Eleven NanoHUB Simulation Tools Using RASPA Software To Demonstrate Classical Atomistic Simulations of Fluids and Nanoporous MaterialsClick to copy article linkArticle link copied!
- Julian C. UmehJulian C. UmehChemical and Materials Engineering, New Mexico State University, Jett Hall 268, Las Cruces, New Mexico88003-3805, United StatesMore by Julian C. Umeh
- Thomas A. Manz*Thomas A. Manz*Email: [email protected]Chemical and Materials Engineering, New Mexico State University, Jett Hall 268, Las Cruces, New Mexico88003-3805, United StatesMore by Thomas A. Manz
Abstract
Eleven interactive simulation tools were created on nanoHUB to help users learn how to perform classical atomistic simulations. These tools enable users to perform classical Monte Carlo and molecular dynamics simulations using RASPA software. These tools use comparatively small numbers of production cycles to keep the runtimes short, so that users will not be discouraged by long wait times to see results. Here, we show that these tools produce results of sufficient accuracy and reproducibility for learning purposes. The 11 tools developed were as follows: (1) calculation of the self-diffusion constant of gas molecules in metal–organic frameworks (MOFs), (2) gas adsorption in MOFs using the grand canonical ensemble, (3) Henry’s coefficient calculator for gas molecules in MOFs and a zeolite, (4) adsorption of a gas mixture in a MOF, (5) self-diffusion of a gas mixture in a MOF, (6) void fraction calculation for several MOFs and zeolites, (7) surface area calculation for several MOFs and zeolites, (8) calculation of radial distribution function and self-diffusion constant for several pure gases, (9) energy distribution of adsorption sites using a probe molecule in MOFs, (10) molecular dynamics simulation of pure fluids in the NPT ensemble, and (11) gas adsorption in MOFs using the Gibbs ensemble.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
1. Introduction
2. Methods
2.1. Classical Forcefields Used
Figure 1
Figure 1. Lennard-Jones potential for argon. The energy unit is in Kelvin (used by RASPA); this can be multiplied by the Boltzmann constant to convert to Joules per particle.
system | forcefield | flexible | ref | tools used in |
---|---|---|---|---|
Ar | Lennard-Jones | N | a | 1, 4, 5, 11 |
H2 | (GenericMOFs)b | N | (21) | 1, 2, 9, 11 |
N2 | TraPPE | N | (20) | 1, 2, 8, 9, 11 |
O2 | TraPPE | N | (22) | 2 |
CO2 | TraPPE | N | (20) | 1, 2, 8, 9, 10, 11 |
CH4 | TraPPE-UA | N | (19) | 1, 2, 4, 5, 7–11 |
C2H6 | TraPPE-UA | Y | (19) | 8, 9, 10 |
C3H8 | TraPPE-UA | Y | (19) | 8, 10 |
C4H10 | TraPPE-UA | Y | (19) | 8 |
n-pentane | TraPPE-UA | Y | (19) | 3 |
n-hexane | TraPPE-UA | Y | (19) | 3 |
n-heptane | TraPPE-UA | Y | (19) | 3 |
n-octane | TraPPE-UA | Y | (19) | 3 |
n-nonane | TraPPE-UA | Y | (19) | 3 |
IRMOF-1 | Dubbeldam | Nc | (16) | 1–7, 9, 11 |
IRMOF-2 | (GenericMOFs)d | N | (9,23) | 7 |
IRMOF-3 | (GenericMOFs)d | N | (9,23) | 7 |
IRMOF-12 | (GenericMOFs)d | N | (9,23) | 7 |
IRMOF-16 | Dubbeldam | Nc | (16) | 1–3, 6, 7, 9, 11 |
ITQ-1 | GenericZeolites | N | e | 6, 7 |
ITQ-3 | GenericZeolites | N | e | 6, 7 |
ITQ-7 | GenericZeolites | N | e | 6, 7 |
ITQ-12 | GenericZeolites | N | e | 6, 7 |
ITQ-29 | GenericZeolites | N | e | 6, 7 |
KFI | GenericZeolites | N | e | 6, 7 |
LTA4A | GenericZeolites | N | e | 6, 7 |
LTA5A | GenericZeolites | N | e | 6, 7 |
MFI_SI | GenericZeolites | N | e | 3 |
Lennard-Jones parameters for argon were set to σ = 3.34 Å and ε = 119.8 K to reproduce the well depth and rmin = 21/6 σ value for the dimer curve.
These parameters for the H2 molecule are part of the GenericMOFs forcefield collection in RASPA.
This forcefield included flexibility but it was not used.
IRMOF-2, IRMOF-3, and IRMOF-12 used generic atom types and parameters from the UFF and Dreiding forcefields, as compiled within the GenericMOFs forcefield collection in RASPA.
The GenericZeolites forcefield in RASPA uses Lennard-Jones parameter values collected from different sources: (a) for Si and O atoms, the GenericZeolites forcefield uses TraPPE-zeo (ref (24)) Lennard-Jones parameter values, (b) for Al atoms, the GenericZeolites forcefield used the same Lennard-Jones parameter values as for Si, (c) although these atoms are not present in any of the structures here, for Na and Ca atoms, the GenericZeolites forcefield uses UFF (ref (9)) Lennard-Jones parameter values.
2.2. Monte Carlo, Molecular Dynamics, and Widom Insertion Simulations in RASPA Software
The simulation type (e.g., Monte Carlo or Molecular Dynamics)
The number of initialization, equilibration, and production cycles
The molecules and/or framework that make up the system to be simulated
The forcefields and mixing rules to be used
If a framework is included, whether framework flexibility is allowed or the framework is to be held rigid
For Monte Carlo calculations, the different kinds of trial moves and their probabilities
Keywords to specify various quantities to be computed and printed
Depending on the simulation type, the external temperature and/or pressure may need to be specified
The cutoff radius for interactions between atoms
Translation: This move stochastically displaces (translates) a random molecule. The internal configuration of the molecule remains unchanged.
Rotation: This move randomly selects a molecule and then randomly rotates it about the chosen starting bead. The internal configuration of the molecule remains unchanged. When a molecule is modeled as a sphere, then rotation is not needed. For example, rotation is not needed when running a Monte Carlo simulation of methane using the united-atom model, which models methane as a sphere.
Swap: This move randomly attempts to insert or delete a random molecule from the system. An insertion is attempted 50% of the time, and a deletion is attempted the other 50% of the time.
Reinsertion: “A full reinsertion move for the current component.” (15) “The ‘reinsertion’ move removes a randomly selected molecule and reinserts it at a random position. For rigid molecules, it uses orientational biasing, and for chains, the molecule is fully regrown (the internal configuration is modified).” (25) “Multiple first beads are chosen, and one of these is selected according to its Boltzmann weight. The remaining part of the molecule is grown using biasing. This move is very useful, and often necessary, to change the internal configuration of flexible molecules.” (15)
Volume change: This move isotopically changes the volume of the simulation box.
Gibbs ensemble calculations use two boxes. (28,29) In these computations, a “Gibbs swap” moves a molecule from one box to the other; a new internal configuration is regrown for the molecule using CBMC in the second box.
Figure 2
Figure 2. Monte Carlo and molecular dynamics cycles.
Microcanonical ensemble (NVE)─This statistical ensemble holds the number of particles N, the volume V, and the energy E constant during the course of the simulation.
Canonical ensemble (NVT)─This ensemble holds the number of particles N, the volume V, and the average temperature T constant.
Grand Canonical ensemble (μVT)─This ensemble holds the chemical potential μ, the volume V, and the average temperature T constant.
Isobaric–isothermal ensemble (NPT)─This ensemble holds the number of particles N, the average pressure P, and the average temperature T constant.
Isoenthalpic–isobaric ensemble (NPH)─This ensemble holds the number of particles N, the average pressure P, and the enthalpy H constant.
2.3. Creating NanoHUB Tools That Run RASPA Simulations
Figure 3
Figure 3. Flow diagram illustrating the process to create a new nanoHUB tool.
Figure 4
Figure 4. NanoHUB workspace.
Figure 5
Figure 5. Rappture builder GUI.
Figure 6
Figure 6. TortoiseSVN repository browser.
2.4. Computing a Molecule’s Self-Diffusivity
Figure 7
Figure 7. Left panel: A log–log plot of mean-squared displacement versus time that shows the separate ballistic, transition, and diffusive regimes. Right panel: A linear–linear plot of mean-squared displacement versus time using the starting and ending times selected by the algorithm that extracts the self-diffusion constant Ds from a data subset within the diffusive regime.
1. | Loop over j = 1, 2 to (N-20), where N is the total number of datapoints. (The 20 points at the end are reserved to ensure that there will always be a large enough curve segment remaining for us to perform an adequate fit to extract the Ds.) During each iteration of this loop, if avg_slopej >1.25, then set k = j. | ||||
2. | Ds is computed as |
initialization cycles | equilibration cycles | production cycles | self-diffusion (Å2/picosecond) | |
---|---|---|---|---|
nanoHUB | 5000 | 5000 | 30,000 | 11 ± 2 |
SDSC | 50,000 | 50,000 | 250,000 | 11.1 ± 0.6 |
3. Simulation Tool Descriptions
3.1. Gas Diffusion Coefficient in MOFs (Tool 1)
Figure 8
Figure 8. Screenshots of Tool 1 for calculating the self-diffusion constant of a gas in a MOF.
3.2. Gas Adsorption Calculator (Tool 2)
Figure 9
Figure 9. Screenshots of Tool 2 for calculating the absolute and excess adsorption of a gas in a MOF.
3.3. Henry’s Coefficients Simulator (Tool 3)
Figure 10
Figure 10. Screenshot of Tool 3 for calculating the Henry’s coefficients of n-alkanes in porous materials.
3.4. Mixed Gas Adsorption Calculator (Tool 4)
Figure 11
Figure 11. Screenshots of Tool 4 for calculating adsorption of a gas mixture in a porous material.
3.5. Mixed Gas Diffusion Calculator (Tool 5)
Figure 12
Figure 12. Screenshots of Tool 5 for calculating diffusion of a gas mixture in a porous material.
3.6. Void Fraction Calculator (Tool 6)
Figure 13
Figure 13. Screenshot of Tool 6 for calculating the void fraction of a porous material.
3.7. Surface Area Calculator (Tool 7)
Figure 14
Figure 14. Screenshot of Tool 7 for calculating the surface area of a porous material.
3.8. Radial Distribution Function Calculator (Tool 8)
Figure 15
Figure 15. Screenshots of Tool 8 for calculating the radial distribution function of a gas.
3.9. Adsorption Energy Calculator (Tool 9)
Figure 16
Figure 16. Screenshot of Tool 9 for calculating the adsorption energy histogram.
3.10. NPT Simulator (Tool 10)
Figure 17
Figure 17. Screenshot of Tool 10 for calculating the density and total energy of a gas in the NPT ensemble.
3.11. Gibbs Adsorption Simulator (Tool 11)
Figure 18
Figure 18. Screenshots of Tool 11 for calculating gas adsorption in a MOF using the Gibbs ensemble.
4. Results and Discussion
4.1. Convergence Tests
tool 2 (grand canonical ensemble) | tool 11 (Gibbs ensemble) | ||||
---|---|---|---|---|---|
initialization cycles | production cycles | absolute adsorption (mol/kg framework) | excess adsorption (mol/kg framework) | absolute adsorption (mol/kg framework) | |
nanoHUB | 5000 | 10,000 | 8.7 ± 0.2 | 8.0 ± 0.2 | 8.5 ± 0.2 |
SDSC | 25,000 | 50,000 | 8.56 ± 0.05 | 7.91 ± 0.05 | 8.85 ± 0.01 |
For comparison, the last column lists the absolute adsorption from Tool 11 in the Gibbs ensemble.
production cycles | C5 | C6 | C7 | C8 | C9 | |
---|---|---|---|---|---|---|
nanoHUB | 5000 | 30.2 ± 0.7 | 60 ± 1 | 123 ± 1 | 243 ± 10 | 446 ± 13 |
SDSC | 25,000 | 28.8 ± 0.2 | 57.83 ± 0.04 | 116 ± 1 | 223 ± 6 | 410 ± 9 |
Units are 10–7 mol/(kg Pa).
initialization cycles | production cycles | argon absolute adsorption (mol/kg framework) | argon excess adsorption (mol/kg framework) | methane absolute adsorption (mol/kg framework) | methane excess adsorption (mol/kg framework) | |
---|---|---|---|---|---|---|
nanoHUB | 5000 | 10,000 | 0.366 ± 0.002 | 0.203 ± 0.002 | 0.8849 ± 0.0013 | 0.6414 ± 0.0013 |
SDSC | 25,000 | 50,000 | 0.3654 ± 0.0005 | 0.203 ± 0.001 | 0.883 ± 0.002 | 0.640 ± 0.002 |
initialization cycles | equilibration cycles | production cycles | argon self-diffusion (Å2/picosecond) | methane self-diffusion (Å2/picosecond) | |
---|---|---|---|---|---|
nanoHUB | 50,000 | 50,000 | 100,000 | 2.4 ± 0.9 | 3.7 ± 0.3 |
SDSC | 50,000 | 50,000 | 250,000 | 3.4 ± 0.5 | 3.2 ± 0.5 |
production cycles | void fraction | surface area (m2/g) | |
---|---|---|---|
nanoHUB | 2000 | 0.4623 ± 0.0013 | 758.9 ± 0.2 |
SDSC | 10,000 | 0.4634 ± 0.0003 | 759.26 ± 0.06 |
initialization cycles | equilibration cycles | production cycle | self-diffusion (Å2/ps) | |
---|---|---|---|---|
nanoHUB | 50,000 | 50,000 | 100,000 | 14.4 ± 0.2 |
SDSC | 50,000 | 50,000 | 250,000 | 14.8 ± 0.7 |
initialization cycles | production cycles | adsorption energy (K) | |
---|---|---|---|
nanoHUB | 5000 | 10,000 | –1173 ± 2 |
SDSC | 25,000 | 50,000 | –1168 ± 2 |
initialization cycles | equilibration cycles | production cycle | density (kg/m3) | system energy (K) | |
---|---|---|---|---|---|
nanoHUB | 5000 | 5000 | 10,000 | 5.11 ± 0.14 | –428 ± 20 |
SDSC | 25,000 | 25,000 | 50,000 | 5.01 ± 0.07 | –447 ± 4 |
tool | run time (min) | tool | run time (min) |
---|---|---|---|
1 | 160 ± 30 | 7 | 120.7 ± 0.2 |
2 | 24.6 ± 0.7 | 8 | 8 ± 2 |
3 | 32 ± 1 | 9 | 6.5 ± 0.4 |
4 | 33.0 ± 0.2 | 10 | 4.77 ± 0.01 |
5 | 153 ± 2 | 11 | 28.0 ± 0.5 |
6 | 0.676 ± 0.008 |
The average and standard deviation of three runs is listed.
