Next Article in Journal
Comparison of Pore Structure Characteristics of Shale-Oil and Tight-Oil Reservoirs in the Fengcheng Formation in Mahu Sag
Previous Article in Journal
A Deep Learning Quantile Regression Photovoltaic Power-Forecasting Method under a Priori Knowledge Injection
Previous Article in Special Issue
Biofuels in Aviation: Exploring the Impact of Sustainable Aviation Fuels in Aircraft Engines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Determination of the Diffusion Coefficients of Binary CH4 and C2H6 in a Supercritical CO2 Environment (500–2000 K and 100–1000 atm) by Molecular Dynamics Simulations

by
Chun-Hung Wang
1,2,
K. R. V. Manikantachari (Raghu)
3,
Artëm E. Masunov
2,4 and
Subith S. Vasu
3,*
1
Department of Science, Northland Pioneer College, Little Colorado Campus, Winslow, AZ 86047, USA
2
NanoScience Technology Center, University of Central Florida, Orlando, FL 32826, USA
3
Center for Advanced Turbomachinery and Energy Research (CATER), Department of Mechanical and Aerospace Engineering, University of Central Florida, Orlando, FL 32816, USA
4
School of Modeling, Simulation, and Training, University of Central Florida, Orlando, FL 32826, USA
*
Author to whom correspondence should be addressed.
Energies 2024, 17(16), 4028; https://doi.org/10.3390/en17164028
Submission received: 20 June 2024 / Revised: 3 August 2024 / Accepted: 11 August 2024 / Published: 14 August 2024

Abstract

:
The self-diffusion coefficients of carbonaceous fuels in a supercritical CO2 environment provide transport information that can help us understand the Allam Cycle mechanism at a high pressure of 300 atm. The diffusion coefficients of pure CO2 and binary CO2/CH4 and CO2/C2H6 at high temperatures (500 K~2000 K) and high pressures (100 atm~1000 atm) are determined by molecular dynamics simulations in this study. Increasing the temperature leads to an increase in the diffusion coefficient, and increasing the pressure leads to a decrease in the diffusion coefficients for both methane and ethane. The diffusion coefficient of methane at 300 atm is approximately 0.012 cm2/s at 1000 K and 0.032 cm2/s at 1500 K. The diffusion coefficient of ethane at 300 atm is approximately 0.016 cm2/s at 1000 K and 0.045 cm2/s at 1500 K. The understanding of diffusion coefficients potentially leads to the reduction in fuel consumption and minimization of greenhouse gas emissions in the Allam Cycle.

