Skip to main content
Log in

Numerical investigation of microscale dynamic contact angles of the CO2–water–silica system using coarse-grained molecular approach

  • Original Paper
  • Published:
Computational Mechanics Aims and scope Submit manuscript

Abstract

The dynamic contact angle of a gas–liquid–solid system depends on the contact line velocity and ignoring this effect could lead to inaccurate estimations of the capillary pressures in microporous media. While most existing coarse-grained molecular dynamics (CGMD) models use one particle to represent a few molecules, we present a novel CGMD framework to model microscale CO2/water flows in silica with each particle representing hundreds of thousands of molecules. The framework can reproduce the densities and viscosities of water and CO2, water–CO2 interfacial tension, and static contact angle over a wide range of pressures. The validated framework is applied to study the velocity-dependency of contact angle of the microscale CO2–water–silica system. The results indicate that the assumption in the molecular kinetic theory that liquid–solid interaction is similar to the reversible work of adhesion between liquid and solid may not hold for CO2–water–silica systems.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  1. Chukwudeme E, Hamouda A (2009) Enhanced oil recovery (EOR) by miscible CO2 and water flooding of asphaltenic and non-asphaltenic oils. Energies 2(3):714–737

    Google Scholar 

  2. Middleton RS, Carey JW, Currier RP, Hyman JD, Kang Q, Karra S, Jiménez-Martínez J, Porter ML, Viswanathan HS (2015) Shale gas and non-aqueous fracturing fluids: opportunities and challenges for supercritical CO2. Appl Energy 147:500–509

    Google Scholar 

  3. Koh DY, Kang H, Kim DO, Park J, Cha M, Lee H (2012) Recovery of methane from gas hydrates intercalated within natural sediments using CO2 and a CO2/N2 gas mixture. Chemsuschem 5(8):1443–1448

    Google Scholar 

  4. Jung J-W, Wan J (2012) Supercritical CO2 and ionic strength effects on wettability of silica surfaces: equilibrium contact angle measurements. Energy Fuels 26(9):6053–6059

    Google Scholar 

  5. Sarmadivaleh M, Al-Yaseri AZ, Iglauer S (2015) Influence of temperature and pressure on quartz–water–CO2 contact angle and CO2–water interfacial tension. J Colloid Interface Sci 441:59–64

    Google Scholar 

  6. Espinoza DN, Santamarina JC (2010) Water–CO2–mineral systems: interfacial tension, contact angle, and diffusion—implications to CO2 geological storage. Water Resour Res 46(7):W07537

    Google Scholar 

  7. Chen C, Dong B, Zhang N, Li W, Song Y (2016) Pressure and temperature dependence of contact angles for CO2/water/silica systems predicted by molecular dynamics simulations. Energy Fuels 30(6):5027–5034

    Google Scholar 

  8. Iglauer S, Mathew M, Bresme F (2012) Molecular dynamics computations of brine–CO2 interfacial tensions and brine–CO2–quartz contact angles and their effects on structural and residual trapping mechanisms in carbon geo-sequestration. J Colloid Interface Sci 386(1):405–414

    Google Scholar 

  9. Javanbakht G, Sedghi M, Welch W, Goual L (2015) Molecular dynamics simulations of CO2/water/quartz interfacial properties: impact of CO2 dissolution in water. Langmuir 31(21):5812–5819

    Google Scholar 

  10. Chen C, Zhang N, Li W, Song Y (2015) Water contact angle dependence with hydroxyl functional groups on silica surfaces under CO2 sequestration conditions. Environ Sci Technol 49(24):14680–14687

    Google Scholar 

  11. Iglauer S, Salamah A, Sarmadivaleh M, Liu K, Phan C (2014) Contamination of silica surfaces: impact on water–CO2–quartz and glass contact angle measurements. Int J Greenhouse Gas Control 22:325–328

    Google Scholar 

  12. Huang P, Shen L, Gan Y, Maggi F, El-Zein A, Pan Z (2019) Atomistic study of dynamic contact angles in CO2–water–silica system. Langmuir 35(15):5324–5332

    Google Scholar 

  13. Friedman SP (1999) Dynamic contact angle explanation of flow rate-dependent saturation–pressure relationships during transient liquid flow in unsaturated porous media. J Adhes Sci Technol 13(12):1495–1518

    Google Scholar 

  14. Liu H, Ju Y, Wang N, Xi G, Zhang Y (2015) Lattice Boltzmann modeling of contact angle and its hysteresis in two-phase flow with large viscosity difference. Phys Rev E 92(3):033306

    MathSciNet  Google Scholar 

  15. Li L, Shen L, Nguyen GD, El-Zein A, Maggi F (2018) A smoothed particle hydrodynamics framework for modelling multiphase interactions at meso-scale. Comput Mech 62(5):1071–1085

    MathSciNet  MATH  Google Scholar 

  16. Bao Y, Li L, Shen L, Lei C, Gan Y (2019) Modified smoothed particle hydrodynamics approach for modelling dynamic contact angle hysteresis. Acta Mech Sin 35(3):472–485

    MathSciNet  Google Scholar 

  17. Shadloo M, Zainali A, Yildiz M (2013) Simulation of single mode Rayleigh–Taylor instability by SPH method. Comput Mech 51(5):699–715

    MathSciNet  MATH  Google Scholar 

  18. Saha AA, Mitra SK (2009) Effect of dynamic contact angle in a volume of fluid (VOF) model for a microfluidic capillary flow. J Colloid Interface Sci 339(2):461–480

    Google Scholar 

  19. Li S, Fan H (2015) On multiscale moving contact line theory. Proc R Soc A 471(2179):20150224

    MathSciNet  MATH  Google Scholar 

  20. Minaki H, Li S (2014) Multiscale modeling and simulation of dynamic wetting. Comput Methods Appl Mech Eng 273:273–302

    MathSciNet  MATH  Google Scholar 

  21. Cook BK, Noble DR, Williams JR (2004) A direct simulation method for particle-fluid systems. Eng Comput 21(2–4):151–168

    MATH  Google Scholar 

  22. Jing L, Kwok C, Leung Y, Sobral Y (2016) Extended CFD–DEM for free-surface flow with multi-size granules. Int J Numer Anal Methods Geomech 40(1):62–79

    Google Scholar 

  23. Das R, Cleary PW (2015) Evaluation of accuracy and stability of the classical SPH method under uniaxial compression. J Sci Comput 64(3):858–897

    MathSciNet  MATH  Google Scholar 

  24. Avendano C, Lafitte T, Galindo A, Adjiman CS, Jackson G, Müller EA (2011) SAFT-γ force field for the simulation of molecular fluids. 1. A single-site coarse grained model of carbon dioxide. J Phys Chem B 115(38):11154–11169

    Google Scholar 

  25. Lobanova O, Avendaño C, Lafitte T, Müller EA, Jackson G (2015) SAFT-γ force field for the simulation of molecular fluids: 4. A single-site coarse-grained model of water applicable over a wide temperature range. Mol Phys 113(9–10):1228–1249

    Google Scholar 

  26. Marrink SJ, De Vries AH, Mark AE (2004) Coarse grained model for semiquantitative lipid simulations. J Phys Chem B 108(2):750–760

    Google Scholar 

  27. Chiu S-W, Scott HL, Jakobsson E (2010) A coarse-grained model based on Morse potential for water and n-alkanes. J Chem Theory Comput 6(3):851–863

    Google Scholar 

  28. Groot RD, Warren PB (1997) Dissipative particle dynamics: bridging the gap between atomistic and mesoscopic simulation. J Chem Phys 107(11):4423–4435

    Google Scholar 

  29. Warren P (2003) Vapor–liquid coexistence in many-body dissipative particle dynamics. Phys Rev E 68(6):066702

    Google Scholar 

  30. Pagonabarraga I, Frenkel D (2001) Dissipative particle dynamics for interacting systems. J Chem Phys 115(11):5015–5026

