, C.B. Exploration of the Application of Green Mining Technology in Coal Mines under the New Situation. Appl. Sci. Innov. Res. 2024, 8, 1. [Google Scholar] [CrossRef]
  6. Luo, H.T.; Zhou, W.; Mithal, I.J.; Wang, Z.M. Analyzing Characteristics of Particulate Matter Pollution in Open-Pit Coal Mines: Implications for Green Mining. Energies 2021, 14, 2680. [Google Scholar] [CrossRef]
  7. Sahu, A.; Mishra, P.D. Prevention and suppression of coal dust explosion in underground coal mines: Role of rock dust type, particle size, proportion, concentration, and thermal properties. Adv. Powder Technol. 2024, 35, 104343. [Google Scholar] [CrossRef]
  8. Liu, T.Q.; Gao, Z.Y.; Yanying Xu, Y.Y.; Duan, G.S.; Wang, X. Research on Explosion Pressure Characteristics of Long Flame Coal Dust and the Inhibition Effect of Different Explosion Suppressants. ACS Omega 2023, 8, 35919–35928. [Google Scholar] [CrossRef]
  9. Cheng, C.L.; Si, R.J.; Wang, L.; Jia, Q.S.; Xin, C.P. Explosion and explosion suppression of gas/deposited coal dust in a realistic environment. Fuel 2024, 357, 129710. [Google Scholar] [CrossRef]
  10. Wang, P.F.; Tan, X.H.; Zhang, L.Y.; Li, Y.J.; Liu, R.H. Influence of particle diameter on the wettability of coal dust and the dust suppression efficiency via spraying. Process Saf. Environ. Prot. 2019, 132, 189–199. [Google Scholar] [CrossRef]
  11. Han, H.; Wang, P.F.; Li, Y.J.; Liu, R.H.; Tian, C. Effect of water supply pressure on atomization characteristics and dust-reduction efficiency of internal mixing air atomizing nozzle. Adv. Powder Technol. 2020, 31, 252–268. [Google Scholar] [CrossRef]
  12. Wang, P.F.; Gao, R.Z.; Liu, R.H.; Yang, F.Q. CFD-based optimization of the installation location of the wall-mounted air duct in a fully mechanized excavation face. Process Saf. Environ. Prot. 2020, 141, 234–245. [Google Scholar] [CrossRef]
  13. Shen, H.G.; Xu, X.Y.; Wei, J.P.; Jiang, W.; Wang, M.Y. Experimental study on the effect of cutting parameters to dust production patterns of different brittle coal. Int. J. Coal Prep. Util. 2024, 44, 154–169. [Google Scholar] [CrossRef]
  14. Zhou, G.; Duan, J.; Sun, B.; Jing, B.; Kong, Y.; Zhang, Y.; Ni, G.; Sun, L. Numerical analysis on pollution law for dust and diesel exhaust particles in multi-ventilation parameter environment of mechanized excavation face. Process Saf. Environ. Prot. 2022, 157, 320–333. [Google Scholar] [CrossRef]
  15. Lu, X.X.; Shen, C.; Xing, Y.; Zhang, H.; Wang, C.Y.; Shi, G.Y.; Wang, M.Y. The spatial diffusion rule and pollution region of disorganized dust in the excavation roadway at different roadheader cutting positions. Powder Technol. 2022, 396, 167–180. [Google Scholar] [CrossRef]
  16. Wang, H.Q.; Ye, Q.; Zhang, H.D.; Sun, X.; Li, T. Prevention and Treatment of Pneumoconiosis in the Context of Healthy China 2030. China CDC Wkly. 2023, 5, 927–932. [Google Scholar] [CrossRef]
  17. Wang, J.; Shi, H.D.; Wang, P.; Wu, Y.G.; Wang, L.L.; Song, Z.L. Analysis of early diagnosis and prevention techniques for occupational pneumoconiosis based on patent perspectives in China’s coal industry. Zhonghua Lao Dong Wei Sheng Zhi Ye Bing Za Zhi = Zhonghua Laodong Weisheng Zhiyebing Zazhi = Chin. J. Ind. Hyg. Occup. Dis. 2023, 41, 948–955. [Google Scholar]
  18. Liao, X.X.; Wang, B.; Wang, L.; Zhu, J.T.; Chu, P.; Zhu, Z.B.; Zheng, S.W. Experimental Study on the Wettability of Coal with Different Metamorphism Treated by Surfactants for Coal Dust Control. ACS Omega 2021, 6, 21925–21938. [Google Scholar] [CrossRef]
  19. Solodyankin, S.S.; Kolmakov, N.G.; Manin, N.S.; Fritslet, V.K.; Kazantsev, A.V.; Miroshnikov, A.M. Using a solution of the surfactant for increasing collection efficiency of coal dust in the exhaust system. Coke Chem. 2016, 59, 333–337. [Google Scholar] [CrossRef]
  20. Yang, L.; Ge, S.C.; Huang, Z.H.; Jing, D.J.; Chen, X. The influence of surfactant on the wettability of coal dust and dust reduction efficiency. Arab. J. Geosci. 2021, 14, 1336. [Google Scholar] [CrossRef]