1. Introduction

Diffusion is a fundamental process driven by the random thermal motion of molecules, which significantly impacts chemical and biological reaction rates [1]. It manifests differently across gas, liquid, and solid phases, and governs the overall rate of these processes. This study explores diffusion within supercritical fluid (SCF), which is a unique state where liquid and gas properties coexist. Such unique properties pose experimental and computational challenges [2]. SCF mass transfer processes hold immense potential in various fields: (i) the selective extraction of valuable components from food and environmental samples; and (ii) design and optimization of SCF reactors with precise control of temperature and pressure parameters to ensure proper mixing for efficient reactions. Supercritical carbon dioxide (sCO2) is a promising tool in materials science and biological applications. Notably, the Allam Power Cycle, a recent breakthrough technology, utilizes sCO2 for thermal energy conversion from carbon fuels [3]. The cycle transforms greenhouse gas CO2 in traditional power plants into a reusable working fluid for spin turboexpanders and then generates electricity. The power generation cycle was developed in the last 10 years and has turned the air pollution problem into a solution [3]. However, further research is necessary to fully unlock its potential. Understanding diffusion is crucial for optimizing several key aspects of the technology of the Allam Cycle. First, diffusion dictates the mixing of fuels, like natural gas, with CO2 at the molecular level. Improved mixing enhances combustion completeness. Second, efficient heat transfer from hot combustion products to CO2 working fluid relies on diffusion within the gas stream. Third, diffusion influences the separation mechanism of CO2 from flue gas after combustion. This study investigates the diffusion of hydrocarbon in sCO2 due to its unique liquid-like and gaseous-like properties, which influences the diffusion coefficients of various hydrocarbons. Understanding the behavior of hydrocarbons in high-temperature and -pressure supercritical environments is vital [4].
The direct-fired sCO2 cycle that utilizes CO2 as the working fluid directly is an alternative to current electricity production. The Allam Cycle is a novel power generation technology and it converts carbon fuels into thermal energy and electricity while capturing the generated CO2 and O2. The cycle involves four stages. First, the cycle begins with the burning of fuels and oxygen. Fuels and pure O2 combust within high-pressure and high-temperature sCO2 in a combustor. The purpose of using pure O2 is to replace air for cleaner combustion and to avoid nitrogen dilution. Second, the hot combustion products, CO2 and water vapor, transfer heat to sCO2 in a heat exchanger. Third, the heated CO2 expands through a turbine and generates electricity. Lastly, cooled-down sCO2 is collected in a compressor to complete the cycle. The relatively low critical temperature (37 °C) and critical pressure (8.4 MPa) of sCO2 make it an attractive choice as an SCF solvent in the Allam Cycle. By elucidating the diffusion coefficients, we aim to gain a deeper understanding of the molecular transport phenomena within this cycle, ultimately paving the way for further power cycle design improvements. The Allam Cycle operates at high pressures (30–300 atm) and high temperatures (500–1150 °C) [3,4].
SCFs offer significant advantages due to their low viscosity and high solute diffusivity, leading to a multitude of industrial applications. The combination of CO2 with hydrocarbons, like methane and ethane, in supercritical environments holds particular interest in various fields. For example, efficient carbon capture and storage are crucial for mitigating climate change. sCO2 can effectively capture CO2 emissions from industrial processes. sCO2 can also be used to selectively separate and purify desired components from mixtures to avoid the usage of hazardous organic solvents. Furthermore, sCO2 is a valuable solvent for extracting materials from various sources. These characteristics show the importance of understanding diffusion properties in the sCO2 environment to optimize these industrial procedures. With its near-perfect CO2 capture capability, the Allam Cycle requires stable diffusion in the combustor. This is the main reason for investigating the diffusion coefficients of methane and ethane in an sCO2 environment to see how stable the diffusion of hydrocarbon is in such gaseous-like and liquid-like conditions. The Allam Cycle operates at a pressure range of 30–300 atm. At 300 atm, CO2 exhibits optimal thermodynamic properties and translates to high thermal efficiency and reduced material selection challenge. In addition, the turbine inlet temperature can reach as high as 1150 °C, which is comparable to modern natural gas combined cycle plants. The broad pressure and temperature windows play significant roles in the Allam Cycle’s performance and design [3,4], and this is the goal of our study.
This paper describes the results of the determination of the binary diffusion coefficients CO2/CH4 and CO2/C2H6 by classical molecular dynamics (MD) simulations. CH4 and C2H6 are prevalent hydrocarbon fuels employed in power cycles. In the past, Stubbs comprehensively reviewed MD and Monte Carlo (MC) simulations of supercritical H2O as well as sCO2 systems in detail, encompassing crucial aspects of the selection of force fields, the size of the simulation boxes, and the pH at different temperatures, pressures, and densities [5]. The thermophysical properties in the sCO2 environment have been summarized in his review article. Several studied have employed various force fields to evaluate diffusion coefficients in systems relevant to this work. Aimoli et al. [6] and Moultos et al. [7] utilized the transferable potentials for phase equilibria (TraPPE) force field to assess the diffusion coefficients of pure CO2, CH4, and CO2-H2O mixtures across a broad range of temperatures (273–623 K) and pressures (0.1–100 MPa). Zhu et al. [8] employed the elementary physical model (EPM2) force field for a Gibbs ensemble MC simulation to determine diffusion coefficients at 304 K. The radial distribution function of CO2 was analyzed to evaluate the validity of their simulations that the first-neighbor C-C distance is around 4.0 Å. Abbaspour and Nameni used a two-body Hartree–Fock dispersion-like potential to determine the self-diffusion coefficients of CO2 and CO2-CH4 mixtures at approximately 300 K [9].
Several studies have investigated diffusion in an sCO2 environment relevant to our research. Guevara-Carrion et al. performed CH4 diffusion experiments with the Taylor dispersion technique and MD simulations [10] at temperatures in the range of 293–333 K and pressures in the range of 9.0–14.7 MPa, which is typical of Allam Cycle conditions. This group determined the self-diffusion, Fick, and Maxwell–Stefan coefficients of CH4 diluted in an sCO2 environment. Feng et al. collected a series of diffusion coefficients of n-hydrocarbon (C1–C14) in near-critical and sCO2 environments (308 K and 323 K, respectively, at 10.5 MPa) from experiments and simulations at an infinite dilution. The ratios of carbon dioxide to hydrocarbon in their simulations are CO2:CH4 = 4000:110 and CO2:C2H6 = 4000:58 to mimic the infinite dilution of hydrocarbon molecules [11]. Furthermore, in recent years, Asadov et al. experimentally studied the diffusion coefficients of CO2-C2H6-heavy oil and CO2-C3H8-heavy oil in an sCO2 environment at temperatures in the range of 320–355 K and pressures in the range of 2–15 MPa [12]. It is important to note that these examples of previous studies focused on conditions below 623 K and 100 MPa.
More MD and MC simulations were applied to study various physical and chemical properties of hydrocarbons in an sCO2 environment. Other MD studies included of the following: (1) the thermodynamic properties of CH4 in an sCO2 environment (CO2:CH4 = 400:100, 323 K at 9.94 MPa), such as potential energy and pressure, mean square force, and torque, were studied by Skarmoutsos et al. [13]; and (2) Gong et al. studied the evaporation mode transition of hydrocarbon fuels in subcritical and supercritical fluids (750–3600 K and 4–36 MPa) to gain insights into air–fuel mixing and combustion processes [14]. Other MC studies included the following: (1) the chemical potential of non-polar hydrocarbon in an sCO2 environment (300–350 K at 10–500 bar) was examined by Chang [15]; and (2) the free energy of the solvation and structural properties of CH4 in an sCO2 environment (304 K at 80 and 200 atm) were studied by Tafazzoli et al. [16] These studies show the importance of providing thermodynamic parameters to understand hydrocarbon reactions in an sCO2 environment.
Here, we focus on the self-diffusion coefficients of CO2/CH4 and CO2/C2H6 mixtures in extreme temperatures (500–2000 K) and pressures (100–1000 atm) that were not easily studied experimentally and/or computationally in the past to mimic the condition of the Allam Cycle. Studying diffusion under such extreme conditions presents significant challenges for both experimental and theoretical approaches. From the perspective of the experimental challenges, specialized equipment is needed to handle such conditions safely and accurately. The high mobility of molecules and potential for chemical reactions at these extremes introduce noise to the measurements. Some materials used for the equipment might be unstable and potentially lead to contamination [3,4,17]. From the perspective of computer simulation challenges, the accuracy of MD simulations relies heavily on the selection of force fields for high-temperature and -pressure systems. In addition, diffusion is a relatively slow process. Simulating realistic timescales at high temperatures and pressures can be computationally prohibitive [18]. To the best of our knowledge, it is the first time extremely high temperatures (~2000 K) and pressures (~1000 atm) have been achieved by computer simulations, and this is systematically discussed under such conditions. Our simulations explore diffusion behavior at high temperatures and pressures relevant to the Allam Cycle. While these conditions might be challenging to achieve experimentally, our simulations provide valuable data for the understanding the system’s behavior at its operating limits.