4.2. Comparing Simulations to Experiments
material | area (m2/g) Langmuirb | area (m2/g) BETb | area (m2/g) (N2 kinetic diam)b | area (m2/g) (sigma) | area (m2/g) (rmin) |
---|---|---|---|---|---|
IRMOF-2 | 2544 | 1722 | 2780 | 2684.4 ± 1.0 | 2349.4 ± 0.5 |
IRMOF-3 | 3062 | 2446 | 3613 | 3296.4 ± 0.9 | 2834.0 ± 1.3 |
Sigma and rmin are surface areas calculated using two different probe distances. The calculations used 2000 production cycles.
From ref (44).
Figure 19
5. Conclusions
Acknowledgments
National Science Foundation (NSF) CAREER Award DMR-1555376 provided financial support. Supercomputing resources were provided by the Extreme Science and Engineering Discovery Environment (XSEDE). (48) XSEDE is funded by NSF Grant ACI-1548562. XSEDE project Grant TG-CTS100027 provided allocations on the Comet and Expanse clusters at the San Diego Supercomputing Center (SDSC). Special thanks to the nanoHUB support team for answering our many questions about setting up tools and for their technical assistance.
References
This article references 48 other publications.
- 1Madhavan, K.; Zentner, L.; Farnsworth, V.; Shivarajapura, S.; Zentner, M.; Denny, N.; Klimeck, G. nanoHUB.org: cloud-based services for nanoscale modeling, simulation, and education. Nanotechnol. Rev. 2013, 2, 107– 117, DOI: 10.1515/ntrev-2012-0043Google ScholarThere is no corresponding record for this reference.
- 2Magana, A. J.; Brophy, S. P.; Bodner, G. M. An exploratory study of engineering and science students’ perceptions of nanoHUB.org simulations. Int. J. Eng. Educ. 2012, 28, 1019– 1032Google ScholarThere is no corresponding record for this reference.
- 3Nanohub.org usage statistics. https://nanohub.org/usage, (accessed March 1, 2022).Google ScholarThere is no corresponding record for this reference.
- 4Dubbeldam, D.; Calero, S.; Ellis, D. E.; Snurr, R. Q. RASPA: molecular simulation software for adsorption and diffusion in flexible nanoporous materials. Mol. Simul. 2016, 42, 81– 101, DOI: 10.1080/08927022.2015.1010082Google Scholar4RASPA: molecular simulation software for adsorption and diffusion in flexible nanoporous materialsDubbeldam, David; Calero, Sofia; Ellis, Donald E.; Snurr, Randall Q.Molecular Simulation (2016), 42 (2), 81-101CODEN: MOSIEA; ISSN:0892-7022. (Taylor & Francis Ltd.)A new software package, RASPA, for simulating adsorption and diffusion of mols. in flexible nanoporous materials is presented. The code implements the latest state-of-the-art algorithms for mol. dynamics and Monte Carlo (MC) in various ensembles including symplectic/measure-preserving integrators, Ewald summation, configurational-bias MC, continuous fractional component MC, reactive MC and Baker's minimisation. We show example applications of RASPA in computing coexistence properties, adsorption isotherms for single and multiple components, self- and collective diffusivities, reaction systems and visualisation. The software is released under the GNU General Public License.
- 5Rogge, S. M. J.; Goeminne, R.; Demuynck, R.; Gutiérrez-Sevillano, J. J.; Vandenbrande, S.; Vanduyfhuys, L.; Waroquier, M.; Verstraelen, T.; Van Speybroeck, V. Modeling gas adsorption in flexible metal-organic frameworks via hybrid Monte Carlo/molecular dynamics schemes. Adv. Theory Simul. 2019, 2, 1800177 DOI: 10.1002/adts.201800177Google Scholar5Modeling Gas Adsorption in Flexible Metal-Organic Frameworks via Hybrid Monte Carlo/Molecular Dynamics SchemesRogge, Sven M. J.; Goeminne, Ruben; Demuynck, Ruben; Gutierrez-Sevillano, Juan Jose; Vandenbrande, Steven; Vanduyfhuys, Louis; Waroquier, Michel; Verstraelen, Toon; Van Speybroeck, VeroniqueAdvanced Theory and Simulations (2019), 2 (4), 1800177CODEN: ATSDCW; ISSN:2513-0390. (Wiley-VCH Verlag GmbH & Co. KGaA)Herein, a hybrid Monte Carlo (MC)/mol. dynamics (MD) simulation protocol that properly accounts for the extraordinary structural flexibility of metal-org. frameworks (MOFs) is developed and validated. This is vital to accurately predict gas adsorption isotherms and guest-induced flexibility of these materials. First, the performance of three recent models to predict adsorption isotherms and flexibility in MOFs is critically investigated. While these methods succeed in providing qual. insight in the gas adsorption process in MOFs, their accuracy remains limited as the intrinsic flexibility of these materials is very hard to account for. To overcome this challenge, a hybrid MC/MD simulation protocol that is specifically designed to handle the flexibility of the adsorbent, including the shape flexibility, is introduced, thereby unifying the strengths of the previous models. It is demonstrated that the application of this new protocol to the adsorption of neon, argon, xenon, methane, and carbon dioxide in MIL-53(Al), a prototypical flexible MOF, substantially decreases the inaccuracy of the obtained adsorption isotherms and predicted guest-induced flexibility. As a result, this method is ideally suited to rationalize the adsorption performance of flexible nanoporous materials at the mol. level, paving the way for the conscious design of MOFs as industrial adsorbents.
- 6Lin, L. C.; Lee, K.; Gagliardi, L.; Neaton, J. B.; Smit, B. Force-field development from electronic structure calculations with periodic boundary conditions: applications to gaseous adsorption and transport in metal-organic frameworks. J. Chem. Theory Comput. 2014, 10, 1477– 1488, DOI: 10.1021/ct500094wGoogle Scholar6Force-Field Development from Electronic Structure Calculations with Periodic Boundary Conditions: Applications to Gaseous Adsorption and Transport in Metal-Organic FrameworksLin, Li-Chiang; Lee, Kyuho; Gagliardi, Laura; Neaton, Jeffrey B.; Smit, BerendJournal of Chemical Theory and Computation (2014), 10 (4), 1477-1488CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)A systematic and efficient methodol. to derive accurate (nonpolarizable) force fields from periodic d. functional theory (DFT) calcns. for use in classical mol. simulations has been developed. The methodol. requires reduced computation cost compared with other conventional ways. The whole process is performed self-consistently in a fully periodic system. The force fields derived by using this methodol. accurately predict the CO2 and H2O adsorption isotherms inside Mg-MOF-74, and is transferable to Zn-MOF-74; by replacing the Mg-CO2 interactions with the corresponding Zn-CO2 interactions, we obtain an accurate prediction of the corresponding isotherm. This methodol. was used to address the effect of water on the sepn. of flue gases in these materials. In general, the mixt. isotherms of CO2 and H2O calcd. with these derived force fields show a significant redn. in CO2 uptake with the existence of trace amts. of water vapor. The effect of water, however, was found to be quant. different between Mg- and Zn-MOF-74.
- 7Dubbeldam, D.; Walton, K. S.; Vlugt, T. J. H.; Calero, S. Design, parameterization, and implementation of atomic force fields for adsorption in nanoporous materials. Adv. Theory Simul. 2019, 2, 1900135 DOI: 10.1002/adts.201900135Google Scholar7Design, Parameterization, and Implementation of Atomic Force Fields for Adsorption in Nanoporous MaterialsDubbeldam, David; Walton, Krista S.; Vlugt, Thijs J. H.; Calero, SofiaAdvanced Theory and Simulations (2019), 2 (11), 1900135CODEN: ATSDCW; ISSN:2513-0390. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Mol. simulations are an excellent tool to study adsorption and diffusion in nanoporous materials. Examples of nanoporous materials are zeolites, carbon nanotubes, clays, metal-org. frameworks (MOFs), covalent org. frameworks (COFs) and zeolitic imidazolate frameworks (ZIFs). The mol. confinement these materials offer has been exploited in adsorption and catalysis for almost 50 years. Mol. simulations have provided understanding of the underlying shape selectivity, and adsorption and diffusion effects. Much of the reliability of the modeling predictions depends on the accuracy and transferability of the force field. However, flexibility and the chem. and structural diversity of MOFs add significant challenges for engineering force fields that are able to reproduce exptl. obsd. structural and dynamic properties. Recent developments in design, parameterization, and implementation of force fields for MOFs and zeolites are reviewed.
- 8Kaplan, I. G. Intermolecular Interactions: Physical Picture, Computational Methods and Model Potentials; John Wiley & Sons: West Sussex, U.K., 2006.Google ScholarThere is no corresponding record for this reference.
- 9Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.; Skiff, W. M. UFF, a full periodic-table force-field for molecular mechanics and molecular-dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024– 10035, DOI: 10.1021/ja00051a040Google Scholar9UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulationsRappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A., III; Skiff, W. M.Journal of the American Chemical Society (1992), 114 (25), 10024-35CODEN: JACSAT; ISSN:0002-7863.A new mol. mechanics force field, the Universal force field (UFF), is described wherein the force field parameters are estd. using general rules based only on the element, its hybridization and its connectivity. The force field functional forms, parameters, and generating formulas for the full periodic table are presented.
- 10Walters, E. T.; Mohebifar, M.; Johnson, E. R.; Rowley, C. N. Evaluating the London dispersion coefficients of protein force fields using the exchange-hole dipole moment model. J. Phys. Chem. B 2018, 122, 6690– 6701, DOI: 10.1021/acs.jpcb.8b02814Google Scholar10Evaluating the London Dispersion Coefficients of Protein Force Fields Using the Exchange-Hole Dipole Moment ModelWalters, Evan T.; Mohebifar, Mohamad; Johnson, Erin R.; Rowley, Christopher N.Journal of Physical Chemistry B (2018), 122 (26), 6690-6701CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)London dispersion is one of the fundamental interactions involved in protein folding and dynamics. The popular CHARMM36, Amber ff14sb, and OPLS-AA force fields represent these interactions through the C6/r6 term of the Lennard-Jones potential, where the C6 parameters are assigned empirically. Here, dispersion coeffs. of these three force fields are shown to be roughly 50% larger than values calcd. using the quantum-mech.-derived exchange-hole dipole moment (XDM) model. The CHARMM36 and Amber OL15 force fields for nucleic acids also exhibit this trend. The hydration energies of the side-chain models were calcd. using REMD-TI for the CHARMM36, Amber ff14sb, and OPLS-AA force fields. These force fields predict side-chain hydration energies that are in generally good agreement with the exptl. values, which suggests the total strength of aq. dispersion interactions is correct, despite C6 coeffs. that are considerably larger than XDM predicts. An anal. expression for the dispersion hydration energy using XDM coeffs. shows that higher-order dispersion terms (i.e., C8 and C10) account for roughly 37.5% of the hydration energy of methane. This suggests that the C6 dispersion coeffs. used in contemporary force fields are elevated to account for the neglected higher-order terms.
- 11Jones, J. E. On the determination of molecular fields - II From the equation of state of a gas. Proc. R. Soc. London, Ser. A 1924, 106, 463– 477, DOI: 10.1098/rspa.1924.0082Google Scholar11The determination of molecular fields (II) From the equation of state of a gasJones, J. R.Proceedings of the Royal Society of London, Series A: Mathematical, Physical and Engineering Sciences (1924), 106 (), 463-77CODEN: PRLAAZ; ISSN:1364-5021.The virial coeffs. are evaluated for an equation of state of gases for the case of a mol. with attractive and repelling forces, as described above. The values obtained for the consts. of these forces for A agree fairly well with those calcd. from the viscosity data. In an appendix, recalcn. on the basis of more recent data on the isotherms of A give better agreement, pointing to 14 1/3 as the best value for the index of the repulsive force.
- 12Boda, D.; Henderson, D. The effects of deviations from Lorentz–Berthelot rules on the properties of a simple mixture. Mol. Phys. 2008, 106, 2367– 2370, DOI: 10.1080/00268970802471137Google Scholar12The effects of deviations from Lorentz-Berthelot rules on the properties of a simple mixtureBoda, Dezso; Henderson, DouglasMolecular Physics (2008), 106 (20), 2367-2370CODEN: MOPHAM; ISSN:0026-8976. (Taylor & Francis Ltd.)Generally, the parameters in the interaction potential between like mols. in a mixt. can be detd. in a relatively straightforward manner from the properties of the pure components. However, the detn. of the parameters in the interaction potential between the unlike pairs in the mixts. is more difficult. As a result, these parameters are usually estd. from avs. of the like parameters. The most common recipes are the Lorentz-Berthelot mixing rules, where the energy and mol. size parameters are presumed to be geometric and arithmetic avs., resp. There have been some studies of the consequences of deviations from the energy rule but almost no studies of the consequences of deviations from the size rule. Here, we study the effects of deviations from both rules on the radial distribution functions of a simple mixt. We find from simulations that, for this mixt., the effect of deviations from the energy rule on the radial distribution function are rather small but that the effect of deviations from the size rule can be significant and are interesting.