  21. Cummings, J.; Shah, K.; Atkin, R.; Moghtaderi, B. Physicochemical interactions of ionic liquids with coal; the viability of ionic liquids for pre-treatments in coal liquefaction. Fuel 2015, 143, 244–252. [Google Scholar] [CrossRef]
  22. Han, W.B.; Zhou, G.; Zhang, Q.T.; Pan, H.W.; Liu, D. Experimental study on modification of physicochemical characteristics of acidified coal by surfactants and ionic liquids. Fuel 2020, 266, 116966. [Google Scholar] [CrossRef]
  23. Ni, G.H.; Xie, H.C.; Li, S.; Sun, Q.; Huang, D.M.; Cheng, Y.Y.; Wang, N. The effect of anionic surfactant (SDS) on pore-fracture evolution of acidified coal and its significance for coalbed methane extraction. Adv. Powder Technol. 2019, 30, 940–951. [Google Scholar] [CrossRef]
  24. Zhang, Z.; Zhang, S. A new method of coal fine particles humidification and agglomeration: Synergistic dust suppression with composition of soap solution. Process Saf. Environ. Prot. 2022, 159, 146–156. [Google Scholar] [CrossRef]
  25. Zhou, L.; Yang, S.Y.; Hu, B.; Yuan, Z.L.; Wu, H.; Yang, L.J. Evaluating of the performance of a composite wetting dust suppressant on lignite dust. Powder Technol. 2018, 339, 882–893. [Google Scholar] [CrossRef]
  26. Hu, X.; Yang, Z.; Zhao, Y.; Dong, Y.; Wang, C.; Zhang, L.; Yu, Y.; Wu, K.; Ren, L. Medium optimization and dust suppression performance analysis of microbial-based dust suppressant compound by response surface curve method. Environ. Sci. Pollut. Res. Int. 2024, 31, 24525–24535. [Google Scholar] [CrossRef]
  27. Nie, W.; Yuan, M.; Bao, Q.; Yan, J.; Zhou, W.; Guo, C.; Guo, L.; Niu, W.; Yu, F.; Hua, Y. Experimental and molecular dynamics simulation research on compound dust suppressant based on locust bean gum. Adv. Powder Technol. 2022, 33, 103485. [Google Scholar] [CrossRef]
  28. Li, S.; Yan, D.; Yan, M.; Bai, Y.; Zhao, B.; Long, H.; Lin, H. Molecular simulation of alkyl glycoside surfactants with different concentrations inhibiting methane diffusion in coal. Energy 2023, 263, 125771. [Google Scholar] [CrossRef]
  29. Zhao, Y.; Li, H.M.; Lei, J.M.; Xie, J.; Li, L.M.; Gan, Y.; Liu, Y.L. Study on the surface wetting mechanism of bituminous coal based on the microscopic molecular structure. RSC Adv. 2023, 13, 5933–5945. [Google Scholar] [CrossRef]
  30. Wang, P.F.; Han, H.; Tian, C.; Liu, R.H.; Jiang, Y.D. Experimental study on dust reduction via spraying using surfactant solution. Atmos. Pollut. Res. 2020, 11, 32–42. [Google Scholar] [CrossRef]
  31. Crawford, R.J.; Mainwaring, D.E. The influence of surfactant adsorption on the surface characterisation of Australian coals. Fuel 2001, 80, 313–320. [Google Scholar] [CrossRef]
  32. Wang, P.F.; Jiang, Y.D.; Liu, R.H.; Liu, L.M.; He, Y.C. Experimental study on the improvement of wetting performance of OP-10 solution by inorganic salt additives. Atmos. Pollut. Res. 2020, 11, 153–161. [Google Scholar] [CrossRef]
Figure 1. Coking coal dust particle size distribution.
Figure 1. Coking coal dust particle size distribution.
Processes 12 01271 g001
Figure 2. The 13CNMR spectrum of coking coal.
Figure 2. The 13CNMR spectrum of coking coal.
Processes 12 01271 g002
Figure 3. XPS peak fitting: (A) carbon peak profile, (B) nitrogen peak profile, (C) oxygen peak profile, (D) sulfur peak profile.
Figure 3. XPS peak fitting: (A) carbon peak profile, (B) nitrogen peak profile, (C) oxygen peak profile, (D) sulfur peak profile.
Processes 12 01271 g003
Figure 4. A 2D\3D model of coking coal.
Figure 4. A 2D\3D model of coking coal.
Processes 12 01271 g004
Figure 5. Schematic diagram of the construction of system B.
Figure 5. Schematic diagram of the construction of system B.
Processes 12 01271 g005
Figure 6. State diagrams of different systems at the end of the simulation: (A) System A, (B) System B, (C) System C, (D) System D.