    Google Scholar 

  31. Arienti M, Pan W, Li X, Karniadakis G (2011) Many-body dissipative particle dynamics simulation of liquid/vapor and liquid/solid interactions. J Chem Phys 134(20):204114

    Google Scholar 

  32. Kumar A, Asako Y, Abu-Nada E, Krafczyk M, Faghri M (2009) From dissipative particle dynamics scales to physical scales: a coarse-graining study for water flow in microchannel. Microfluid Nanofluidics 7(4):467

    Google Scholar 

  33. Espanol P, Revenga M (2003) Smoothed dissipative particle dynamics. Phys Rev E 67(2):026705

    Google Scholar 

  34. Ellero M, Español P (2018) Everything you always wanted to know about SDPD⋆(⋆ but were afraid to ask). Appl Math Mech 39(1):103–124

    MathSciNet  Google Scholar 

  35. Hu XY, Adams NA (2006) A multi-phase SPH method for macroscopic and mesoscopic flows. J Comput Phys 213(2):844–861

    MathSciNet  MATH  Google Scholar 

  36. Lei H, Baker NA, Wu L, Schenter GK, Mundy CJ, Tartakovsky AM (2016) Smoothed dissipative particle dynamics model for mesoscopic multiphase flows in the presence of thermal fluctuations. Phys Rev E 94(2):023304

    Google Scholar 

  37. Huang P, Shen L, Gan Y, Nguyen GD, El-Zein A, Maggi F (2018) Coarse-grained modeling of multiphase interactions at microscale. J Chem Phys 149(12):124505

    Google Scholar 

  38. Ruckenstein E, Liu H (1997) Self-diffusion in gases and liquids. Ind Eng Chem Res 36(9):3927–3936

    Google Scholar 

  39. Liu H, Silva CM, Macedo EA (1997) New equations for tracer diffusion coefficients of solutes in supercritical and liquid solvents based on the Lennard–Jones fluid model. Ind Eng Chem Res 36(1):246–252

    Google Scholar 

  40. Lemmon EW, McLinden MO, Friend DG. Thermophysical properties of fluid systems. In: Linstrom PJ, Mallard WG (eds) NIST Chemistry WebBook, NIST Standard Reference Database Number 69. National Institute of Standards and Technology, Gaithersburg MD, 20899. https://doi.org/10.18434/t4d303

  41. Shinoda W, Shiga M, Mikami M (2004) Rapid estimation of elastic constants by molecular dynamics simulation under constant stress. Phys Rev B 69(13):134103

    Google Scholar 

  42. Allen MP, Tildesley DJ (1987) Computer simulation of liquids. Oxford University Press, New York

    MATH  Google Scholar 

  43. Georgiadis A, Maitland G, Trusler JM, Bismarck A (2010) Interfacial tension measurements of the (H2O + CO2) system at elevated pressures and temperatures. J Chem Eng Data 55(10):4168–4175

    Google Scholar 

  44. Groot RD, Rabone K (2001) Mesoscopic simulation of cell membrane damage, morphology change and rupture by nonionic surfactants. Biophys J 81(2):725–736

    Google Scholar 

  45. Ghoufi A, Malfreyt P (2011) Mesoscale modeling of the water liquid–vapor interface: a surface tension calculation. Phys Rev E 83(5):051601

    Google Scholar 

  46. Farokhpoor R, Bjørkvik BJ, Lindeberg E, Torsæter O (2013) Wettability behaviour of CO2 at storage conditions. Int J Greenhouse Gas Control 12:18–25

    Google Scholar 

  47. Li X, Fan X (2015) Effect of CO2 phase on contact angle in oil-wet and water-wet pores. Int J Greenhouse Gas Control 36:106–113

    Google Scholar 

  48. Plimpton S (1995) Fast parallel algorithms for short-range molecular dynamics. J Comput Phys 117(1):1–19

    MATH  Google Scholar 

  49. Rushton M (2014) atsim.potentials. http://atsimpotentials.readthedocs.io/en/latest/. Accessed 14 Apr 2016

  50. Frenkel D, Smit B (2001) Understanding molecular simulation: from algorithms to applications, vol 1. Academic Press, San Diego