- 13Berthelot, D. Sur le mélange des gaz. Compt. Rendus 1898, 126, 1703– 1706Google Scholar13On mixtures of gasesBerthelot, D.Comptes Rendus Hebdomadaires des Seances de l'Academie des Sciences (1898), 126 (), 1703CODEN: COREAF ISSN:.The author proposes a formula of the van der Waals type for a mixture of two gases and shows how the constants A and B for the mixture can be calculated from the corresponding constants for the single gases.
- 14Lorentz, H. A. Ueber die anwendung des satzes vom virial in der kinetischen theorie der gase. Ann. Phys. 1881, 248, 127– 136, DOI: 10.1002/andp.18812480110Google ScholarThere is no corresponding record for this reference.
- 15Dubbeldam, D.; Calero, S.; Vlugt, T. J. H.; Ellis, D. E.; Snurr, R. Q. RASPA 2.0: Molecular Software Package for Adsorption and Diffusion in (Flexible) Nanoporous Materials , program manual, March 2015, pp 1− 145.Google ScholarThere is no corresponding record for this reference.
- 16Dubbeldam, D.; Walton, K. S.; Ellis, D. E.; Snurr, R. Q. Exceptional negative thermal expansion in isoreticular metal-organic frameworks. Angew. Chem., Int. Ed. 2007, 46, 4496– 4499, DOI: 10.1002/anie.200700218Google Scholar16Exceptional negative thermal expansion in isoreticular metal-organic frameworksDubbeldam, David; Walton, Krista S.; Ellis, Donald E.; Snurr, Randall Q.Angewandte Chemie, International Edition (2007), 46 (24), 4496-4499CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A new model for flexible frameworks is used to simulate the structures and adsorption properties of isoreticular metal-org. frameworks (IRMOFs), such as IRMOF-1. The structural simulations suggest that the IRMOFs have neg. thermal-expansion coeffs. over their full temp. ranges of stability.
- 17Chen, B.; Siepmann, J. I. Transferable potentials for phase equilibria. 3. Explicit-hydrogen description of normal alkanes. J. Phys. Chem. B 1999, 103, 5370– 5379, DOI: 10.1021/jp990822mGoogle Scholar17Transferable potentials for phase equilibria. 3. Explicit-hydrogen description of normal alkanesChen, Bin; Siepmann, J. IljaJournal of Physical Chemistry B (1999), 103 (25), 5370-5379CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)Motivated by shortcomings of the available united-atom models for alkanes, a new explicit-H model for n-alkanes (TraPPE-EH, transferable potentials for phase equil.-explicit H) is developed from fitting to 1-component fluid-phase properties. In addn. to Lennard-Jones sites on C atoms, this model utilizes Lennard-Jones sites on the centers of C-H bonds. Configurational-bias Monte Carlo simulations in the Gibbs and canonical ensembles were carried out to calc. the 1-component vapor-liq. phase equil. for CH4 to n-dodecane, to det. the phase diagram of supercrit. C2H6 and n-heptane mixts., to obtain the Gibbs free energies of transfer for n-pentane (I) and n-hexane between He vapor and n-heptane liq. phases, and to study the high-pressure region of the equation of state for I and n-decane. The explicit-H representation, with its more faithful description of the mol. shape of alkanes, allows us to find a set of Lennard-Jones parameters that yields significantly better agreement with expt. for 1- and multicomponent phase equil. than our united-atom alkane model, but the price is higher computational cost.
- 18Eggimann, B. L.; Sunnarborg, A. J.; Stern, H. D.; Bliss, A. P.; Siepmann, J. I. An online parameter and property database for the TraPPE force field. Mol. Simul. 2014, 40, 101– 105, DOI: 10.1080/08927022.2013.842994Google Scholar18An online parameter and property database for the TraPPE force fieldEggimann, Becky L.; Sunnarborg, Amara J.; Stern, Hudson D.; Bliss, Andrew P.; Siepmann, J. IljaMolecular Simulation (2014), 40 (1-3), 101-105CODEN: MOSIEA; ISSN:0892-7022. (Taylor & Francis Ltd.)A review. The transferable potentials for phase equil. (TraPPE) force field aims to be accurate, computationally efficient and applicable to a wide range of chem. compds., state points and thermophys. properties. When new users wish to implement TraPPE models into their chosen simulation program, they face several obstacles: the TraPPE models are dispersed over many sep. publications and misinterpretations of the primary literature are possible; the TraPPE force field makes specific choices for std. conventions that may require non-trivial code modifications for some simulation software. Therefore, the TraPPE developers report here a resource website and online searchable parameter and property database designed to provide new and experienced users with tools for successful implementation and validation (http://www.chem.umn.edu/groups/siepmann/trappe/).
- 19Martin, M. G.; Siepmann, J. I. Transferable potentials for phase equilibria. 1. United-atom description of n-alkanes. J. Phys. Chem. B 1998, 102, 2569– 2577, DOI: 10.1021/jp972543+Google Scholar19Transferable Potentials for Phase Equilibria. 1. United-Atom Description of n-AlkanesMartin, Marcus G.; Siepmann, J. IljaJournal of Physical Chemistry B (1998), 102 (14), 2569-2577CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A new set of united-atom Lennard-Jones interaction parameters for n-alkanes is proposed from fitting to crit. temps. and satd. liq. densities. Configurational-bias Monte Carlo simulations in the Gibbs ensemble were carried out to det. the vapor-liq. coexistence curves for methane to dodecane using three united-atom force fields: OPLS [Jorgensen, et al. J. Am. Chem. Soc. 1984, 106, 813], SKS [Siepmann, et al. Nature 1993, 365, 330], and TraPPE. Std. specific densities and the high-pressure equation-of-state for the transferable potentials for phase equil. (TraPPE) model were studied by simulations in the isobaric-isothermal and canonical ensembles, resp. It is found that one set of Me and methylene parameters is sufficient to accurately describe the fluid phases of all n-alkanes with two or more carbon atoms. Whereas other n-alkane force fields employ Me groups that are either equal or larger in size than the methylene groups, it is demonstrated here that using a smaller Me group yields a better fit to the set of exptl. data. As should be expected from an effective pair potential, the new parameters do not reproduce exptl. second virial coeffs. Satd. vapor pressures and densities show small, but systematic deviation from the exptl. data.
- 20Potoff, J. J.; Siepmann, J. I. Vapor–liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogen. AIChE J. 2001, 47, 1676– 1682, DOI: 10.1002/aic.690470719Google Scholar20Vapor-liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogenPotoff, Jeffrey J.; Siepmann, J. IljaAIChE Journal (2001), 47 (7), 1676-1682CODEN: AICEAC; ISSN:0001-1541. (American Institute of Chemical Engineers)New force fields for carbon dioxide and nitrogen are introduced that quant. reproduce the vapor-liq. equil. (VLE) of the neat systems and their mixts. with alkanes. In addn. to the usual VLE calcns. for pure CO2 and N2, calcns. of the binary mixts. with propane were used in the force-field development to achieve a good balance between dispersive and electrostatic (quadrupole-quadrupole) interactions. The transferability of the force fields was then assessed from calcns. of the VLE for the binary mixts. with n-hexane, the binary mixt. of CO2/N2, and the ternary mixt. of CO2/N2/propane. The VLE calcns. were carried out using configurational-bias Monte Carlo simulations in either the grand canonical ensemble with histogram-reweighting or in the Gibbs ensemble.
- 21Darkrim, F.; Levesque, D. Monte Carlo simulations of hydrogen adsorption in single-walled carbon nanotubes. J. Chem. Phys. 1998, 109, 4981– 4984, DOI: 10.1063/1.477109Google Scholar21Monte Carlo simulations of hydrogen adsorption in single-walled carbon nanotubesDarkrim, Farida; Levesque, DominiqueJournal of Chemical Physics (1998), 109 (12), 4981-4984CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Within the framework of a study on the properties of carbon nanotubes, a promising new material, we performed numerical simulation of hydrogen adsorption at room temp. in single-walled nanotubes. The structure of this material is favorable to the adsorption phenomenon because of the narrow size distribution of the nanotube diams., which have dimensions on the order of the range of the carbon attractive interaction. We discuss the influence of the single-walled carbon nanotube diams. on the relative arrangement of carbon atoms and hydrogen mols. within an array of parallel single-walled carbon nanotubes. We also studied the influence on adsorption of the distance between the nearest-neighbor nanotubes.
- 22Zhang, L.; Siepmann, J. I. Direct calculation of Henry’s law constants from Gibbs ensemble Monte Carlo simulations: nitrogen, oxygen, carbon dioxide and methane in ethanol. Theor. Chem. Acc. 2006, 115, 391– 397, DOI: 10.1007/s00214-005-0073-1Google Scholar22Direct calculation of Henry's law constants from Gibbs ensemble Monte Carlo simulations: nitrogen, oxygen, carbon dioxide and methane in ethanolZhang, Ling; Siepmann, J. IljaTheoretical Chemistry Accounts (2006), 115 (5), 391-397CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)Configurational-bias Monte Carlo simulations in the Gibbs ensemble were used to calc. Henry's law consts., Ostwald solubilities, and Gibbs free energies of transfer for oxygen, nitrogen, methane, and carbon dioxide in ethanol at 323 and 373 K. These three soly. descriptors can be expressed as functions of mech. properties that are directly observable in the Gibbs ensemble approach, thereby allowing for very precise detn. of the descriptors. Addnl., the Henry's law consts. of multiple solutes can be computed from a single simulation. Most of the simulations were carried out for systems contg. 1000 solvent and up to 8 solute mols., and further simulations using either 500 or 2000 solvent mols. point to negligible system size effects. A comparison with exptl. data shows that the united-atom version of the transferable potential for phase equil. force field yields Henry's law consts. that reproduce well the differences between the four solutes and the changes upon increase of the temp.
- 23Mayo, S. L.; Olafson, B. D.; Goddard, W. A. DREIDING - A generic force field for molecular simulations. J. Phys. Chem. 1990, 94, 8897– 8909, DOI: 10.1021/j100389a010Google Scholar23DREIDING: a generic force field for molecular simulationsMayo, Stephen L.; Olafson, Barry D.; Goddard, William A., IIIJournal of Physical Chemistry (1990), 94 (26), 8897-909CODEN: JPCHAX; ISSN:0022-3654.The parameters are given for a new generic force field, DREIDING, that is useful for predicting structures and dynamics of org., biol., and main-group inorg. mols. The philosophy in DREIDING is to use general force consts. and geometry parameters based on simple hybridization considerations rather than individual force consts. and geometric parameters that depend on the particular combination of atoms involved in the bond, angle, or torsion terms. Thus all bond distances are derived from at. radii, and there is only one force const. each for bonds, angles, and inversions and only six different values for torsional barriers. Parameters are defined for all possible combinations of atoms and new atoms can be added to the force field rather simply. The parameters are given for the "nonmetallic" main-group elements (B, C, N, O, F columns for the C, Si, Ge, and Sn rows) plus H and a few metals (Na, Ca, Zn, Fe). The accuracy of the DREIDING force field is tested by comparing with (i) 76 accurately detd. crystal structures of org. compds. involving H, C, N, O, F, P, S, Cl, and Br, (ii) rotational barriers of a no. of mols., and (iii) relative conformational energies and barriers of a no. of mols. There is excellent agreement.
- 24Bai, P.; Tsapatsis, M.; Siepmann, J. I. TraPPE-zeo: Transferable potentials for phase equilibria force field for all-silica zeolites. J. Phys. Chem. C 2013, 117, 24375– 24387, DOI: 10.1021/jp4074224Google Scholar24TraPPE-zeo: Transferable Potentials for Phase Equilibria Force Field for All-Silica ZeolitesBai, Peng; Tsapatsis, Michael; Siepmann, J. IljaJournal of Physical Chemistry C (2013), 117 (46), 24375-24387CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)The transferable potentials for phase equil. (TraPPE) force field is extended to all-silica zeolites. This novel force field is parametrized to match the exptl. adsorption isotherms of n-heptane, propane, carbon dioxide, and ethanol with the Lennard-Jones parameters for sorbate-framework interactions detd. in a consistent manner using the Lorentz-Berthelot combining rules as for other parts of the TraPPE force field. The TraPPE-zeo force field allows for accurate predictions for both adsorption and diffusion of alkanes, alcs., carbon dioxide, and water over a wide range of pressures and temps. To achieve transferability to a wider range of mol. types, ranging from nonpolar to dipolar and hydrogen-bonding compds., Lennard-Jones interaction sites and partial charges are placed at both the oxygen and the silicon atoms of the zeolite lattice, which allows for a better balance of dispersive and 1st-order electrostatic interactions than is achievable with the Lennard-Jones potential used only for the oxygen atoms. The use of the Lorentz-Berthelot combining rules for unlike interactions makes the TraPPE-zeo force field applicable to any sorbate as long as the relevant TraPPE sorbate-sorbate parameters are available. The TraPPE-zeo force field allows for greatly improved predictive power compared to force fields that explicitly tabulate the individual cross-interaction parameters.
- 25Dubbeldam, D.; Torres-Knoop, A.; Walton, K. S. On the inner workings of Monte Carlo codes. Mol. Simul. 2013, 39, 1253– 1292, DOI: 10.1080/08927022.2013.819102Google Scholar25On the inner workings of Monte Carlo codesDubbeldam, David; Torres-Knoop, Ariana; Walton, Krista S.Molecular Simulation (2013), 39 (14-15), 1253-1292CODEN: MOSIEA; ISSN:0892-7022. (Taylor & Francis Ltd.)We review state-of-the-art Monte Carlo (MC) techniques for computing fluid coexistence properties (Gibbs simulations) and adsorption simulations in nanoporous materials such as zeolites and metal-org. frameworks. Conventional MC is discussed and compared to advanced techniques such as reactive MC, configurational-bias Monte Carlo and continuous fractional MC. The latter technique overcomes the problem of low insertion probabilities in open systems. Other modern methods are (hyper-)parallel tempering, Wang-Landau sampling and nested sampling. Details on the techniques and acceptance rules as well as to what systems these techniques can be applied are provided. We highlight consistency tests to help validate and debug MC codes.
- 26Heinen, J.; Burtch, N. C.; Walton, K. S.; Dubbeldam, D. Flexible force field parameterization through fitting on the ab initio-derived elastic tensor. J. Chem. Theory Comput. 2017, 13, 3722– 3730, DOI: 10.1021/acs.jctc.7b00310Google Scholar26Flexible Force Field Parameterization through Fitting on the Ab Initio-Derived Elastic TensorHeinen, Jurn; Burtch, Nicholas C.; Walton, Krista S.; Dubbeldam, DavidJournal of Chemical Theory and Computation (2017), 13 (8), 3722-3730CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Constructing functional forms and their corresponding force field parameters for the metal-linker interface of metal-org. frameworks is challenging. We propose to fit these parameters on the elastic tensor, computed from ab initio d. functional theory calcns. The advantage of this top-down approach is that it becomes evident if functional forms are missing when components of the elastic tensor are off. As a proof-of-concept, a new flexible force field for MIL-47(V) is derived. Neg. thermal expansion is obsd. and framework flexibility has a negligible effect on adsorption and transport properties for small guest mols. We believe this force field parameterization approach can serve as a useful tool for developing accurate flexible force field models that capture the correct mech. behavior of the full periodic structure.
- 27Heinen, J.; Dubbeldam, D. On flexible force fields for metal-organic frameworks: Recent developments and future prospects. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2018, 8, e1363 DOI: 10.1002/wcms.1363Google ScholarThere is no corresponding record for this reference.
- 28Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids, 2nd ed.; Oxford University Press: Oxford, U.K., 2017.Google ScholarThere is no corresponding record for this reference.
- 29Frenkel, D.; Smit, B. Understanding Molecular Simulation: From Algorithms to Applications, 2nd ed.; Academic Press: San Diego, California, 2002.Google ScholarThere is no corresponding record for this reference.
- 30Hastings, W. K. Monte Carlo sampling methods using Markov chains and their applications. Biometrika 1970, 57, 97– 109, DOI: 10.1093/biomet/57.1.97Google ScholarThere is no corresponding record for this reference.
- 31Metropolis, N.; Rosenbluth, A. W.; Rosenbluth, M. N.; Teller, A. H.; Teller, E. Equation of state calculations by fast computing machines. J. Chem. Phys. 1953, 21, 1087– 1092, DOI: 10.1063/1.1699114Google Scholar31Equation-of-state calculations by fast computing machinesMetropolis, Nicholas; Rosenbluth, Arianna W.; Rosenbluth, Marshall N.; Teller, Augusta H.; Teller, EdwardJournal of Chemical Physics (1953), 21 (), 1087-92CODEN: JCPSA6; ISSN:0021-9606.A general method, suitable for fast computing machines, is described for investigating such properties as equations of state for substances consisting of interacting individual mols. The method consists of a modified Monte Carlo integration over configuration space. Results for the 2-dimensional rigid-sphere system were obtained on the Los Alamos MANIAC. These results were compared to the free-vol. equation of state and to a 4-term virial coeff. expansion.
- 32Hirotani, A.; Mizukami, K.; Miura, R.; Takaba, H.; Miya, T.; Fahmi, A.; Stirling, A.; Kubo, M.; Miyamoto, A. Grand canonical Monte Carlo simulation of the adsorption of CO2 on silicalite and NaZSM-5. Appl. Surf. Sci. 1997, 120, 81– 84, DOI: 10.1016/S0169-4332(97)00222-5Google Scholar32Grand canonical Monte Carlo simulation of the adsorption of CO2 on silicalite and NaZSM-5Hirotani, Akiyasu; Mizukami, Koichi; Miura, Ryuji; Takaba, Hiromitsu; Miya, Takeshi; Fahmi, Adil; Stirling, Andras; Kubo, Momoji; Miyamoto, AkiraApplied Surface Science (1997), 120 (1/2), 81-84CODEN: ASUSEE; ISSN:0169-4332. (Elsevier)The adsorption of carbon dioxide in silicalite and NaZSM-5 zeolite has been studied using new Monte Carlo software. In this program, sodium cations and framework are movable during the simulation. The calcd. adsorption isotherms are in good agreement with the exptl. results. The energy distribution of carbon dioxide over silicalite and NaZSM-5 shows that the increase of the adsorption energy for NaZSM-5 is mainly due the elec. field generated by sodium cations.
- 33Gao, G.; Wang, W. Gibbs Ensemble Monte Carlo simulation of binary vapor-liquid equilibria for CFC alternatives. Fluid Phase Equilib. 1997, 130, 157– 166, DOI: 10.1016/S0378-3812(96)03195-0Google Scholar33Gibbs Ensemble Monte Carlo simulation of binary vapor-liquid equilibria for CFC alternativesGao, Guangtu; Wang, WenchuanFluid Phase Equilibria (1997), 130 (1-2), 157-166CODEN: FPEQDT; ISSN:0378-3812. (Elsevier)The Gibbs Ensemble Monte Carlo (GEMC) simulation method has been used for vapor-liq. equil. calcns. of the binary systems R22-R142b and R22-R152a, which are considered to be promising mixt. refrigerants for replacing R12. All the mols. have been treated in terms of an effective Lennard-Jones potential, with temp. dependent parameters regressed by fitting vapor pressures and satd. liq. densities at various temps. for pure substances of interest. Good agreement between exptl. and calcd. data from the vol.-translation Peng-Robinson equation of state and the simulated results, including the compns., molar volumes for both the vapor and liq. phases, and heats of vaporization, indicates that the GEMC method can be applied to the description of phase behavior for these CFC alternative binary systems.
- 34Slepoy, A.; Thompson, A. P.; Plimpton, S. J. A constant-time kinetic Monte Carlo algorithm for simulation of large biochemical reaction networks. J. Chem. Phys. 2008, 20, 205101 DOI: 10.1063/1.2919546Google ScholarThere is no corresponding record for this reference.
- 35Dove, M. T. An introduction to atomistic simulation methods. Seminarios de la SEM 2008, 4, 7– 37Google ScholarThere is no corresponding record for this reference.
- 36Braun, E.; Gilmer, J.; Mayes, H. B.; Mobley, D. L.; Monroe, J. I.; Prasad, S.; Zuckerman, D. M. Best practices for foundations in molecular simulations. Living J. Comput. Mol. Sci. 2019, 1, 5957 DOI: 10.33011/livecoms.1.1.5957Google ScholarThere is no corresponding record for this reference.
- 37Abouelnasr, M. K. F.; Smit, B. Diffusion in confinement: kinetic simulations of self- and collective diffusion behavior of adsorbed gases. Phys. Chem. Chem. Phys. 2012, 14, 11600– 11609, DOI: 10.1039/c2cp41147dGoogle Scholar37Diffusion in confinement: kinetic simulations of self- and collective diffusion behavior of adsorbed gasesAbouelnasr, Mahmoud K. F.; Smit, BerendPhysical Chemistry Chemical Physics (2012), 14 (33), 11600-11609CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The self- and collective-diffusion behaviors of adsorbed methane, helium, and isobutane in zeolite frameworks LTA, MFI, AFI, and SAS were examd. at various concns. using a range of mol. simulation techniques including Mol. Dynamics (MD), Monte Carlo (MC), Bennett-Chandler (BC), and kinetic Monte Carlo (kMC). This paper has three main results. (1) A novel model for the process of adsorbate movement between two large cages was created, allowing the formulation of a mixing rule for the re-crossing coeff. between two cages of unequal loading. The predictions from this mixing rule were found to agree quant. with explicit simulations. (2) A new approach to the dynamically cor. Transition State Theory method to anal. calc. self-diffusion properties was developed, explicitly accounting for nanoscale fluctuations in concn. This approach was demonstrated to quant. agree with previous methods, but is uniquely suited to be adapted to a kMC simulation that can simulate the collective-diffusion behavior. (3) While at low and moderate loadings the self- and collective-diffusion behaviors in LTA are obsd. to coincide, at higher concns. they diverge. A change in the adsorbate packing scheme was shown to cause this divergence, a trait which is replicated in a kMC simulation that explicitly models this behavior. These phenomena were further investigated for isobutane in zeolite MFI, where MD results showed a sepn. in self- and collective- diffusion behavior that was reproduced with kMC simulations.
- 38Skoulidas, A. I.; Sholl, D. S. Self-diffusion and transport diffusion of light gases in metal-organic framework materials assessed using molecular dynamics simulations. J. Phys. Chem. B 2005, 109, 15760– 15768, DOI: 10.1021/jp051771yGoogle Scholar38Self-Diffusion and Transport Diffusion of Light Gases in Metal-Organic Framework Materials Assessed Using Molecular Dynamics SimulationsSkoulidas, Anastasios I.; Sholl, David S.Journal of Physical Chemistry B (2005), 109 (33), 15760-15768CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Metal-org. framework (MOF) materials pose an interesting alternative to more traditional nanoporous materials for a variety of sepn. processes. Sepn. processes involving nanoporous materials can be controlled by either adsorption equil., diffusive transport rates, or a combination of these factors. Adsorption equil. was studied for a variety of gases in MOFs, but almost nothing is currently known about mol. diffusion rates in MOFs. The authors used equil. mol. dynamics (MD) to probe the self-diffusion and transport diffusion of a no. of small gas species in several MOFs as a function of pore loading at room temp. Specifically, the authors have studied Ar, CH4, CO2, N2, and H2 diffusion in MOF-5. The diffusion of Ar in MOF-2, MOF-3, and Cu-BTC was assessed in a similar manner. Results greatly expand the range of MOFs for which data describing mol. diffusion is available. The authors discuss the prospects for exploiting mol. transport properties in MOFs in practical sepn. processes and the future role of MD simulations in screening families of MOFs for these processes.
- 39Werder, T.; Walther, J. H.; Jaffe, R. L.; Halicioglu, T.; Koumoutsakos, P. On the water-carbon interaction for use in molecular dynamics simulations of graphite and carbon nanotubes. J. Phys. Chem. B 2003, 107, 1345– 1352, DOI: 10.1021/jp0268112Google Scholar39On the Water-Carbon Interaction for Use in Molecular Dynamics Simulations of Graphite and Carbon NanotubesWerder, T.; Walther, J. H.; Jaffe, R. L.; Halicioglu, T.; Koumoutsakos, P.Journal of Physical Chemistry B (2003), 107 (6), 1345-1352CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)A systematic mol. dynamics study shows that the contact angle of a H2O droplet on graphite changes significantly as a function of the H2O-C interaction energy. Together with the observation that a linear relation can be established between the contact angle and the H2O monomer binding energy on graphite, a new route to calibrate interaction potential parameters is presented. Through a variation of the droplet size at 1000-17,500 H2O mols., the authors det. the line tension to be pos. and ∼2 × 10-10 J/m. To recover a macroscopic contact angle of 86°, a H2O monomer binding energy of -6.33 kJ mol-1 is required, which is obtained by applying a C-oxygen Lennard-Jones potential with the parameters εCO = 0.392 kJ mol-1 and σCO = 3.19 Å. For this new H2O-C interaction potential, the authors present d. profiles and hydrogen bond distributions for a H2O droplet on graphite.
- 40Henry, A. S.; Chen, G. Spectral phonon transport properties of silicon based on molecular dynamics simulations and lattice dynamics. J. Comput. Theor. Nanosci. 2008, 5, 141– 152, DOI: 10.1166/jctn.2008.2454Google Scholar40Spectral phonon transport properties of silicon based on molecular dynamics simulations and lattice dynamicsHenry, Asegun S.; Chen, GangJournal of Computational and Theoretical Nanoscience (2008), 5 (2), 141-152CODEN: JCTNAB; ISSN:1546-1955. (American Scientific Publishers)Although the thermal cond. of silicon has been studied before, current estns. for the phonon mean free paths have not provided full explanation of the strong size effects exptl. obsd. for various silicon micro and nanostructures. Since phonon relaxation time models are mostly semi-empirical, the mean free paths cannot be detd. reliably and questions remain as to which polarizations, frequencies, and wavelengths are dominant heat carriers. Here we have used a combination of equil. mol. dynamics simulations and lattice dynamics calcns. to fully detail the spectral dependence of phonon transport properties in bulk silicon. By considering the frequency dependence of the sp. heat, group velocities, and mean free paths, we address these unresolved questions and examine the errors assocd. with isotropic and frequency averaged approxns. Simulation details, such as the convergence of results on the simulation time and extn. of phonon transport properties in different crystallog. directions, are also discussed.
- 41Karplus, M.; McCammon, J. A. Molecular dynamics simulations of biomolecules. Nat. Struct. Biol. 2002, 9, 646– 652, DOI: 10.1038/nsb0902-646Google Scholar41Molecular dynamics simulations of biomoleculesKarplus, Martin; McCammon, J. AndrewNature Structural Biology (2002), 9 (9), 646-652CODEN: NSBIEW; ISSN:1072-8368. (Nature Publishing Group)A review. Mol. dynamics simulations are important tools for understanding the phys. basis of the structure and function of biol. macromols. The early view of proteins as relatively rigid structures has been replaced by a dynamic model in which the internal motions and resulting conformational changes play an essential role in their function. This review presents a brief description of the origin and early uses of biomol. simulations. It then outlines some recent studies that illustrate the utility of such simulations and closes with a discussion of their ever-increasing potential for contributing to biol.
- 42Umeh, J. C. Development of Simulation Tools on Nanuhub as Learning Modules. Master’s thesis; New Mexico State University: Las Cruces, NM, 2019.Google ScholarThere is no corresponding record for this reference.
- 43Humbert, M. T.; Zhang, Y.; Maginn, E. J. PyLAT: Python LAMMPS analysis tools. J. Chem. Inf. Model. 2019, 59, 1301– 1305, DOI: 10.1021/acs.jcim.9b00066Google Scholar43PyLAT: Python LAMMPS Analysis ToolsHumbert, Michael T.; Zhang, Yong; Maginn, Edward J.Journal of Chemical Information and Modeling (2019), 59 (4), 1301-1305CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Reproducibility and accuracy have become increasingly important issues for the mol. simulation community. The availability of validated open-source postprocessing tools to analyze simulation trajectories and compute properties is key to helping researchers conduct more accurate and reproducible simulations. Here the authors describe a suite of open-source Python-based postprocessing routines the authors have developed called PyLAT. PyLAT is compatible with the popular mol. dynamics package LAMMPS and enables users to compute viscosities, self-diffusivities, ionic conductivities, mol. or ion pair lifetimes, dielec. consts., and radial distribution functions using best-practice methods.
- 44Rowsell, J. L. C.; Yaghi, O. M. Effects of functionalization, catenation, and variation of the metal oxide and organic linking units on the low-pressure hydrogen adsorption properties of metal-organic frameworks. J. Am. Chem. Soc. 2006, 128, 1304– 1315, DOI: 10.1021/ja056639qGoogle Scholar44Effects of Functionalization, Catenation, and Variation of the Metal Oxide and Organic Linking Units on the Low-Pressure Hydrogen Adsorption Properties of Metal-Organic FrameworksRowsell, Jesse L. C.; Yaghi, Omar M.Journal of the American Chemical Society (2006), 128 (4), 1304-1315CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The dihydrogen adsorption isotherms of eight metal-org. frameworks (MOFs), measured at 77 K up to a pressure of 1 atm, were examd. for correlations with their structural features. All materials display approx. Type I isotherms with no hysteresis, and satn. was not reached for any of the materials under these conditions. Among the six iso-reticular MOFs (IRMOFs) studied, the catenated materials exhibit the largest capacities on a molar basis, up to 9.8 H2 per formula unit. The addn. of functional groups (-Br, -NH2, -C2H4-) to the phenylene links of IRMOF-1 (MOF-5), or their replacement with thieno[3,2-b]thiophene moieties in IRMOF-20, altered the adsorption behavior by a minor amt. despite large variations in the pore vols. of the resulting materials. But replacement of the metal oxide units with those contg. coordinatively unsatd. metal sites resulted in greater H2 uptake. The enhanced affinities of these materials, MOF-74 and HKUST-1, were further demonstrated by calcn. of the isosteric heats of adsorption, which were larger across much of the range of coverage examd., compared to those of representative IRMOFs. Probably under low-loading conditions, the H2 adsorption behavior of MOFs can be improved by imparting larger charge gradients on the metal oxide units and adjusting the link metrics to constrict the pore dimensions; however, a large pore vol. is still a prerequisite feature.
- 45Siberio-Perez, D. Y.; Wong-Foy, A. G.; Yaghi, O. M.; Matzger, A. J. Raman spectroscopic investigation of CH4 and N2 adsorption in metal-organic frameworks. Chem. Mater. 2007, 19, 3681– 3685, DOI: 10.1021/cm070542gGoogle Scholar45Raman Spectroscopic Investigation of CH4 and N2 Adsorption in Metal-Organic FrameworksSiberio-Perez, Diana Y.; Wong-Foy, Antek G.; Yaghi, Omar M.; Matzger, Adam J.Chemistry of Materials (2007), 19 (15), 3681-3685CODEN: CMATEX; ISSN:0897-4756. (American Chemical Society)The adsorption behavior of CH4 and N2 (298 K, 30 bar) in a series of isoreticular metal-org. frameworks (IRMOFs) was investigated by Raman spectroscopy. For CH4, the ν1 vibrational mode shifted to lower frequency by 7.6, 8.4, 11.0, 10.3, and 10.1 cm-1 from 2917 cm-1 when adsorbed to IRMOF-1, -6, -8, -11, and -18, resp. Along this same series, the adsorbed N2 stretch exhibited smaller shifts of 2.7, 3.1, 4.2, 4.1, and 3.7 cm-1. These shifts arise because of interactions within the framework pores, and not with the outer crystal surface. In all cases, Raman spectra at pressures up to 30 bar showed that satn. of the sorption sites does not occur. The obsd. shifts of the vibrational modes for each gas indicate different chem. environments within different IRMOFs, pointing to the important role the linkers play in the adsorption of gases.
- 46Duren, T.; Sarkisov, L.; Yaghi, O. M.; Snurr, R. Q. Design of new materials for methane storage. Langmuir 2004, 20, 2683– 2689, DOI: 10.1021/la0355500Google Scholar46Design of new materials for methane storageDuren Tina; Sarkisov Lev; Yaghi Omar M; Snurr Randall QLangmuir : the ACS journal of surfaces and colloids (2004), 20 (7), 2683-9 ISSN:0743-7463.One of the strategic goals of the modern automobile manufacturing industry is to replace gasoline and diesel with alternative fuels such as natural gas. In this report, we elucidate the desired characteristics of an optimal adsorbent for gas storage. The U.S. Department of Energy has outlined several requirements that adsorbents must fulfill for natural gas to become economically viable, with a key criterion being the amount adsorbed at 35 bar. We explore the adsorption characteristics of novel metal-organic materials (IRMOFs and molecular squares) and contrast them with the characteristics of two zeolites, MCM-41, and different carbon nanotubes. Using molecular simulations, we uncover the complex interplay of the factors influencing methane adsorption, especially the surface area, the capacity or free volume, the strength of the energetic interaction, and the pore size distribution. We also explain the extraordinary adsorption properties of IRMOF materials and propose new, not yet synthesized IRMOF structures with adsorption characteristics that are predicted to exceed the best experimental results to date by up to 36%.
- 47Walton, K. S.; Millward, A. R.; Dubbeldam, D.; Frost, H.; Low, J. J.; Yaghi, O. M.; Snurr, R. Q. Understanding inflection and steps in carbon dioxide adsorption isotherms in metal-organic frameworks. J. Am. Chem. Soc. 2007, 130, 406– 407, DOI: 10.1021/ja076595gGoogle ScholarThere is no corresponding record for this reference.
- 48Towns, J.; Cockerill, T.; Dahan, M.; Foster, I.; Gaither, K.; Grimshaw, A.; Hazlewood, V.; Lathrop, S.; Lifka, D.; Peterson, G. D.; Roskies, R.; Scott, J. R.; Wilkins-Diehr, N. XSEDE: accelerating scientific discovery. Comput. Sci. Eng. 2014, 16, 62– 74, DOI: 10.1109/MCSE.2014.80Google ScholarThere is no corresponding record for this reference.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 5 publications.
- Guozhao Liu, Hongyu Wang, Xiaolong Yang, Guanwen Su, Hongyuan Wei, Leping Dang. Insight on Crystal Morphology of ε-CL-20 in Binary Green Solvents: From Molecular Simulations to Experiments. Industrial & Engineering Chemistry Research 2025, 64
(3)
, 1762-1773. https://doi.org/10.1021/acs.iecr.4c04400
- Felipe L. Oliveira, Binquan Luan, Pierre M. Esteves, Mathias Steiner, Rodrigo Neumann Barros Ferreira. pyMSER─An Open-Source Library for Automatic Equilibration Detection in Molecular Simulations. Journal of Chemical Theory and Computation 2024, 20
(19)
, 8559-8568. https://doi.org/10.1021/acs.jctc.4c00417
- Hongyu Wang, Guozhao Liu, Guanwen Su, Hongyuan Wei, Leping Dang. Design of solvent systems for preparation of ε-CL-20 crystals with high sphericity assisted by molecular simulation. CrystEngComm 2025, 27
(4)
, 547-558. https://doi.org/10.1039/D4CE01181C
- Yutong Liu, Yawen Dong, Hua Wu. Comprehensive overview of machine learning applications in MOFs: from modeling processes to latest applications and design classifications. Journal of Materials Chemistry A 2025, 13
(4)
, 2403-2440. https://doi.org/10.1039/D4TA06740A
- Haixia He, Xiaoxi Guo, Yameng Wan, Jingwen Zhang, Xinyu Niu, Ruiai Wang, Qing Liu. Solubility, thermodynamic investigation and molecular dynamics simulation of guanidine hydrochloride in four binary solvents. Journal of Molecular Liquids 2023, 389 , 122902. https://doi.org/10.1016/j.molliq.2023.122902
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Figure 1
Figure 1. Lennard-Jones potential for argon. The energy unit is in Kelvin (used by RASPA); this can be multiplied by the Boltzmann constant to convert to Joules per particle.
Figure 2
Figure 2. Monte Carlo and molecular dynamics cycles.
Figure 3
Figure 3. Flow diagram illustrating the process to create a new nanoHUB tool.
Figure 4
Figure 4. NanoHUB workspace.
Figure 5
Figure 5. Rappture builder GUI.
Figure 6
Figure 6. TortoiseSVN repository browser.
Figure 7
Figure 7. Left panel: A log–log plot of mean-squared displacement versus time that shows the separate ballistic, transition, and diffusive regimes. Right panel: A linear–linear plot of mean-squared displacement versus time using the starting and ending times selected by the algorithm that extracts the self-diffusion constant Ds from a data subset within the diffusive regime.
Figure 8
Figure 8. Screenshots of Tool 1 for calculating the self-diffusion constant of a gas in a MOF.
Figure 9
Figure 9. Screenshots of Tool 2 for calculating the absolute and excess adsorption of a gas in a MOF.
Figure 10
Figure 10. Screenshot of Tool 3 for calculating the Henry’s coefficients of n-alkanes in porous materials.
Figure 11
Figure 11. Screenshots of Tool 4 for calculating adsorption of a gas mixture in a porous material.
Figure 12
Figure 12. Screenshots of Tool 5 for calculating diffusion of a gas mixture in a porous material.
Figure 13
Figure 13. Screenshot of Tool 6 for calculating the void fraction of a porous material.
Figure 14
Figure 14. Screenshot of Tool 7 for calculating the surface area of a porous material.
Figure 15
Figure 15. Screenshots of Tool 8 for calculating the radial distribution function of a gas.
Figure 16
Figure 16. Screenshot of Tool 9 for calculating the adsorption energy histogram.
Figure 17
Figure 17. Screenshot of Tool 10 for calculating the density and total energy of a gas in the NPT ensemble.
Figure 18
Figure 18. Screenshots of Tool 11 for calculating gas adsorption in a MOF using the Gibbs ensemble.
Figure 19
References
This article references 48 other publications.
- 1Madhavan, K.; Zentner, L.; Farnsworth, V.; Shivarajapura, S.; Zentner, M.; Denny, N.; Klimeck, G. nanoHUB.org: cloud-based services for nanoscale modeling, simulation, and education. Nanotechnol. Rev. 2013, 2, 107– 117, DOI: 10.1515/ntrev-2012-0043There is no corresponding record for this reference.
- 2Magana, A. J.; Brophy, S. P.; Bodner, G. M. An exploratory study of engineering and science students’ perceptions of nanoHUB.org simulations. Int. J. Eng. Educ. 2012, 28, 1019– 1032There is no corresponding record for this reference.
- 3Nanohub.org usage statistics. https://nanohub.org/usage, (accessed March 1, 2022).There is no corresponding record for this reference.
- 4Dubbeldam, D.; Calero, S.; Ellis, D. E.; Snurr, R. Q. RASPA: molecular simulation software for adsorption and diffusion in flexible nanoporous materials. Mol. Simul. 2016, 42, 81– 101, DOI: 10.1080/08927022.2015.10100824RASPA: molecular simulation software for adsorption and diffusion in flexible nanoporous materialsDubbeldam, David; Calero, Sofia; Ellis, Donald E.; Snurr, Randall Q.Molecular Simulation (2016), 42 (2), 81-101CODEN: MOSIEA; ISSN:0892-7022. (Taylor & Francis Ltd.)A new software package, RASPA, for simulating adsorption and diffusion of mols. in flexible nanoporous materials is presented. The code implements the latest state-of-the-art algorithms for mol. dynamics and Monte Carlo (MC) in various ensembles including symplectic/measure-preserving integrators, Ewald summation, configurational-bias MC, continuous fractional component MC, reactive MC and Baker's minimisation. We show example applications of RASPA in computing coexistence properties, adsorption isotherms for single and multiple components, self- and collective diffusivities, reaction systems and visualisation. The software is released under the GNU General Public License.
- 5Rogge, S. M. J.; Goeminne, R.; Demuynck, R.; Gutiérrez-Sevillano, J. J.; Vandenbrande, S.; Vanduyfhuys, L.; Waroquier, M.; Verstraelen, T.; Van Speybroeck, V. Modeling gas adsorption in flexible metal-organic frameworks via hybrid Monte Carlo/molecular dynamics schemes. Adv. Theory Simul. 2019, 2, 1800177 DOI: 10.1002/adts.2018001775Modeling Gas Adsorption in Flexible Metal-Organic Frameworks via Hybrid Monte Carlo/Molecular Dynamics SchemesRogge, Sven M. J.; Goeminne, Ruben; Demuynck, Ruben; Gutierrez-Sevillano, Juan Jose; Vandenbrande, Steven; Vanduyfhuys, Louis; Waroquier, Michel; Verstraelen, Toon; Van Speybroeck, VeroniqueAdvanced Theory and Simulations (2019), 2 (4), 1800177CODEN: ATSDCW; ISSN:2513-0390. (Wiley-VCH Verlag GmbH & Co. KGaA)Herein, a hybrid Monte Carlo (MC)/mol. dynamics (MD) simulation protocol that properly accounts for the extraordinary structural flexibility of metal-org. frameworks (MOFs) is developed and validated. This is vital to accurately predict gas adsorption isotherms and guest-induced flexibility of these materials. First, the performance of three recent models to predict adsorption isotherms and flexibility in MOFs is critically investigated. While these methods succeed in providing qual. insight in the gas adsorption process in MOFs, their accuracy remains limited as the intrinsic flexibility of these materials is very hard to account for. To overcome this challenge, a hybrid MC/MD simulation protocol that is specifically designed to handle the flexibility of the adsorbent, including the shape flexibility, is introduced, thereby unifying the strengths of the previous models. It is demonstrated that the application of this new protocol to the adsorption of neon, argon, xenon, methane, and carbon dioxide in MIL-53(Al), a prototypical flexible MOF, substantially decreases the inaccuracy of the obtained adsorption isotherms and predicted guest-induced flexibility. As a result, this method is ideally suited to rationalize the adsorption performance of flexible nanoporous materials at the mol. level, paving the way for the conscious design of MOFs as industrial adsorbents.
- 6Lin, L. C.; Lee, K.; Gagliardi, L.; Neaton, J. B.; Smit, B. Force-field development from electronic structure calculations with periodic boundary conditions: applications to gaseous adsorption and transport in metal-organic frameworks. J. Chem. Theory Comput. 2014, 10, 1477– 1488, DOI: 10.1021/ct500094w6Force-Field Development from Electronic Structure Calculations with Periodic Boundary Conditions: Applications to Gaseous Adsorption and Transport in Metal-Organic FrameworksLin, Li-Chiang; Lee, Kyuho; Gagliardi, Laura; Neaton, Jeffrey B.; Smit, BerendJournal of Chemical Theory and Computation (2014), 10 (4), 1477-1488CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)A systematic and efficient methodol. to derive accurate (nonpolarizable) force fields from periodic d. functional theory (DFT) calcns. for use in classical mol. simulations has been developed. The methodol. requires reduced computation cost compared with other conventional ways. The whole process is performed self-consistently in a fully periodic system. The force fields derived by using this methodol. accurately predict the CO2 and H2O adsorption isotherms inside Mg-MOF-74, and is transferable to Zn-MOF-74; by replacing the Mg-CO2 interactions with the corresponding Zn-CO2 interactions, we obtain an accurate prediction of the corresponding isotherm. This methodol. was used to address the effect of water on the sepn. of flue gases in these materials. In general, the mixt. isotherms of CO2 and H2O calcd. with these derived force fields show a significant redn. in CO2 uptake with the existence of trace amts. of water vapor. The effect of water, however, was found to be quant. different between Mg- and Zn-MOF-74.
- 7Dubbeldam, D.; Walton, K. S.; Vlugt, T. J. H.; Calero, S. Design, parameterization, and implementation of atomic force fields for adsorption in nanoporous materials. Adv. Theory Simul. 2019, 2, 1900135 DOI: 10.1002/adts.2019001357Design, Parameterization, and Implementation of Atomic Force Fields for Adsorption in Nanoporous MaterialsDubbeldam, David; Walton, Krista S.; Vlugt, Thijs J. H.; Calero, SofiaAdvanced Theory and Simulations (2019), 2 (11), 1900135CODEN: ATSDCW; ISSN:2513-0390. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Mol. simulations are an excellent tool to study adsorption and diffusion in nanoporous materials. Examples of nanoporous materials are zeolites, carbon nanotubes, clays, metal-org. frameworks (MOFs), covalent org. frameworks (COFs) and zeolitic imidazolate frameworks (ZIFs). The mol. confinement these materials offer has been exploited in adsorption and catalysis for almost 50 years. Mol. simulations have provided understanding of the underlying shape selectivity, and adsorption and diffusion effects. Much of the reliability of the modeling predictions depends on the accuracy and transferability of the force field. However, flexibility and the chem. and structural diversity of MOFs add significant challenges for engineering force fields that are able to reproduce exptl. obsd. structural and dynamic properties. Recent developments in design, parameterization, and implementation of force fields for MOFs and zeolites are reviewed.
- 8Kaplan, I. G. Intermolecular Interactions: Physical Picture, Computational Methods and Model Potentials; John Wiley & Sons: West Sussex, U.K., 2006.There is no corresponding record for this reference.
- 9Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.; Skiff, W. M. UFF, a full periodic-table force-field for molecular mechanics and molecular-dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024– 10035, DOI: 10.1021/ja00051a0409UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulationsRappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A., III; Skiff, W. M.Journal of the American Chemical Society (1992), 114 (25), 10024-35CODEN: JACSAT; ISSN:0002-7863.A new mol. mechanics force field, the Universal force field (UFF), is described wherein the force field parameters are estd. using general rules based only on the element, its hybridization and its connectivity. The force field functional forms, parameters, and generating formulas for the full periodic table are presented.
- 10Walters, E. T.; Mohebifar, M.; Johnson, E. R.; Rowley, C. N. Evaluating the London dispersion coefficients of protein force fields using the exchange-hole dipole moment model. J. Phys. Chem. B 2018, 122, 6690– 6701, DOI: 10.1021/acs.jpcb.8b0281410Evaluating the London Dispersion Coefficients of Protein Force Fields Using the Exchange-Hole Dipole Moment ModelWalters, Evan T.; Mohebifar, Mohamad; Johnson, Erin R.; Rowley, Christopher N.Journal of Physical Chemistry B (2018), 122 (26), 6690-6701CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)London dispersion is one of the fundamental interactions involved in protein folding and dynamics. The popular CHARMM36, Amber ff14sb, and OPLS-AA force fields represent these interactions through the C6/r6 term of the Lennard-Jones potential, where the C6 parameters are assigned empirically. Here, dispersion coeffs. of these three force fields are shown to be roughly 50% larger than values calcd. using the quantum-mech.-derived exchange-hole dipole moment (XDM) model. The CHARMM36 and Amber OL15 force fields for nucleic acids also exhibit this trend. The hydration energies of the side-chain models were calcd. using REMD-TI for the CHARMM36, Amber ff14sb, and OPLS-AA force fields. These force fields predict side-chain hydration energies that are in generally good agreement with the exptl. values, which suggests the total strength of aq. dispersion interactions is correct, despite C6 coeffs. that are considerably larger than XDM predicts. An anal. expression for the dispersion hydration energy using XDM coeffs. shows that higher-order dispersion terms (i.e., C8 and C10) account for roughly 37.5% of the hydration energy of methane. This suggests that the C6 dispersion coeffs. used in contemporary force fields are elevated to account for the neglected higher-order terms.
- 11Jones, J. E. On the determination of molecular fields - II From the equation of state of a gas. Proc. R. Soc. London, Ser. A 1924, 106, 463– 477, DOI: 10.1098/rspa.1924.008211The determination of molecular fields (II) From the equation of state of a gasJones, J. R.Proceedings of the Royal Society of London, Series A: Mathematical, Physical and Engineering Sciences (1924), 106 (), 463-77CODEN: PRLAAZ; ISSN:1364-5021.The virial coeffs. are evaluated for an equation of state of gases for the case of a mol. with attractive and repelling forces, as described above. The values obtained for the consts. of these forces for A agree fairly well with those calcd. from the viscosity data. In an appendix, recalcn. on the basis of more recent data on the isotherms of A give better agreement, pointing to 14 1/3 as the best value for the index of the repulsive force.
- 12Boda, D.; Henderson, D. The effects of deviations from Lorentz–Berthelot rules on the properties of a simple mixture. Mol. Phys. 2008, 106, 2367– 2370, DOI: 10.1080/0026897080247113712The effects of deviations from Lorentz-Berthelot rules on the properties of a simple mixtureBoda, Dezso; Henderson, DouglasMolecular Physics (2008), 106 (20), 2367-2370CODEN: MOPHAM; ISSN:0026-8976. (Taylor & Francis Ltd.)Generally, the parameters in the interaction potential between like mols. in a mixt. can be detd. in a relatively straightforward manner from the properties of the pure components. However, the detn. of the parameters in the interaction potential between the unlike pairs in the mixts. is more difficult. As a result, these parameters are usually estd. from avs. of the like parameters. The most common recipes are the Lorentz-Berthelot mixing rules, where the energy and mol. size parameters are presumed to be geometric and arithmetic avs., resp. There have been some studies of the consequences of deviations from the energy rule but almost no studies of the consequences of deviations from the size rule. Here, we study the effects of deviations from both rules on the radial distribution functions of a simple mixt. We find from simulations that, for this mixt., the effect of deviations from the energy rule on the radial distribution function are rather small but that the effect of deviations from the size rule can be significant and are interesting.
- 13Berthelot, D. Sur le mélange des gaz. Compt. Rendus 1898, 126, 1703– 170613On mixtures of gasesBerthelot, D.Comptes Rendus Hebdomadaires des Seances de l'Academie des Sciences (1898), 126 (), 1703CODEN: COREAF ISSN:.The author proposes a formula of the van der Waals type for a mixture of two gases and shows how the constants A and B for the mixture can be calculated from the corresponding constants for the single gases.
- 14Lorentz, H. A. Ueber die anwendung des satzes vom virial in der kinetischen theorie der gase. Ann. Phys. 1881, 248, 127– 136, DOI: 10.1002/andp.18812480110There is no corresponding record for this reference.
- 15Dubbeldam, D.; Calero, S.; Vlugt, T. J. H.; Ellis, D. E.; Snurr, R. Q. RASPA 2.0: Molecular Software Package for Adsorption and Diffusion in (Flexible) Nanoporous Materials , program manual, March 2015, pp 1− 145.There is no corresponding record for this reference.
- 16Dubbeldam, D.; Walton, K. S.; Ellis, D. E.; Snurr, R. Q. Exceptional negative thermal expansion in isoreticular metal-organic frameworks. Angew. Chem., Int. Ed. 2007, 46, 4496– 4499, DOI: 10.1002/anie.20070021816Exceptional negative thermal expansion in isoreticular metal-organic frameworksDubbeldam, David; Walton, Krista S.; Ellis, Donald E.; Snurr, Randall Q.Angewandte Chemie, International Edition (2007), 46 (24), 4496-4499CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A new model for flexible frameworks is used to simulate the structures and adsorption properties of isoreticular metal-org. frameworks (IRMOFs), such as IRMOF-1. The structural simulations suggest that the IRMOFs have neg. thermal-expansion coeffs. over their full temp. ranges of stability.
- 17Chen, B.; Siepmann, J. I. Transferable potentials for phase equilibria. 3. Explicit-hydrogen description of normal alkanes. J. Phys. Chem. B 1999, 103, 5370– 5379, DOI: 10.1021/jp990822m17Transferable potentials for phase equilibria. 3. Explicit-hydrogen description of normal alkanesChen, Bin; Siepmann, J. IljaJournal of Physical Chemistry B (1999), 103 (25), 5370-5379CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)Motivated by shortcomings of the available united-atom models for alkanes, a new explicit-H model for n-alkanes (TraPPE-EH, transferable potentials for phase equil.-explicit H) is developed from fitting to 1-component fluid-phase properties. In addn. to Lennard-Jones sites on C atoms, this model utilizes Lennard-Jones sites on the centers of C-H bonds. Configurational-bias Monte Carlo simulations in the Gibbs and canonical ensembles were carried out to calc. the 1-component vapor-liq. phase equil. for CH4 to n-dodecane, to det. the phase diagram of supercrit. C2H6 and n-heptane mixts., to obtain the Gibbs free energies of transfer for n-pentane (I) and n-hexane between He vapor and n-heptane liq. phases, and to study the high-pressure region of the equation of state for I and n-decane. The explicit-H representation, with its more faithful description of the mol. shape of alkanes, allows us to find a set of Lennard-Jones parameters that yields significantly better agreement with expt. for 1- and multicomponent phase equil. than our united-atom alkane model, but the price is higher computational cost.
- 18Eggimann, B. L.; Sunnarborg, A. J.; Stern, H. D.; Bliss, A. P.; Siepmann, J. I. An online parameter and property database for the TraPPE force field. Mol. Simul. 2014, 40, 101– 105, DOI: 10.1080/08927022.2013.84299418An online parameter and property database for the TraPPE force fieldEggimann, Becky L.; Sunnarborg, Amara J.; Stern, Hudson D.; Bliss, Andrew P.; Siepmann, J. IljaMolecular Simulation (2014), 40 (1-3), 101-105CODEN: MOSIEA; ISSN:0892-7022. (Taylor & Francis Ltd.)A review. The transferable potentials for phase equil. (TraPPE) force field aims to be accurate, computationally efficient and applicable to a wide range of chem. compds., state points and thermophys. properties. When new users wish to implement TraPPE models into their chosen simulation program, they face several obstacles: the TraPPE models are dispersed over many sep. publications and misinterpretations of the primary literature are possible; the TraPPE force field makes specific choices for std. conventions that may require non-trivial code modifications for some simulation software. Therefore, the TraPPE developers report here a resource website and online searchable parameter and property database designed to provide new and experienced users with tools for successful implementation and validation (http://www.chem.umn.edu/groups/siepmann/trappe/).
- 19Martin, M. G.; Siepmann, J. I. Transferable potentials for phase equilibria. 1. United-atom description of n-alkanes. J. Phys. Chem. B 1998, 102, 2569– 2577, DOI: 10.1021/jp972543+19Transferable Potentials for Phase Equilibria. 1. United-Atom Description of n-AlkanesMartin, Marcus G.; Siepmann, J. IljaJournal of Physical Chemistry B (1998), 102 (14), 2569-2577CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A new set of united-atom Lennard-Jones interaction parameters for n-alkanes is proposed from fitting to crit. temps. and satd. liq. densities. Configurational-bias Monte Carlo simulations in the Gibbs ensemble were carried out to det. the vapor-liq. coexistence curves for methane to dodecane using three united-atom force fields: OPLS [Jorgensen, et al. J. Am. Chem. Soc. 1984, 106, 813], SKS [Siepmann, et al. Nature 1993, 365, 330], and TraPPE. Std. specific densities and the high-pressure equation-of-state for the transferable potentials for phase equil. (TraPPE) model were studied by simulations in the isobaric-isothermal and canonical ensembles, resp. It is found that one set of Me and methylene parameters is sufficient to accurately describe the fluid phases of all n-alkanes with two or more carbon atoms. Whereas other n-alkane force fields employ Me groups that are either equal or larger in size than the methylene groups, it is demonstrated here that using a smaller Me group yields a better fit to the set of exptl. data. As should be expected from an effective pair potential, the new parameters do not reproduce exptl. second virial coeffs. Satd. vapor pressures and densities show small, but systematic deviation from the exptl. data.
- 20Potoff, J. J.; Siepmann, J. I. Vapor–liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogen. AIChE J. 2001, 47, 1676– 1682, DOI: 10.1002/aic.69047071920Vapor-liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogenPotoff, Jeffrey J.; Siepmann, J. IljaAIChE Journal (2001), 47 (7), 1676-1682CODEN: AICEAC; ISSN:0001-1541. (American Institute of Chemical Engineers)New force fields for carbon dioxide and nitrogen are introduced that quant. reproduce the vapor-liq. equil. (VLE) of the neat systems and their mixts. with alkanes. In addn. to the usual VLE calcns. for pure CO2 and N2, calcns. of the binary mixts. with propane were used in the force-field development to achieve a good balance between dispersive and electrostatic (quadrupole-quadrupole) interactions. The transferability of the force fields was then assessed from calcns. of the VLE for the binary mixts. with n-hexane, the binary mixt. of CO2/N2, and the ternary mixt. of CO2/N2/propane. The VLE calcns. were carried out using configurational-bias Monte Carlo simulations in either the grand canonical ensemble with histogram-reweighting or in the Gibbs ensemble.
- 21Darkrim, F.; Levesque, D. Monte Carlo simulations of hydrogen adsorption in single-walled carbon nanotubes. J. Chem. Phys. 1998, 109, 4981– 4984, DOI: 10.1063/1.47710921Monte Carlo simulations of hydrogen adsorption in single-walled carbon nanotubesDarkrim, Farida; Levesque, DominiqueJournal of Chemical Physics (1998), 109 (12), 4981-4984CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Within the framework of a study on the properties of carbon nanotubes, a promising new material, we performed numerical simulation of hydrogen adsorption at room temp. in single-walled nanotubes. The structure of this material is favorable to the adsorption phenomenon because of the narrow size distribution of the nanotube diams., which have dimensions on the order of the range of the carbon attractive interaction. We discuss the influence of the single-walled carbon nanotube diams. on the relative arrangement of carbon atoms and hydrogen mols. within an array of parallel single-walled carbon nanotubes. We also studied the influence on adsorption of the distance between the nearest-neighbor nanotubes.
- 22Zhang, L.; Siepmann, J. I. Direct calculation of Henry’s law constants from Gibbs ensemble Monte Carlo simulations: nitrogen, oxygen, carbon dioxide and methane in ethanol. Theor. Chem. Acc. 2006, 115, 391– 397, DOI: 10.1007/s00214-005-0073-122Direct calculation of Henry's law constants from Gibbs ensemble Monte Carlo simulations: nitrogen, oxygen, carbon dioxide and methane in ethanolZhang, Ling; Siepmann, J. IljaTheoretical Chemistry Accounts (2006), 115 (5), 391-397CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)Configurational-bias Monte Carlo simulations in the Gibbs ensemble were used to calc. Henry's law consts., Ostwald solubilities, and Gibbs free energies of transfer for oxygen, nitrogen, methane, and carbon dioxide in ethanol at 323 and 373 K. These three soly. descriptors can be expressed as functions of mech. properties that are directly observable in the Gibbs ensemble approach, thereby allowing for very precise detn. of the descriptors. Addnl., the Henry's law consts. of multiple solutes can be computed from a single simulation. Most of the simulations were carried out for systems contg. 1000 solvent and up to 8 solute mols., and further simulations using either 500 or 2000 solvent mols. point to negligible system size effects. A comparison with exptl. data shows that the united-atom version of the transferable potential for phase equil. force field yields Henry's law consts. that reproduce well the differences between the four solutes and the changes upon increase of the temp.
- 23Mayo, S. L.; Olafson, B. D.; Goddard, W. A. DREIDING - A generic force field for molecular simulations. J. Phys. Chem. 1990, 94, 8897– 8909, DOI: 10.1021/j100389a01023DREIDING: a generic force field for molecular simulationsMayo, Stephen L.; Olafson, Barry D.; Goddard, William A., IIIJournal of Physical Chemistry (1990), 94 (26), 8897-909CODEN: JPCHAX; ISSN:0022-3654.The parameters are given for a new generic force field, DREIDING, that is useful for predicting structures and dynamics of org., biol., and main-group inorg. mols. The philosophy in DREIDING is to use general force consts. and geometry parameters based on simple hybridization considerations rather than individual force consts. and geometric parameters that depend on the particular combination of atoms involved in the bond, angle, or torsion terms. Thus all bond distances are derived from at. radii, and there is only one force const. each for bonds, angles, and inversions and only six different values for torsional barriers. Parameters are defined for all possible combinations of atoms and new atoms can be added to the force field rather simply. The parameters are given for the "nonmetallic" main-group elements (B, C, N, O, F columns for the C, Si, Ge, and Sn rows) plus H and a few metals (Na, Ca, Zn, Fe). The accuracy of the DREIDING force field is tested by comparing with (i) 76 accurately detd. crystal structures of org. compds. involving H, C, N, O, F, P, S, Cl, and Br, (ii) rotational barriers of a no. of mols., and (iii) relative conformational energies and barriers of a no. of mols. There is excellent agreement.
- 24Bai, P.; Tsapatsis, M.; Siepmann, J. I. TraPPE-zeo: Transferable potentials for phase equilibria force field for all-silica zeolites. J. Phys. Chem. C 2013, 117, 24375– 24387, DOI: 10.1021/jp407422424TraPPE-zeo: Transferable Potentials for Phase Equilibria Force Field for All-Silica ZeolitesBai, Peng; Tsapatsis, Michael; Siepmann, J. IljaJournal of Physical Chemistry C (2013), 117 (46), 24375-24387CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)The transferable potentials for phase equil. (TraPPE) force field is extended to all-silica zeolites. This novel force field is parametrized to match the exptl. adsorption isotherms of n-heptane, propane, carbon dioxide, and ethanol with the Lennard-Jones parameters for sorbate-framework interactions detd. in a consistent manner using the Lorentz-Berthelot combining rules as for other parts of the TraPPE force field. The TraPPE-zeo force field allows for accurate predictions for both adsorption and diffusion of alkanes, alcs., carbon dioxide, and water over a wide range of pressures and temps. To achieve transferability to a wider range of mol. types, ranging from nonpolar to dipolar and hydrogen-bonding compds., Lennard-Jones interaction sites and partial charges are placed at both the oxygen and the silicon atoms of the zeolite lattice, which allows for a better balance of dispersive and 1st-order electrostatic interactions than is achievable with the Lennard-Jones potential used only for the oxygen atoms. The use of the Lorentz-Berthelot combining rules for unlike interactions makes the TraPPE-zeo force field applicable to any sorbate as long as the relevant TraPPE sorbate-sorbate parameters are available. The TraPPE-zeo force field allows for greatly improved predictive power compared to force fields that explicitly tabulate the individual cross-interaction parameters.
- 25Dubbeldam, D.; Torres-Knoop, A.; Walton, K. S. On the inner workings of Monte Carlo codes. Mol. Simul. 2013, 39, 1253– 1292, DOI: 10.1080/08927022.2013.81910225On the inner workings of Monte Carlo codesDubbeldam, David; Torres-Knoop, Ariana; Walton, Krista S.Molecular Simulation (2013), 39 (14-15), 1253-1292CODEN: MOSIEA; ISSN:0892-7022. (Taylor & Francis Ltd.)We review state-of-the-art Monte Carlo (MC) techniques for computing fluid coexistence properties (Gibbs simulations) and adsorption simulations in nanoporous materials such as zeolites and metal-org. frameworks. Conventional MC is discussed and compared to advanced techniques such as reactive MC, configurational-bias Monte Carlo and continuous fractional MC. The latter technique overcomes the problem of low insertion probabilities in open systems. Other modern methods are (hyper-)parallel tempering, Wang-Landau sampling and nested sampling. Details on the techniques and acceptance rules as well as to what systems these techniques can be applied are provided. We highlight consistency tests to help validate and debug MC codes.
- 26Heinen, J.; Burtch, N. C.; Walton, K. S.; Dubbeldam, D. Flexible force field parameterization through fitting on the ab initio-derived elastic tensor. J. Chem. Theory Comput. 2017, 13, 3722– 3730, DOI: 10.1021/acs.jctc.7b0031026Flexible Force Field Parameterization through Fitting on the Ab Initio-Derived Elastic TensorHeinen, Jurn; Burtch, Nicholas C.; Walton, Krista S.; Dubbeldam, DavidJournal of Chemical Theory and Computation (2017), 13 (8), 3722-3730CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Constructing functional forms and their corresponding force field parameters for the metal-linker interface of metal-org. frameworks is challenging. We propose to fit these parameters on the elastic tensor, computed from ab initio d. functional theory calcns. The advantage of this top-down approach is that it becomes evident if functional forms are missing when components of the elastic tensor are off. As a proof-of-concept, a new flexible force field for MIL-47(V) is derived. Neg. thermal expansion is obsd. and framework flexibility has a negligible effect on adsorption and transport properties for small guest mols. We believe this force field parameterization approach can serve as a useful tool for developing accurate flexible force field models that capture the correct mech. behavior of the full periodic structure.
- 27Heinen, J.; Dubbeldam, D. On flexible force fields for metal-organic frameworks: Recent developments and future prospects. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2018, 8, e1363 DOI: 10.1002/wcms.1363There is no corresponding record for this reference.
- 28Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids, 2nd ed.; Oxford University Press: Oxford, U.K., 2017.There is no corresponding record for this reference.
- 29Frenkel, D.; Smit, B. Understanding Molecular Simulation: From Algorithms to Applications, 2nd ed.; Academic Press: San Diego, California, 2002.There is no corresponding record for this reference.
- 30Hastings, W. K. Monte Carlo sampling methods using Markov chains and their applications. Biometrika 1970, 57, 97– 109, DOI: 10.1093/biomet/57.1.97There is no corresponding record for this reference.
- 31Metropolis, N.; Rosenbluth, A. W.; Rosenbluth, M. N.; Teller, A. H.; Teller, E. Equation of state calculations by fast computing machines. J. Chem. Phys. 1953, 21, 1087– 1092, DOI: 10.1063/1.169911431Equation-of-state calculations by fast computing machinesMetropolis, Nicholas; Rosenbluth, Arianna W.; Rosenbluth, Marshall N.; Teller, Augusta H.; Teller, EdwardJournal of Chemical Physics (1953), 21 (), 1087-92CODEN: JCPSA6; ISSN:0021-9606.A general method, suitable for fast computing machines, is described for investigating such properties as equations of state for substances consisting of interacting individual mols. The method consists of a modified Monte Carlo integration over configuration space. Results for the 2-dimensional rigid-sphere system were obtained on the Los Alamos MANIAC. These results were compared to the free-vol. equation of state and to a 4-term virial coeff. expansion.
- 32Hirotani, A.; Mizukami, K.; Miura, R.; Takaba, H.; Miya, T.; Fahmi, A.; Stirling, A.; Kubo, M.; Miyamoto, A. Grand canonical Monte Carlo simulation of the adsorption of CO2 on silicalite and NaZSM-5. Appl. Surf. Sci. 1997, 120, 81– 84, DOI: 10.1016/S0169-4332(97)00222-532Grand canonical Monte Carlo simulation of the adsorption of CO2 on silicalite and NaZSM-5Hirotani, Akiyasu; Mizukami, Koichi; Miura, Ryuji; Takaba, Hiromitsu; Miya, Takeshi; Fahmi, Adil; Stirling, Andras; Kubo, Momoji; Miyamoto, AkiraApplied Surface Science (1997), 120 (1/2), 81-84CODEN: ASUSEE; ISSN:0169-4332. (Elsevier)The adsorption of carbon dioxide in silicalite and NaZSM-5 zeolite has been studied using new Monte Carlo software. In this program, sodium cations and framework are movable during the simulation. The calcd. adsorption isotherms are in good agreement with the exptl. results. The energy distribution of carbon dioxide over silicalite and NaZSM-5 shows that the increase of the adsorption energy for NaZSM-5 is mainly due the elec. field generated by sodium cations.
- 33Gao, G.; Wang, W. Gibbs Ensemble Monte Carlo simulation of binary vapor-liquid equilibria for CFC alternatives. Fluid Phase Equilib. 1997, 130, 157– 166, DOI: 10.1016/S0378-3812(96)03195-033Gibbs Ensemble Monte Carlo simulation of binary vapor-liquid equilibria for CFC alternativesGao, Guangtu; Wang, WenchuanFluid Phase Equilibria (1997), 130 (1-2), 157-166CODEN: FPEQDT; ISSN:0378-3812. (Elsevier)The Gibbs Ensemble Monte Carlo (GEMC) simulation method has been used for vapor-liq. equil. calcns. of the binary systems R22-R142b and R22-R152a, which are considered to be promising mixt. refrigerants for replacing R12. All the mols. have been treated in terms of an effective Lennard-Jones potential, with temp. dependent parameters regressed by fitting vapor pressures and satd. liq. densities at various temps. for pure substances of interest. Good agreement between exptl. and calcd. data from the vol.-translation Peng-Robinson equation of state and the simulated results, including the compns., molar volumes for both the vapor and liq. phases, and heats of vaporization, indicates that the GEMC method can be applied to the description of phase behavior for these CFC alternative binary systems.
- 34Slepoy, A.; Thompson, A. P.; Plimpton, S. J. A constant-time kinetic Monte Carlo algorithm for simulation of large biochemical reaction networks. J. Chem. Phys. 2008, 20, 205101 DOI: 10.1063/1.2919546There is no corresponding record for this reference.
- 35Dove, M. T. An introduction to atomistic simulation methods. Seminarios de la SEM 2008, 4, 7– 37There is no corresponding record for this reference.