Figure 6. State diagrams of different systems at the end of the simulation: (A) System A, (B) System B, (C) System C, (D) System D.
Processes 12 01271 g006
Figure 7. Relative concentration distribution of different systems: (A) System A, (B) System B, (C) System C, (D) System D.
Figure 7. Relative concentration distribution of different systems: (A) System A, (B) System B, (C) System C, (D) System D.
Processes 12 01271 g007
Figure 8. MSD curves of water molecules.
Figure 8. MSD curves of water molecules.
Processes 12 01271 g008
Table 1. Industrial and elemental analysis of coal sample.
Table 1. Industrial and elemental analysis of coal sample.
CoalMad (%)Aad (%)Vad (%)FCad (%)Cd (%)Hd (%)Od (%)Nd (%)Sd (%)
Coking coal2.5612.7214.9469.7889.385.522.641.880.58
Note: in the table, Mad is air-drying moisture; Aad is ash on air-drying basis; Vad is volatile matter on air-drying basis; FCad is the fixed carbon content.
Table 2. Peak position and peak area corresponding to 13CMRN of coking coal.
Table 2. Peak position and peak area corresponding to 13CMRN of coking coal.
Coal SamplesNumber of PeaksChemical Shift (ppm)Relative Area (%)Attribution
Coking Coal118.597.93Aromatic Methylene Carbon
234.4615.09Methylene Carbon
349.856.20Quaternary and Hypomethyl Carbons
473.025.56Endocyclic Oxidized Fatty Carbons
5107.512.18Protonated Aromatic Carbon
6124.1848.59Protonated Aromatic Carbon
7134.3812.58Bridged Aromatic Carbon
8154.522.63Oxygen-Substituted Aromatic Carbons
9169.624.45Carboxy Carbon
10221.482.89Carboxy Carbon
Table 3. The 12 main structural parameters of coal samples.
Table 3. The 12 main structural parameters of coal samples.
Coal SamplesfafafacfaHfaNfaPfaSfaBfalfal*falHfalO
Coking Coal73.3265.987.3450.7715.212.63012.5834.787.9321.295.56
Notes: fa is SP2 hybridized carbon; fac is carbonyl carbon content; fa is the aromatic carbon ratio of the coal sample; faH is protonated carbon; faN, is non-protonated carbon; faP signifies phenolic hydroxyl or ether oxygen-linked carbon; faS is alkyl-substituted carbon; faB is aromatic bridging carbon; fal is SP3 hybridized carbon content; falH is methylene or methyl carbon; fal* is methyl carbon; and falO is oxygen-linked aliphatic carbon.
Table 4. Forms of aromatic structures present in coking coal molecules.
Table 4. Forms of aromatic structures present in coking coal molecules.
Forms of Existence of Aromatic StructuresCoking Coal
Thiophene (C4H4S)2
Pyridine (C5H5N)6
Pyrrole (C4H5N)2
Table 5. Forms of heteroatoms in coking coal molecules.
Table 5. Forms of heteroatoms in coking coal molecules.
ElementQuantityForm of Existence
N8C5H5N, C4H5N
O11C-O, C=O
S2C4H4S
Table 6. Action energy analysis of different systems.
Table 6. Action energy analysis of different systems.
SystemEint/(kcal·mol−1)Evan/(kcal·mol−1)Eele/(kcal·mol−1)
System A−9416.0062427.363−11891.369
System B−9906.6512395−12252.468
System C−9652.3902353.543−12005.933
System D−9857.3302375.737−12282.388
Table 7. Diffusion coefficients for the four systems.
Table 7. Diffusion coefficients for the four systems.
SystemDiffusion Coefficient (cm2/s)
System A4.77383 × 10−5
System B6.24917 × 10−5
System C5.97467 × 10−5
System D5.63033 × 10−5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, R.; Li, S.; Ling, Y.; Zhao, Y.; Liu, W. Research on Mechanism of Surfactant Improving Wettability of Coking Coal Based on Molecular Dynamics. Processes 2024, 12, 1271. https://doi.org/10.3390/pr12061271

AMA Style

Liu R, Li S, Ling Y, Zhao Y, Liu W. Research on Mechanism of Surfactant Improving Wettability of Coking Coal Based on Molecular Dynamics. Processes. 2024; 12(6):1271. https://doi.org/10.3390/pr12061271

Chicago/Turabian Style

Liu, Ren, Shilin Li, Yuping Ling, Yuanpei Zhao, and Wei Liu. 2024. "Research on Mechanism of Surfactant Improving Wettability of Coking Coal Based on Molecular Dynamics" Processes 12, no. 6: 1271. https://doi.org/10.3390/pr12061271

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop