Skip to main content
Log in

An Alternative High-Density Ratio Pseudo-potential Lattice Boltzmann Model with Surface Tension Adjustment Capability

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

In this paper, an alternative model is presented to simultaneously achieve thermodynamic consistency, leading to high-density ratios, and surface tension adjustment capability based on the pseudo-potential lattice Boltzmann model (LBM). These features are achieved by adding two relatively simple terms, each with an adjustable coefficient, selected from the Taylor expansion of the original interaction force, one for achieving and maintaining thermodynamic consistency, and the other for adjusting the surface tension independent of the density ratio. The model also takes advantage of being independent of using SRT, MRT, 2D or 3D lattice models. The capability of the model is evaluated and verified by performing several static and dynamic benchmark test cases. First, the coexistence densities extracted for a flat interface from the model are compared with those of the Maxwell equal area rule in a wide temperature range. Results show that the thermodynamic consistency is well achieved and the coexistence densities are independent of the surface tension strength as well. Next, the well-known Laplace law for a droplet is evaluated and satisfied with the model. In addition, a wide range of surface tension values is achievable at a fixed temperature by adjusting the surface tension adjustment coefficient. After that, the ellipsoidal droplet oscillation is simulated with a good agreement between the analytical and numerical results for the oscillation period. The spurious velocities around the droplet are then evaluated and shown to be reasonable, and comparatively low even at a relatively high-density ratio around of 2400. It was found that the magnitude of these velocities can be further reduced, especially at high-density ratios, by reducing the liquid to gas dynamic viscosity ratio or equivalently increasing the gas to liquid kinematic viscosity ratio. Next, the stability, and the fascinating phenomenon of symmetric bifurcation of a rotating planar droplet are simulated and formation of two- and four-lobed shapes along with rigid rotation is presented depending on the value of the surface tension. Finally, the droplet impact on a liquid film is simulated at a relatively high-density ratio around of 720, and Reynolds number of 1440. However, the upper limit of physical parameters for simulating this phenomenon is found to be a density ratio of 720 and Re and We numbers of 6600 and 275, respectively. Effects of the surface tension strength are shown as well. The crown spread radius from the model agrees well with the analytical solution reported in the literature.

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

Similar content being viewed by others

References

  1. Chen, L., Kang, Q., Mu, Y., He, Y., Tao, W.: International journal of heat and mass transfer a critical review of the pseudopotential multiphase lattice boltzmann model: methods and applications. Heat Mass Transf. 76, 210–236 (2014)

    Article  Google Scholar 

  2. Huang, H., Sukop, M., Lu, X.: Multiphase Lattice Boltzmann Methods: Theory and Application. Wiley, New York (2015)

    Book  Google Scholar 

  3. Krüger, T., Kusumaatmaja, H., Kuzmin, A., Shardt, O., Silva, G., Viggen, E.M.: The Lattice Boltzmann Method. Springer International Publishing, Cham (2017)

    Book  MATH  Google Scholar 

  4. Huang, H., Krafczyk, M., Lu, X.: Forcing term in single-phase and Shan–Chen-type multiphase lattice Boltzmann models. Phys. Rev. E 84, 46710 (2011)

    Article  ADS  Google Scholar 

  5. Hu, A., Li, L., Uddin, R.: Force method in a pseudo-potential lattice Boltzmann model. J. Comput. Phys. 294, 78–89 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Kupershtokh, A.L., Medvedev, D.A., Karpov, D.I.: On equations of state in a lattice Boltzmann method. Comput. Math. Appl. 58(5), 965–974 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  7. Yu, Z., Fan, L.-S.: Multirelaxation-time interaction-potential-based lattice Boltzmann model for two-phase flow. Phys. Rev. E 82(4), 46708 (2010)

    Article  ADS  Google Scholar 

  8. Sbragaglia, M., Benzi, R., Biferale, L., Succi, S., Sugiyama, K., Toschi, F.: Generalized lattice Boltzmann method with multirange pseudopotential. Phys. Rev. E 75(2), 1–13 (2007)

    Article  MathSciNet  Google Scholar 

  9. Guo, Z., Zheng, C., Shi, B.: Discrete lattice effects on the forcing term in the lattice boltzmann method. Phys. Rev. E 65(4), 4630 (2002)

    Article  Google Scholar 

  10. McCracken, M.E., Abraham, J.: Multiple-relaxation-time lattice-Boltzmann model for multiphase flow. Phys. Rev. E 71(3), 36701 (2005)

    Article  ADS  Google Scholar 

  11. Li, Q., Luo, K.H., Li, X.J.: Forcing scheme in pseudopotential lattice Boltzmann model for multiphase flows. Phys. Rev. E 86, 16709 (2012)

    Article  ADS  Google Scholar 

  12. Li, Q., Luo, K.H., Li, X.J.: Lattice Boltzmann modeling of multiphase flows at large density ratio with an improved pseudopotential model. Phys. Rev. E 87(5), 53301 (2013)

    Article  ADS  Google Scholar 

  13. Zhang, D., Papadikis, K., Gu, S.: Investigations on the droplet impact onto a spherical surface with a high density ratio multi-relaxation time lattice-Boltzmann model. Commun. Comput. Phys. 16(4), 892–912 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  14. Kharmiani, S., Passandideh-Fard, M., Niazmand, H.: Simulation of a single droplet impact onto a thin liquid film using the lattice Boltzmann method. J. Mol. Liq. 222, 1172–1182 (2016)

    Article  Google Scholar 

  15. Fang, W.Z., Chen, L., Kang, Q.J., Tao, W.Q.: Lattice Boltzmann modeling of pool boiling with large liquid-gas density ratio. Int. J. Therm. Sci. 114, 172–183 (2017)

    Article  Google Scholar 

  16. Kharmiani, S.F., Passandideh-Fard, M.: A two-phase lattice Boltzmann study on injection filling of cavities with arbitrary shapes. Int. J. Multiph. Flow 101, 11–23 (2018)

    Article  MathSciNet  Google Scholar 

  17. Zhang, D., Papadikis, K., Gu, S.: Three-dimensional multi-relaxation time lattice-Boltzmann model for the drop impact on a dry surface at large density ratio. Int. J. Multiph. Flow 64, 11–18 (2014)

    Article  MathSciNet  Google Scholar 

  18. Xu, T., Zhao, S., An, L., Shi, L.: A three-dimensional pseudo-potential-based lattice Boltzmann model for multiphase flows with large density ratio and variable surface tension. Int. J. Heat Fluid Flow 56, 261–271 (2015)

    Article  Google Scholar 

  19. Khajepor, S., Wen, J., Chen, B.: Multipseudopotential interaction: a solution for thermodynamic inconsistency in pseudopotential lattice Boltzmann models. Phys. Rev. E 91(2), 23301 (2015)

    Article  ADS  Google Scholar 

  20. Huang, R., Wu, H.: Third-order analysis of pseudopotential lattice Boltzmann model for multiphase flow. J. Comput. Phys. 327, 121–139 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Falcucci, G., Falcucci, G., Bella, G., Chiatti, G., Chibbaro, S., Sbragaglia, M., Succi, S.: Lattice boltzmann models with mid-range interactions. Commun. Comput. Phys 2, 1071–1084 (2007)

    Google Scholar 

  22. Chibbaro, S., Falcucci, G., Chiatti, G., Chen, H., Shan, X., Succi, S.: Lattice Boltzmann models for nonideal fluids with arrested phase-separation. Phys. Rev. E 77(3), 36705 (2008)

    Article  ADS  Google Scholar 

  23. Yuan, P., Schaefer, L., Yuan, P., Schaefer, L.: Equations of state in a lattice Boltzmann model. Phys. Fluids 42101(4), 42101 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  24. Li, Q., Luo, K.H.: Achieving tunable surface tension in the pseudopotential lattice Boltzmann modeling of multiphase flows. Phys. Rev. E 88(5), 53307 (2013)

    Article  ADS  Google Scholar 

  25. Hu, A., Li, L., Uddin, R.: Surface tension adjustment in a pseudo-potential lattice Boltzmann model. arXiv Repr. (2014)

  26. Ammar, S., Pernaudat, G., Trépanier, J.Y.: A multiphase three-dimensional multi-relaxation time (MRT) lattice Boltzmann model with surface tension adjustment. J. Comput. Phys. 343, 73–91 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Lycett-Brown, D., Luo, K.H.: Improved forcing scheme in pseudopotential lattice Boltzmann methods for multiphase flow at arbitrarily high density ratios. Phys. Rev. E 91(2), 23305 (2015)

    Article  ADS  Google Scholar 

  28. Lycett-Brown, D., Luo, K.H.: Cascaded lattice Boltzmann method with improved forcing scheme for large-density-ratio multiphase flow at high Reynolds and Weber numbers. Phys. Rev. E 94(5), 53313 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  29. Li, Q., Luo, K.H., Kang, Q.J., He, Y.L., Chen, Q., Liu, Q.: Lattice Boltzmann methods for multiphase flow and phase-change heat transfer. Prog. Energy Combust. Sci. 52, 62–105 (2016)