    MATH  Google Scholar 

  51. Martyna GJ, Tobias DJ, Klein ML (1994) Constant pressure molecular dynamics algorithms. J Chem Phys 101(5):4177–4189

    Google Scholar 

  52. Stukowski A (2009) Visualization and analysis of atomistic simulation data with OVITO—the Open Visualization Tool. Modell Simul Mater Sci Eng 18(1):015012

    MathSciNet  Google Scholar 

  53. Li X, Fan X, Askounis A, Wu K, Sefiane K, Koutsos V (2013) An experimental study on dynamic pore wettability. Chem Eng Sci 104:988–997

    Google Scholar 

  54. Jiang T-S, Soo-Gun O, Slattery JC (1979) Correlation for dynamic contact angle. J Colloid Interface Sci 69(1):74–77

    Google Scholar 

  55. Bachu S, Bennion DB (2008) Interfacial tension between CO2, freshwater, and brine in the range of pressure from (2 to 27) MPa, temperature from (20 to 125) °C, and water salinity from (0 to 334 000) mg L − 1. J Chem Eng Data 54(3):765–775

    Google Scholar 

  56. Bikkina PK, Shoham O, Uppaluri R (2011) Equilibrated interfacial tension data of the CO2–water system at high pressures and moderate temperatures. J Chem Eng Data 56(10):3725–3733

    Google Scholar 

  57. Bikkina PK (2011) Contact angle measurements of CO2–water–quartz/calcite systems in the perspective of carbon sequestration. Int J Greenhouse Gas Control 5(5):1259–1271

    Google Scholar 

  58. Rohatgi A WebPlotDigitizer, Version 4.1. https://automeris.io/WebPlotDigitizer. Accessed Jan 2018

  59. Blake T, Haynes J (1969) Kinetics of liquid–liquid displacement. J Colloid Interface Sci 30(3):421–423

    Google Scholar 

  60. Blake T, De Coninck J (2002) The influence of solid–liquid interactions on dynamic wetting. Adv Colloid Interface Sci 96(1–3):21–36

    Google Scholar 

  61. Glasstone S, Laidler K, Eyring H (1941) The theory of rate processes. McGraw-Hill Book Co. Inc., New York

    Google Scholar 

  62. Bertrand E, Blake TD, De Coninck J (2009) Influence of solid–liquid interactions on dynamic wetting: a molecular dynamics study. J Phys: Condens Matter 21(46):464124

    Google Scholar 

  63. Duvivier D, Blake TD, De Coninck J (2013) Toward a predictive theory of wetting dynamics. Langmuir 29(32):10132–10140

    Google Scholar 

  64. Müller-Plathe F (1999) Reversing the perturbation in nonequilibrium molecular dynamics: an easy way to calculate the shear viscosity of fluids. Phys Rev E 59(5):4894

    Google Scholar 

  65. Friedberg R, Cameron JE (1970) Test of the Monte Carlo method: fast simulation of a small Ising lattice. J Chem Phys 52(12):6049–6058

    Google Scholar 

  66. Fincham D, Quirke N, Tildesley D (1986) Computer simulation of molecular liquid mixtures. I. A diatomic Lennard-Jones model mixture for CO2/C2H6. J Chem Phys 84(8):4535–4546

    Google Scholar 

Download references

Acknowledgements

This work was supported in part by the Australian Research Council through Discovery Projects (DP170102886 and DP190102954). P.H. acknowledges the financial support from The University of Sydney Nano Institute Postgraduate Top-Up Scholarship. This research was undertaken and supported with the assistance of resources and services from the National Computational Infrastructure (NCI), which is supported by the Australian Government, and the University of Sydney HPC service at The University of Sydney.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Luming Shen.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendices

1.1 Appendix 1: Calculation of density

The calculations of density at different pressures for MDPD and DPD particles follow a procedure similar to our previous work [37]. 8000 particles are created in a cubic simulation box with periodic boundary conditions in all three directions. The system is run in the NPT ensemble at a constant temperature T* and pressure P* for 100,000 time steps. The instantaneous density is calculated at every 20 time steps. The average density and the standard deviation are calculated from the instantaneous density obtained from the last 50,000 time steps. A time step Δt* = 0.0025 is used.

1.2 Appendix 2: Calculation of viscosity

For the calculation of viscosity at a particular pressure, a system composed of 3000 particles (MDPD or DPD) is created in a periodic simulation box with a particle density calculated at the studied pressure. The aspect ratio of the simulation box is L*x :L*y :L*z  = 1:1:3. The system is first equilibrated with a Langevin thermostat for 100,000 time steps. The Müller-Plathe algorithm is then applied to calculate the viscosity [64]. A schematic diagram of the simulated system is shown in Fig. 11. The system is divided into 20 layers in the z-direction. At every 50 time steps, the particle in the first layer with the largest speed in the positive x-direction is selected. Similarly, the particle in the middle layer (11th layer) with the largest speed in the negative x-direction is selected. The x-components of the momenta of the two particles are then exchanged. The exchange of momenta induces a shear velocity profile in the system with periodic boundary conditions. The simulations are run for 1,100,000 time steps at the NVT ensemble using the Nosé–Hoover chains thermostat. The first 100,000 time steps are considered as the equilibrium run, for the development of the velocity profile. The last 1,000,000 time steps are considered as the production run. The velocity profile is calculated at every 10,000 time steps by averaging the x-component of the particle velocities in each layer. The total momentum exchanged p*x is calculated at every 10,000 time steps. In the steady state, the momentum flux J* can be calculated as:

$$ J^{*} = p_{x}^{*} /2t^{*} L_{x}^{*} L_{y}^{*} , $$
(22)

where t* is the time duration of the production run; L*x and L*y are the length of the simulation box in the x- and y-directions, respectively. The factor of 2 is used here due to the periodic boundary conditions applied. Therefore, the viscosity η* can be calculated by:

$$ J^{*} = - \eta^{*} \frac{{\partial v_{x}^{*} }}{{\partial z^{*} }}, $$
(23)

where ∂v*x /∂z* is the gradient of the x-component of the fluid velocity with respect to the z-direction within the region between the first and middle layers. The averaged viscosity can be calculated using Eqs. (22) and (23) from the production run. The value of J* is obtained at the end of the simulation. ∂v*x /∂z* is calculated from the slope of the time-averaged velocity profile by linear regression at every 10,000 time steps. The average of ∂v*x /∂z* is then calculated from the 100 samples of the slope. The simulations are repeated five times for each case, and the average and standard deviation of the viscosity are then calculated accordingly.

Fig. 11
figure 11

A schematic diagram of the simulated system for the calculation of viscosity using the Müller–Plathe algorithm [64]. The x-component velocity profile is plotted in red colour. (Color figure online)

1.3 Appendix 3: Calculation of interfacial tension

The CO2–water interface simulation is performed to calculate the interfacial tension of the CO2–water system at a certain T* and P*. To construct the system, an 8000-water-particle cubic box, which is equilibrated at targeted T* and P* with periodic boundary conditions applied in all three dimensions, is prepared. The CO2 box is also prepared with the same size as the water box. The particle numbers of CO2 in the box is calculated from the equilibrium density at the targeted T* and P*. The two boxes are replicated twice in the z-direction and placed together in a periodic simulation box with a ratio of L*x :L*y :L*z  = 1:1:4. An energy minimization process is applied to the system at the beginning to avoid the particle overlapping at the two CO2–water interfaces. The water phase is first relaxed by allowing the water particles to move at the Langevin thermostat for 10,000 time steps, while the CO2 particles are fixed. The CO2 phase is then relaxed at the Langevin thermostat with water particles being fixed for 10,000 time steps. Next, the whole system is relaxed together at the Langevin thermostat for another 10,000 time steps. An example of a relaxed system is shown in Fig. 12. The relaxed system is then further equilibrated under the NPzT ensemble for 100,000 time steps to reach the targeted T* and P*. The system is then switched to the NVT ensemble and run for 1,000,000 time steps. The last 600,000 time steps are considered as the production run. The interfacial tension γ* is calculated as:

$$ \gamma^{*} = \frac{1}{2}L_{z}^{*} \left\langle {P_{zz}^{ *} - \frac{{P_{xx}^{ *} + P_{yy}^{*} }}{2}} \right\rangle , $$
(24)

where P*xx , P*yy , and P*zz are the x-, y-, and z-components of the virial pressure of the system; L*z is the length of the simulation box in the z-direction; the time averages are performed inside of < > for the data recorded every 20 time steps. The statistical errors of the surface tension are estimated by calculating the statistical inefficiency [42, 65, 66].

Fig. 12
figure 12

A snapshot of the system of the water–CO2 interface simulation after being relaxed at the Langevin thermostat

1.4 Appendix 4: Water bridge simulation

A liquid bridge simulation is performed to ensure that the selected parameters of water–silica interactions would be able to generate a complete wetting of the silica surface. A periodic simulation box with L*x  = 50.67, L*y  = 10.48, and L*z  = 26.38 is created with water sandwiched between two silica slabs. The particle density of water is initially close to the bulk water density of the equilibrium water–vapour system. The water particles are equilibrated at the Langevin thermostat for 100,000 time steps, with the fixed solid particles. Next, the left and right portions of the water particles are deleted with around one-third of the water particles remaining in the middle between the two silica slabs. Figure 13 shows a snapshot of the initial state of the remained water particles between the two silica slabs. The system is further equilibrated under the NVT ensemble with a Nosé–Hoover chains thermostat until it reaches equilibrium. If the water particles spread over and cover all the silica surfaces inside, it indicates a complete wetting of the simulated silica surface.

Fig. 13
figure 13

A snapshot of the initial state of the system for the liquid bridge simulation

1.5 Appendix 5: Contact angle measurements

There are different ways of measuring contact angles. In this work, the contact angle is measured by fitting a circle to the interface by assuming a circular shape of the interface. The simulation box is divided into small square bins with a size of 0.5rc × 0.5rc (equivalent to 0.01768 × 0.01768 μm2) in the x–z plane, and the time-averaged density in each bin is calculated at every 100,000 time steps throughout the simulations, with instantaneous density recorded every 20 time steps. For the simulations with the piston moving at a large velocity greater than 0.7 m/s, the 2D density profile is calculated at every 10,000 time steps and the movement of the piston in the x-axis is also checked at every 10,000 time steps. When the piston has moved one bin width forward (0.5rc), the calculated density profile should be shifted one bin width backward in order to have a stable water–CO2 interface at large piston velocity. In addition, the first bin at the x-direction should be moved to the end due to the periodic boundary condition applied along the x-axis. Hence, the shifted density profiles are averaged again to obtain the density profile at every 100,000 time steps, in order to calculate the contact angle. The 2D density contour is plotted using Matlab. The location of the interface between CO2 and water can be calculated from the 2D density contour by calculating the density at the interface ρint = 0.5(ρwater + ρCO2), where ρwater and ρCO2 are the bulk densities of water and CO2, respectively, which can be estimated from the bulk regions in water and CO2 phases. Since the density perturbations can happen near the solid surface, only the interface data that are 2rc away from the solid surface are used to fit the circle. An example of the circular fitting is shown in Fig. 14. The contact angle is calculated from the tangent of the circle where the circle and solid surfaces meet. The contact angle values are calculated by averaging all the top and bottom contact angles during the production run, and the standard deviations of the contact angles are calculated accordingly.

Fig. 14
figure 14

The measurement of the contact angle by fitting a circle to the interface. The markers of small circles are the coordinates of the water–CO2 interfaces. The solid blue lines indicate the location of the solid surface, and the interface data between the two blue dashed lines are used in the circle fitting

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Huang, P., Shen, L., Gan, Y. et al. Numerical investigation of microscale dynamic contact angles of the CO2–water–silica system using coarse-grained molecular approach. Comput Mech 66, 707–722 (2020). https://doi.org/10.1007/s00466-020-01873-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-020-01873-7

Keywords

Navigation