2. Computational Methods

The initial pure CO2 diffusion coefficients at various temperatures and pressures were determined by ChemKin II Fortran version for comparison [19]. All the molecular dynamics simulations were performed using the LAMMPS program [20]. The force fields of CO2 [21,22], CH4 [23], and C2H6 [23,24] we selected were united-atom TraPPE because of its broad applicability and transferability to carbon dioxide and hydrocarbons. TraPPE has been successfully validated across various diffusion coefficient evaluations involving sCO2 by comparison with available computational CO2 self-diffusion coefficient results [7]. The cutoff radius of Coulomb and Lennard–Jones potentials was set to 14.0 Å. A time step of 1.0 fs was adopted. Simulations were performed under periodic boundary conditions in a cubic box measuring 35.0 × 35.0 × 35.0 Å3 and 40.0 × 40.0 × 40.0 Å3, with a particle–particle particle–mesh calculation in the k-space to better estimate long-range interactions. The initial geometry of all the simulations was generated by Packmol v20.15.0 [25]. A total of 77 CO2 molecules and binary CO2:CH4 = 16:16, 32:32, or 42:42, and CO2:C2H6 = 16:16 or 32:32 molecules were placed in the simulation box. Different numbers of molecules in simulation boxes of different sizes will lead to different pressures. The reason for using a one-to-one ratio of carbon dioxide versus hydrocarbon is to achieve high pressure in the simulations. The selected simulation temperatures are 750, 1000, 1250, 1500, 1750, and 2000 K for pure CO2, and 300, 500, 1000, 1500, and 2000 K for both CO2/CH4 and CO2/C2H6 mixtures. Energy minimization was the first step to stabilize the simulation box, which was generated originally using Packmol software. The criteria for stopping minimization were: a tolerance for energy of 10−5, a tolerance for force of 10−7 kcal/mol/Å, max iterations of the minimizer of 5 million steps, and a max number of force/energy evaluations of 10 million steps. After energy minimization, a constant volume and temperature ensemble (NVT) were applied for the first 5 million steps (5 ns), followed by a constant volume and energy ensemble (NVE) for 5 million steps (5 ns) as the production run. A Nosé–Hoover thermostat was applied for the NVT [26,27]. The damping parameter was 200.0, meaning we relaxed the temperature at a timespan of 200 femtoseconds for the NVT. At least 5 to 10 independent test runs were averaged (with different initial geometries) to reduce fluctuations and then receive statistically reasonable results.
The calculations of the self-diffusion coefficients of the Einstein relation are based upon the mean squared displacement (MSD):
M S D | r t r ( 0 ) | 2 .
The equation describes the rate at which individual molecules move around due to random thermal motion within a medium. The slope of the MSD versus time is proportional to the diffusion coefficient of the diffusing atoms. The displacement of an atom is from its reference position, which is the original position at the time the simulation was started. MSD reflects the average squared distance a molecule travels over time, indicating how far it explores its surroundings. The MSD was collected in the NVE production runs. The unit of diffusion is cm2/s.

3. Results and Discussion

Pure CO2. In order to validate our simulation methods, we compared our 16, 49, 148, 288, 411, and 538 atm MD simulation results for diffusion coefficient Dij to the ideal gas kinetic theory (IGKT, the selected pressures are 1, 49, 99, 147, and 296 atm) predictions by ChemKin II in Figure 1. In the MD simulation, the NVT and NVE are not able to assign the same pressure as the IGKT results. Therefore, the selection of pressures in the MD simulation should be as close to the IGKT pressure as possible. This comparison serves a two-fold purpose. Firstly, it ensures consistency with established principles of the IGKT. Secondly, it provides a reference point for interpreting the behavior observed in the MD simulations. Dij of pure CO2 from the IGKT creates a gas where particles are point masses with negligible interactions. The equation for calculating Dij by the ideal gas kinetic theory is presented as:
D i j = 3 16 2 π k B 3 T 3 m i j P π σ i j 2 Ω 1,1 * ,
where kB is the Boltzmann constant and mij is the reduced molecular mass for the (i,j) species pair:
m i j = m i m j m i + m j
Σij is the reduced collision diameter, and Ω(1,1)* is the collision integral based on the Stockmayer potential [28]. Figure 1 shows the similarities between our MD simulation results and fitted ideal gas theory results, except at a low pressure of 16 atm at 750 K. The increase in temperature leads to the increase in Dij. A higher temperature leads to faster molecular motion and a higher diffusion coefficient. The increased density in SCFs due to the pressure increase results in more frequent collisions and more efficient shuffling between molecules. At a high pressure (above 100 atm), the MD simulation results are closer to those of the ideal gas kinetic theory (within a factor of 10 of each other). At a very high pressure, potentially, sCO2 can start to resemble a solid and diffusion slows down. This causes the slope of Dij to decline as the temperature increases. An underestimate of Dij at a lower temperature (750 K) aligns with our previous study of the chemical kinetics of combustion reactions in an sCO2 environment well [29,30]. Table 1 shows the average diffusion coefficients of pure CO2 under various temperatures and pressures by MD simulations and how spread out the data are. The results support the validity of using the chosen simulation conditions, such as force field TraPPE, to account for intermolecular interactions for the binary systems containing carbon dioxide and hydrocarbon, particularly at higher pressures above 50 atm, for accurate predictions compared to ideal gas assumptions.
Binary CO2 and CH4 mixtures. To understand the combustion of methane in an sCO2 environment, it is necessary to study the diffusion of methane in CO2/CH4 mixtures. Figure 2 presents the diffusion coefficient of binary mixtures at various temperatures and pressures. The general trend of diffusion coefficient Dij is that, as the temperature increases, the molecules in the mixture gain kinetic energy and result in the increase in Dij. Both CO2 and CH4 move faster and collide with each other more frequently. A higher temperature provides enough thermal energy for molecules to diffuse. As the pressure increases at a constant temperature, Dij decreases. A higher pressure depresses the diffusion of molecules as the mean free path of the fluid decreases and starts to resemble a solid so that a more frequent collision between molecules occurs and then hinders diffusion. Our simulation results reach as high as 930 atm at 2000 K. Dij at near 300 atm is 0.012 at 1000 K and 0.032 cm2/s at 1500 K. Interestingly, compared with the results of Dij at 300 K from different groups, our results are ten-times higher than those of Guevara-Carrion et al.’s result of ~0.0002 cm2/s at 300 K, 9 MPa [10], and Feng et al.’s result of 0.000272 cm2/s at 299 K, 10.5 MPa [11]. It is important to remember that Dij at low temperatures (300 and 500 K) in our simulation results is systematically tenfold higher than those the other groups reported. However, when we compare the diffusion coefficients in terms of absolute values, they are quite similar. These differences may be from the relatively low mole fraction of CH4 in their study, while our study has a 1:1 carbon dioxide and hydrocarbon ratio. Such a discrepancy at lower temperatures aligns with our previous study of the chemical kinetics of combustion reactions in an sCO2 environment well [29,30]. We are currently looking into the additional reasons behind this discrepancy to better understand the implications for our research.
Binary CO2 and C2H6 mixtures. To understand the combustion of ethane in an sCO2 environment, it is necessary to study the diffusion of ethane in CO2/C2H6 mixtures. Figure 3 presents the trend of the diffusion coefficient (Dij) of CO2/C2H6 that is similar to that of CO2/CH4 mixtures. We successfully obtained Dij for all studied temperatures (500, 1000, 1500, and 2000 K), except 300 K and pressures below 600 atm. As the temperature increases, Dij increases as a higher thermal energy for molecules. As the pressure increases at a constant temperature, Dij decreases as more frequent collisions hinder diffusion. Our simulation can reach as high as 560 atm at 2000 K. Dij at almost 300 atm is 0.016 cm2/s at 1000 K and 0.045 cm2/s at 1500 K. Our results can be compared with Feng et al.’s results, 0.000597 cm2/s at 323 K, 10.5 MPa as well [11], though we are not able to achieve the simulation condition of 300 K at a mole fraction of 0.50 for C2H6.
Overall, compared with the CO2/CH4 and CO2/C2H6 results in Figure 2 and Figure 3, respectively, the diffusion coefficient is inversely proportional to the molar volume of the solute and decreases quickly with the increasing carbon chain for short-chain n-alkanes. A smaller molar volume indicates a smaller molecule that experiences less friction as it moves through a solvent molecule. In general, both methane and ethane will experience similar increases in Dij with the increasing temperature. The pressure influences the Dij values of methane and ethane differently due to their subtle size difference. Table 2 shows the value of the diffusion coefficients of CO2/CH4 and CO2/C2H6 under various temperatures and pressures. It is expected that the diffusion coefficient decreases as the mass and size of the molecule increase, while the difference is limited. This table also shows, in general, a greater fluctuation in Dij as the pressure reduces. Feng et al. showed the limited difference of Dij by experiments and MD simulations from CH4 to C14H30 at 10.5 MPa [11]. For example, Dij ranges from 0.000691 to 0.000220 cm2/s at 323 K. Our simulations show the importance of the diffusion coefficients of binary CO2/CH4 and CO2/C2H6 in an sCO2 environment at high temperatures and pressures, which are challenging to achieve experimentally [2].
Table 3 shows the fitted equations under various temperatures in the range of 300–2000 K. The temperature is around 1100 °C (1423 K) and the pressure is around 300 bar (296 atm) of the Allam Power Cycle. The fitted Dij for methane at 1500 K is 0.038 cm2/s and for ethane at 1500 K, it is 0.045 cm2/s.

4. Conclusions

We determined the self-diffusion coefficients of pure CO2 and simplest hydrocarbons in binary CO2/CH4 and CO2/C2H6 mixtures at temperatures in the range of 300–2000 K and pressures in the range of 10–1000 atm. For the comparison to other experimental and computational results, we tested lower temperatures (as low as ~300 K) and pressures (as low as ~10 atm) as well. Our simulations at higher temperatures (as high as ~2000 K) and pressures (as high as ~900 atm) provide the data to better understand the Allam Power Cycle at 300 atm and make improvements in the future. The diffusion coefficients of methane at 300 atm are approximately 0.012 cm2/s at 1000 K and 0.032 cm2/s at 1500 K. The diffusion coefficients of ethane at 300 atm are approximately 0.016 cm2/s at 1000 K and 0.045 cm2/s at 1500 K. Our simulation results are far from a perfect fit and still need improvements, especially at lower temperatures and pressures. Nonetheless, this study provides valuable information for the improvement of the Allam Cycle in the future. The technology of the Allam Cycle is still under development. Besides the understanding of diffusion coefficients, the Allam Cycle requires further research and experiments to improve its overall performance and effectiveness to achieve commercial viability.

Author Contributions

Design of molecular dynamics simulations: C.-H.W.; ideal gas model and concept of equation of state: K.R.V.M.; simulation planning: A.E.M.; funding acquisition and manuscript review: S.S.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work was mainly supported by the Department of Energy (DOE) under the grant number DE-FE0025260. Some aspects of the research were made possible with support from UCF.

Data Availability Statement

In the research and the article, only materials from publicly available sources of information are used. The simulation results obtained with LAMMPS software 2 August 2023 version are presented.

Acknowledgments

This study was prepared as an account of work sponsored by the United States Government agency. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed or represented that its use would not infringe privately owned rights. Referring to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of the authors expressed here do not necessarily state or reflect those of the United States Government or any agency thereof.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cussler, E.L. Diffusion: Mass Transfer in Fluid Systems; Cambridge University Press: Cambridge, UK, 2009. [Google Scholar]
  2. Medina, I. Determination of diffusion coefficients for supercritical fluids. J. Chromatogr. A 2012, 1250, 124–140. [Google Scholar] [CrossRef] [PubMed]
  3. Allam, R.; Martin, S.; Forrest, B.; Fetvedt, J.; Lu, X.; Freed, D.; Brown, G.W., Jr.; Sasaki, T.; Itoh, M.; Manning, J. Demonstration of the Allam Cycle: An update on the development status of a high efficiency supercritical carbon dioxide power process employing full carbon capture. Energy Procedia 2017, 114, 5948–5966. [Google Scholar] [CrossRef]
  4. Chan, W.; Morosuk, T.; Li, X.; Li, H. Allam cycle: Review of research and development. Energy Convers. Manag. 2023, 294, 117607. [Google Scholar] [CrossRef]
  5. Stubbs, J.M. Molecular simulations of supercritical fluid systems. J. Supercrit. Fluids 2016, 108, 104–122. [Google Scholar] [CrossRef]
  6. Aimoli, C.; Maginn, E.; Abreu, C. Transport properties of carbon dioxide and methane from molecular dynamics simulations. J. Chem. Phys. 2014, 141, 134101. [Google Scholar] [CrossRef] [PubMed]
  7. Moultos, O.A.; Orozco, G.A.; Tsimpanogiannis, I.N.; Panagiotopoulos, A.Z. Atomistic molecular dynamics simulations of H2O diffusivity in liquid and supercritical CO2. Mol. Phys. 2015, 113, 2805–2814. [Google Scholar] [CrossRef]
  8. Zhu, A.; Zhang, X.; Liu, Q.; Zhang, Q. A fully flexible potential model for carbon dioxide. Chin. J. Chem. Eng. 2009, 17, 268–272. [Google Scholar] [CrossRef]
  9. Abbaspour, M.; Nameni, E. Many-body effects in some thermodynamic properties of supercritical CO2, CO2-Ar, and CO2-CH4 using HFD-like potentials from molecular dynamics simulation. J. Supercrit. Fluids 2013, 74, 61–69. [Google Scholar] [CrossRef]
  10. Guevara-Carrion, G.; Ancherbak, S.; Mialdun, A.; Vrabec, J.; Shevtsova, V. Diffusion of methane in supercritical carbon dioxide across the Widom line. Sci. Rep. 2019, 9, 8466. [Google Scholar] [CrossRef] [PubMed]
  11. Feng, H.; Gao, W.; Sun, Z.; Lei, B.; Li, G.; Chen, L. Molecular dynamics simulation of diffusion and structure of some n-alkanes in near critical and supercritical carbon dioxide at infinite dilution. J. Phys. Chem. B 2013, 117, 12525–12534. [Google Scholar] [CrossRef] [PubMed]
  12. Asadov, M.M.; Ramazanova, E.E.; Aliyev, E.N. Solubility, diffusion of components, and equilibrium in CO2-ethane-heavy oil and CO2-propane-heavy oil system. Russ. J. Phys. Chem. A 2021, 95, 868–874. [Google Scholar] [CrossRef]
  13. Skarmoutsos, I.; Samios, J. Local intermolecular structure and dynamics in binary supercritical solutions. A molecular dynamics simulation study of methane in carbon dioxide. J. Mol. Liq. 2006, 125, 181–186. [Google Scholar] [CrossRef]
  14. Gong, Y.; Ma, X.; Luo, K.H.; Xu, H.; Shuai, S. A molecular dynamics study of evaporation mode transition of hydrocarbon fuels under supercritical conditions. Combust. Flame 2022, 246, 112397. [Google Scholar] [CrossRef]
  15. Chang, J. Expanded ensemble Monte Carlo simulations for the chemical potentials of supercritical carbon dioxide and hydrocarbon studies. Korean J. Chem. Eng. 2011, 28, 597–601. [Google Scholar] [CrossRef]
  16. Tafazzoli, M.; Khanlarkhani, A. Investigation of the enhanced solubility of fluorinated methanes in CO2 by Monte Carlo simulation: Absolute free energy of solvation and structural properties of solution. J. Supercrit. Fluids 2007, 40, 40–49. [Google Scholar] [CrossRef]
  17. Fernandes, D.; Wang, S.; Xu, Q.; Buss, R.; Chen, D. Process and Carbon Footprint Analyses of the Allam Cycle Power Plant Integrated with an Air Separation Unit. Clean Technol. 2019, 1, 325–340. [Google Scholar] [CrossRef]
  18. Frenkel, D.; Smit, B. Understanding Molecular Simulation: From Algorithms to Applications, 2nd ed.; Academic Press: Cambridge, MA, USA, 2002. [Google Scholar]
  19. Kee, R.J.; Rupley, F.M.; Miller, J.A. Chemkin-II: A Fortran Chemical Kinetics Package for the Analysis of Gas-Phase Chemical Kinetics. United States: N. p. 1989. Web. Available online: https://www.osti.gov/biblio/5681118 (accessed on 10 August 2024).
  20. Thompson, A.P.; Aktulga, H.M.; Berger, R.; Bolintineanu, D.S.; Brown, W.M.; Crozier, P.S.; In’t Veld, P.J.; Kohlmeyer, A.; Moore, S.G.; Nguyen, T.D.; et al. LAMMPS–A flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Comput. Phys. Commun. 2022, 271, 108171. [Google Scholar] [CrossRef]
  21. Potoff, J.J.; Siepmann, J.I. Vapor-liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogen. AIChE J. 2001, 47, 1676–1682. [Google Scholar] [CrossRef]
  22. Eggimann, B.L.; Sun, Y.; DeJaco, R.F.; Singh, R.; Ahsan, M.; Josephson, T.R.; Siepmann, J.I. Assessing the quality of molecular simulations for vapor-liquid equilibria: An analysis of the TraPPE database. J. Chem. Eng. Data 2020, 65, 1330–1334. [Google Scholar] [CrossRef]
  23. Martin, M.G.; Siepmann, J.I. Transferable potentials for phase equilibria. 1. United-atom description on n-alkanes. J. Phys. Chem. B 1998, 102, 2569–2577. [Google Scholar] [CrossRef]
  24. Shah, M.S.; Tsapatsis, M.; Siepmann, J.I. Transferable potentials for phase equilibria. Improved united-atom description of ethane and ethylene. AIChE J. 2017, 63, 5098–5110. [Google Scholar] [CrossRef]
  25. Martinez, L.; Andrade, R.; Birgin, E.G.; Martinez, J.M. Packmol: A package for building initial configurations for molecular dynamics simulations. J. Comput. Chem. 2009, 30, 2157–2164. [Google Scholar] [CrossRef] [PubMed]
  26. Nosé, S. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81, 511–519. [Google Scholar] [CrossRef]
  27. Hoover, W.G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 1985, 31, 1695–1697. [Google Scholar] [CrossRef] [PubMed]
  28. Monchick, L.; Mason, E.A. Transport properties of polar gases. J. Chem. Phys. 1961, 35, 1676–1697. [Google Scholar] [CrossRef]
  29. Wang, C.-H.; Panteleev, S.V.; Masunov, A.E.; Allison, T.C.; Chang, S.; Lim, C.; Jin, Y.; Vasu, S.S. Molecular dynamics of combustion reactions in supercritical carbon dioxide. Part 5: Computational study of ethane dissociation and recombination reactions C2H6 ⇌ CH3 + CH3. J. Phys. Chem. A 2019, 123, 4776–4784. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, C.-H.; Masunov, A.E.; Allison, T.C.; Chang, S.; Lim, C.; Jin, Y.; Vasu, S.S. Molecular dynamics of combustion reactions in supercritical carbon dioxide. 6. Computational kinetics of reactions between hydrogen atom and oxygen molecule H + O2 ⇌ HO + O and H + O2 ⇌ HO2. J. Phys. Chem. A 2019, 123, 10772–10781. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Self-diffusion coefficients (Dij) of CO2 computed by the ideal gas kinetic theory (1, 49, 99, 197, and 296 atm) and MD simulations (16, 49, 148, 288, 411, and 538 atm).
Figure 1. Self-diffusion coefficients (Dij) of CO2 computed by the ideal gas kinetic theory (1, 49, 99, 197, and 296 atm) and MD simulations (16, 49, 148, 288, 411, and 538 atm).
Energies 17 04028 g001
Figure 2. Self-diffusion coefficients (Dij) of CH4 in binary CO2/CH4 mixtures at various temperatures (300, 500, 1000, 1500, and 2000 K) and pressures below 1000 atm.
Figure 2. Self-diffusion coefficients (Dij) of CH4 in binary CO2/CH4 mixtures at various temperatures (300, 500, 1000, 1500, and 2000 K) and pressures below 1000 atm.
Energies 17 04028 g002
Figure 3. Self-diffusion coefficients (Dij) of C2H6 in binary CO2/C2H6 mixtures at various temperatures (500, 1000, 1500, and 2000 K) and pressures below 600 atm.
Figure 3. Self-diffusion coefficients (Dij) of C2H6 in binary CO2/C2H6 mixtures at various temperatures (500, 1000, 1500, and 2000 K) and pressures below 600 atm.
Energies 17 04028 g003
Table 1. Diffusion coefficients (Dij) of CO2 under various temperatures (748–2040 K) and pressures (16–538 atm).
Table 1. Diffusion coefficients (Dij) of CO2 under various temperatures (748–2040 K) and pressures (16–538 atm).
T (K)P (atm)Dij (10−4 cm2/s)
74816 ± 32 ± 1
101249 ± 3258 ± 24
1243148 ± 7128 ± 11
1530288 ± 14171 ± 9
1783411 ± 23215 ± 13
2040538 ± 21245 ± 16
Table 2. Diffusion coefficients (Dij) of hydrocarbon in CO2/CH4 and CO2/C2H6 mixtures under various temperatures (300–2000 K) and pressures (10–1000 atm).
Table 2. Diffusion coefficients (Dij) of hydrocarbon in CO2/CH4 and CO2/C2H6 mixtures under various temperatures (300–2000 K) and pressures (10–1000 atm).
CO2/CH4CO2/C2H6
T (K)P (atm)Dij (10−4 cm2/s)P (atm)Dij (10−4 cm2/s)
30012 ± 4185 ± 29
20 ± 11111 ± 24
31 ± 1152 ± 8
104 ± 3832 ± 3
50013 ± 7283 ± 5841 ± 3404 ± 54
42 ± 11113 ± 2774 ± 5192 ± 34
94 ± 469 ± 995 ± 14124 ± 18
121 ± 2156 ± 17
100091 ± 7356 ± 4073 ± 5653 ± 88
140 ± 9229 ± 20179 ± 14266 ± 20
238 ± 18129 ± 21266 ± 16163 ± 15
274 ± 20125 ± 16490 ± 31105 ± 11
341 ± 31119 ± 15
1500232 ± 17484 ± 69144 ± 12884 ± 146
351 ± 27330 ± 51304 ± 22444 ± 43
567 ± 54202 ± 16384 ± 29376 ± 37
726 ± 74167 ± 20557 ± 37242 ± 19
2000371 ± 32624 ± 68275 ± 14928 ± 109
510 ± 54397 ± 53564 ± 25490 ± 31
729 ± 51353 ± 34
936 ± 104276 ± 45
Table 3. Fitted equations of Dij (y) of binary CO2/CH4 and CO2/C2H6 mixtures under various temperatures (300–2000 K) and pressures (x).
Table 3. Fitted equations of Dij (y) of binary CO2/CH4 and CO2/C2H6 mixtures under various temperatures (300–2000 K) and pressures (x).
CO2/CH4CO2/C2H6
T (K)Fitted EquationR2 ValueFitted EquationR2 Value
300y = 0.1200x−0.8130.9487
500y = 0.1786x−0.7230.9979y = 7.3009x−1.3920.9979
1000y = 1.8117x−0.8820.9860y = 4.1986x−0.9770.9981
1500y = 8.5061x−0.9490.9993y = 9.5610x−0.9400.9978
2000y = 7.3284x−0.8170.9339y = 13.7450x−0.8901.0000
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

Wang, C.-H.; Manikantachari, K.R.V.; Masunov, A.E.; Vasu, S.S. Determination of the Diffusion Coefficients of Binary CH4 and C2H6 in a Supercritical CO2 Environment (500–2000 K and 100–1000 atm) by Molecular Dynamics Simulations. Energies 2024, 17, 4028. https://doi.org/10.3390/en17164028

AMA Style

Wang C-H, Manikantachari KRV, Masunov AE, Vasu SS. Determination of the Diffusion Coefficients of Binary CH4 and C2H6 in a Supercritical CO2 Environment (500–2000 K and 100–1000 atm) by Molecular Dynamics Simulations. Energies. 2024; 17(16):4028. https://doi.org/10.3390/en17164028

Chicago/Turabian Style

Wang, Chun-Hung, K. R. V. Manikantachari (Raghu), Artëm E. Masunov, and Subith S. Vasu. 2024. "Determination of the Diffusion Coefficients of Binary CH4 and C2H6 in a Supercritical CO2 Environment (500–2000 K and 100–1000 atm) by Molecular Dynamics Simulations" Energies 17, no. 16: 4028. https://doi.org/10.3390/en17164028

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