- 36Braun, E.; Gilmer, J.; Mayes, H. B.; Mobley, D. L.; Monroe, J. I.; Prasad, S.; Zuckerman, D. M. Best practices for foundations in molecular simulations. Living J. Comput. Mol. Sci. 2019, 1, 5957 DOI: 10.33011/livecoms.1.1.5957There is no corresponding record for this reference.
- 37Abouelnasr, M. K. F.; Smit, B. Diffusion in confinement: kinetic simulations of self- and collective diffusion behavior of adsorbed gases. Phys. Chem. Chem. Phys. 2012, 14, 11600– 11609, DOI: 10.1039/c2cp41147d37Diffusion in confinement: kinetic simulations of self- and collective diffusion behavior of adsorbed gasesAbouelnasr, Mahmoud K. F.; Smit, BerendPhysical Chemistry Chemical Physics (2012), 14 (33), 11600-11609CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The self- and collective-diffusion behaviors of adsorbed methane, helium, and isobutane in zeolite frameworks LTA, MFI, AFI, and SAS were examd. at various concns. using a range of mol. simulation techniques including Mol. Dynamics (MD), Monte Carlo (MC), Bennett-Chandler (BC), and kinetic Monte Carlo (kMC). This paper has three main results. (1) A novel model for the process of adsorbate movement between two large cages was created, allowing the formulation of a mixing rule for the re-crossing coeff. between two cages of unequal loading. The predictions from this mixing rule were found to agree quant. with explicit simulations. (2) A new approach to the dynamically cor. Transition State Theory method to anal. calc. self-diffusion properties was developed, explicitly accounting for nanoscale fluctuations in concn. This approach was demonstrated to quant. agree with previous methods, but is uniquely suited to be adapted to a kMC simulation that can simulate the collective-diffusion behavior. (3) While at low and moderate loadings the self- and collective-diffusion behaviors in LTA are obsd. to coincide, at higher concns. they diverge. A change in the adsorbate packing scheme was shown to cause this divergence, a trait which is replicated in a kMC simulation that explicitly models this behavior. These phenomena were further investigated for isobutane in zeolite MFI, where MD results showed a sepn. in self- and collective- diffusion behavior that was reproduced with kMC simulations.
- 38Skoulidas, A. I.; Sholl, D. S. Self-diffusion and transport diffusion of light gases in metal-organic framework materials assessed using molecular dynamics simulations. J. Phys. Chem. B 2005, 109, 15760– 15768, DOI: 10.1021/jp051771y38Self-Diffusion and Transport Diffusion of Light Gases in Metal-Organic Framework Materials Assessed Using Molecular Dynamics SimulationsSkoulidas, Anastasios I.; Sholl, David S.Journal of Physical Chemistry B (2005), 109 (33), 15760-15768CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Metal-org. framework (MOF) materials pose an interesting alternative to more traditional nanoporous materials for a variety of sepn. processes. Sepn. processes involving nanoporous materials can be controlled by either adsorption equil., diffusive transport rates, or a combination of these factors. Adsorption equil. was studied for a variety of gases in MOFs, but almost nothing is currently known about mol. diffusion rates in MOFs. The authors used equil. mol. dynamics (MD) to probe the self-diffusion and transport diffusion of a no. of small gas species in several MOFs as a function of pore loading at room temp. Specifically, the authors have studied Ar, CH4, CO2, N2, and H2 diffusion in MOF-5. The diffusion of Ar in MOF-2, MOF-3, and Cu-BTC was assessed in a similar manner. Results greatly expand the range of MOFs for which data describing mol. diffusion is available. The authors discuss the prospects for exploiting mol. transport properties in MOFs in practical sepn. processes and the future role of MD simulations in screening families of MOFs for these processes.
- 39Werder, T.; Walther, J. H.; Jaffe, R. L.; Halicioglu, T.; Koumoutsakos, P. On the water-carbon interaction for use in molecular dynamics simulations of graphite and carbon nanotubes. J. Phys. Chem. B 2003, 107, 1345– 1352, DOI: 10.1021/jp026811239On the Water-Carbon Interaction for Use in Molecular Dynamics Simulations of Graphite and Carbon NanotubesWerder, T.; Walther, J. H.; Jaffe, R. L.; Halicioglu, T.; Koumoutsakos, P.Journal of Physical Chemistry B (2003), 107 (6), 1345-1352CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)A systematic mol. dynamics study shows that the contact angle of a H2O droplet on graphite changes significantly as a function of the H2O-C interaction energy. Together with the observation that a linear relation can be established between the contact angle and the H2O monomer binding energy on graphite, a new route to calibrate interaction potential parameters is presented. Through a variation of the droplet size at 1000-17,500 H2O mols., the authors det. the line tension to be pos. and ∼2 × 10-10 J/m. To recover a macroscopic contact angle of 86°, a H2O monomer binding energy of -6.33 kJ mol-1 is required, which is obtained by applying a C-oxygen Lennard-Jones potential with the parameters εCO = 0.392 kJ mol-1 and σCO = 3.19 Å. For this new H2O-C interaction potential, the authors present d. profiles and hydrogen bond distributions for a H2O droplet on graphite.
- 40Henry, A. S.; Chen, G. Spectral phonon transport properties of silicon based on molecular dynamics simulations and lattice dynamics. J. Comput. Theor. Nanosci. 2008, 5, 141– 152, DOI: 10.1166/jctn.2008.245440Spectral phonon transport properties of silicon based on molecular dynamics simulations and lattice dynamicsHenry, Asegun S.; Chen, GangJournal of Computational and Theoretical Nanoscience (2008), 5 (2), 141-152CODEN: JCTNAB; ISSN:1546-1955. (American Scientific Publishers)Although the thermal cond. of silicon has been studied before, current estns. for the phonon mean free paths have not provided full explanation of the strong size effects exptl. obsd. for various silicon micro and nanostructures. Since phonon relaxation time models are mostly semi-empirical, the mean free paths cannot be detd. reliably and questions remain as to which polarizations, frequencies, and wavelengths are dominant heat carriers. Here we have used a combination of equil. mol. dynamics simulations and lattice dynamics calcns. to fully detail the spectral dependence of phonon transport properties in bulk silicon. By considering the frequency dependence of the sp. heat, group velocities, and mean free paths, we address these unresolved questions and examine the errors assocd. with isotropic and frequency averaged approxns. Simulation details, such as the convergence of results on the simulation time and extn. of phonon transport properties in different crystallog. directions, are also discussed.
- 41Karplus, M.; McCammon, J. A. Molecular dynamics simulations of biomolecules. Nat. Struct. Biol. 2002, 9, 646– 652, DOI: 10.1038/nsb0902-64641Molecular dynamics simulations of biomoleculesKarplus, Martin; McCammon, J. AndrewNature Structural Biology (2002), 9 (9), 646-652CODEN: NSBIEW; ISSN:1072-8368. (Nature Publishing Group)A review. Mol. dynamics simulations are important tools for understanding the phys. basis of the structure and function of biol. macromols. The early view of proteins as relatively rigid structures has been replaced by a dynamic model in which the internal motions and resulting conformational changes play an essential role in their function. This review presents a brief description of the origin and early uses of biomol. simulations. It then outlines some recent studies that illustrate the utility of such simulations and closes with a discussion of their ever-increasing potential for contributing to biol.
- 42Umeh, J. C. Development of Simulation Tools on Nanuhub as Learning Modules. Master’s thesis; New Mexico State University: Las Cruces, NM, 2019.There is no corresponding record for this reference.
- 43Humbert, M. T.; Zhang, Y.; Maginn, E. J. PyLAT: Python LAMMPS analysis tools. J. Chem. Inf. Model. 2019, 59, 1301– 1305, DOI: 10.1021/acs.jcim.9b0006643PyLAT: Python LAMMPS Analysis ToolsHumbert, Michael T.; Zhang, Yong; Maginn, Edward J.Journal of Chemical Information and Modeling (2019), 59 (4), 1301-1305CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Reproducibility and accuracy have become increasingly important issues for the mol. simulation community. The availability of validated open-source postprocessing tools to analyze simulation trajectories and compute properties is key to helping researchers conduct more accurate and reproducible simulations. Here the authors describe a suite of open-source Python-based postprocessing routines the authors have developed called PyLAT. PyLAT is compatible with the popular mol. dynamics package LAMMPS and enables users to compute viscosities, self-diffusivities, ionic conductivities, mol. or ion pair lifetimes, dielec. consts., and radial distribution functions using best-practice methods.
- 44Rowsell, J. L. C.; Yaghi, O. M. Effects of functionalization, catenation, and variation of the metal oxide and organic linking units on the low-pressure hydrogen adsorption properties of metal-organic frameworks. J. Am. Chem. Soc. 2006, 128, 1304– 1315, DOI: 10.1021/ja056639q44Effects of Functionalization, Catenation, and Variation of the Metal Oxide and Organic Linking Units on the Low-Pressure Hydrogen Adsorption Properties of Metal-Organic FrameworksRowsell, Jesse L. C.; Yaghi, Omar M.Journal of the American Chemical Society (2006), 128 (4), 1304-1315CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The dihydrogen adsorption isotherms of eight metal-org. frameworks (MOFs), measured at 77 K up to a pressure of 1 atm, were examd. for correlations with their structural features. All materials display approx. Type I isotherms with no hysteresis, and satn. was not reached for any of the materials under these conditions. Among the six iso-reticular MOFs (IRMOFs) studied, the catenated materials exhibit the largest capacities on a molar basis, up to 9.8 H2 per formula unit. The addn. of functional groups (-Br, -NH2, -C2H4-) to the phenylene links of IRMOF-1 (MOF-5), or their replacement with thieno[3,2-b]thiophene moieties in IRMOF-20, altered the adsorption behavior by a minor amt. despite large variations in the pore vols. of the resulting materials. But replacement of the metal oxide units with those contg. coordinatively unsatd. metal sites resulted in greater H2 uptake. The enhanced affinities of these materials, MOF-74 and HKUST-1, were further demonstrated by calcn. of the isosteric heats of adsorption, which were larger across much of the range of coverage examd., compared to those of representative IRMOFs. Probably under low-loading conditions, the H2 adsorption behavior of MOFs can be improved by imparting larger charge gradients on the metal oxide units and adjusting the link metrics to constrict the pore dimensions; however, a large pore vol. is still a prerequisite feature.
- 45Siberio-Perez, D. Y.; Wong-Foy, A. G.; Yaghi, O. M.; Matzger, A. J. Raman spectroscopic investigation of CH4 and N2 adsorption in metal-organic frameworks. Chem. Mater. 2007, 19, 3681– 3685, DOI: 10.1021/cm070542g45Raman Spectroscopic Investigation of CH4 and N2 Adsorption in Metal-Organic FrameworksSiberio-Perez, Diana Y.; Wong-Foy, Antek G.; Yaghi, Omar M.; Matzger, Adam J.Chemistry of Materials (2007), 19 (15), 3681-3685CODEN: CMATEX; ISSN:0897-4756. (American Chemical Society)The adsorption behavior of CH4 and N2 (298 K, 30 bar) in a series of isoreticular metal-org. frameworks (IRMOFs) was investigated by Raman spectroscopy. For CH4, the ν1 vibrational mode shifted to lower frequency by 7.6, 8.4, 11.0, 10.3, and 10.1 cm-1 from 2917 cm-1 when adsorbed to IRMOF-1, -6, -8, -11, and -18, resp. Along this same series, the adsorbed N2 stretch exhibited smaller shifts of 2.7, 3.1, 4.2, 4.1, and 3.7 cm-1. These shifts arise because of interactions within the framework pores, and not with the outer crystal surface. In all cases, Raman spectra at pressures up to 30 bar showed that satn. of the sorption sites does not occur. The obsd. shifts of the vibrational modes for each gas indicate different chem. environments within different IRMOFs, pointing to the important role the linkers play in the adsorption of gases.
- 46Duren, T.; Sarkisov, L.; Yaghi, O. M.; Snurr, R. Q. Design of new materials for methane storage. Langmuir 2004, 20, 2683– 2689, DOI: 10.1021/la035550046Design of new materials for methane storageDuren Tina; Sarkisov Lev; Yaghi Omar M; Snurr Randall QLangmuir : the ACS journal of surfaces and colloids (2004), 20 (7), 2683-9 ISSN:0743-7463.One of the strategic goals of the modern automobile manufacturing industry is to replace gasoline and diesel with alternative fuels such as natural gas. In this report, we elucidate the desired characteristics of an optimal adsorbent for gas storage. The U.S. Department of Energy has outlined several requirements that adsorbents must fulfill for natural gas to become economically viable, with a key criterion being the amount adsorbed at 35 bar. We explore the adsorption characteristics of novel metal-organic materials (IRMOFs and molecular squares) and contrast them with the characteristics of two zeolites, MCM-41, and different carbon nanotubes. Using molecular simulations, we uncover the complex interplay of the factors influencing methane adsorption, especially the surface area, the capacity or free volume, the strength of the energetic interaction, and the pore size distribution. We also explain the extraordinary adsorption properties of IRMOF materials and propose new, not yet synthesized IRMOF structures with adsorption characteristics that are predicted to exceed the best experimental results to date by up to 36%.
- 47Walton, K. S.; Millward, A. R.; Dubbeldam, D.; Frost, H.; Low, J. J.; Yaghi, O. M.; Snurr, R. Q. Understanding inflection and steps in carbon dioxide adsorption isotherms in metal-organic frameworks. J. Am. Chem. Soc. 2007, 130, 406– 407, DOI: 10.1021/ja076595gThere is no corresponding record for this reference.
- 48Towns, J.; Cockerill, T.; Dahan, M.; Foster, I.; Gaither, K.; Grimshaw, A.; Hazlewood, V.; Lathrop, S.; Lifka, D.; Peterson, G. D.; Roskies, R.; Scott, J. R.; Wilkins-Diehr, N. XSEDE: accelerating scientific discovery. Comput. Sci. Eng. 2014, 16, 62– 74, DOI: 10.1109/MCSE.2014.80There is no corresponding record for this reference.