    Article  Google Scholar 

  30. Lee, T., Lin, C.-L.: A stable discretization of the lattice Boltzmann equation for simulation of incompressible two-phase flows at high density ratio. J. Comput. Phys. 206(1), 16–47 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Mukherjee, S., Abraham, J.: A pressure-evolution-based multi-relaxation-time high-density-ratio two-phase lattice-Boltzmann model. Comput. Fluids 36(6), 1149–1158 (2007)

    Article  MATH  Google Scholar 

  32. He, X., Doolen, G.D.: Thermodynamic foundations of kinetic theory and lattice Boltzmann models for multiphase flows. J. Stat. Phys. 107(1), 309 (2002)

    Article  MATH  Google Scholar 

  33. Lamb, H.: Hydrodynamics, 6th edn, pp. 366–369. Dover, New York (1994)

    Google Scholar 

  34. Lewis, D., Marsden, J., Ratiu, T.: Stability and bifurcation of a rotating planar liquid drop. J. Math. Phys. 28(10), 2508–2515 (1987)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Marsden, J.E.: Lectures on Mechanics, p. 268. Cambridge University Press, Cambridge (1992)

    Book  Google Scholar 

  36. Holm, D.D., Marsden, J.E., Ratiu, T., Weinstein, A.: Nonlinear stability of fluid and plasma equilibria. Phys. Rep. 123(1–2), 1–116 (1985)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Yarin, A.L., Weiss, D.A.: Impact of drops on solid-surfaces: self-similar capillary waves, and splashing as a new-type of kinematic discontinuity. J. Fluid Mech. 283, 141–173 (1995)

    Article  ADS  Google Scholar 

  38. Yarin, A.L.: Drop impact dynamics: splashing, spreading, receding, bouncing. Annu. Rev. Fluid Mech. 38, 159–192 (2006)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. Shan, X.: Pressure tensor calculation in a class of nonideal gas lattice Boltzmann models. Phys. Rev. E 77, 66702 (2008)

    Article  ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hamid Niazmand.

Appendices

Appendix 1

In this section, the model is analyzed mathematically based on the discrete pressure tensor proposed by Shan [39]. The discrete pressure tensor of the pseudopotential model using the original interaction force given by Eq. (12), is obtained as follows [12]:

$$ \int {(\nabla .{\text{ P}})d} \varOmega = \int {\nabla .(\rho c_{s}^{2} {\text{I}})d} \varOmega - \int {\text{F}} d\varOmega $$
(a1)

where \( d\varOmega \) is a finite volume. Using the Gauss integration theorem:

$$ \int {{\text{ P}} .d} {\text{A}} = \int {\rho c_{s}^{2} {\text{I}} .d} {\text{A}} - \int {\text{F}} d\varOmega $$
(a2)

Then in discrete form:

$$ \sum {{\text{P}} . {\text{A}}} { = }\sum {\rho c_{s}^{2} {\text{I}} . {\text{A}}} - \sum {\text{F}} $$
(a3)
$$ \to \varvec{P}_{d} = \rho c_{s}^{2} \varvec{I} + \frac{G}{2}\psi \left( {\varvec{x},t} \right)\left[ {\mathop \sum \limits_{\alpha } w\left( {\left| {\varvec{e}_{\alpha } } \right|^{2} } \right)\psi \left( {\varvec{x} + \varvec{e}_{\alpha } ,t} \right)\varvec{e}_{\alpha } \varvec{e}_{\alpha } } \right] $$
(a4)

The Taylor series expansion of the above equation leads to the following expression:

$$ \varvec{P}_{\varvec{d}} = \left( {\rho c_{s}^{2} + \frac{{Gc^{2} }}{2}\psi^{2} } \right.\left. { + \frac{{Gc^{4} }}{12}\psi \nabla^{2} \psi } \right)\varvec{I} + \frac{{Gc^{4} }}{6}\psi \nabla \nabla \psi $$
(a5)

The modified interaction force is again presented below:

$$ \begin{aligned} \varvec{F}_{{\varvec{new}}} = \varvec{F} + \lambda \frac{{Gc^{4} }}{12}\nabla \left( {\left| {\nabla \psi } \right|^{2} } \right) - \kappa \frac{{Gc^{4} }}{6}\nabla^{2} \psi \nabla \psi \hfill \\ = \varvec{F} + \lambda \frac{{Gc^{4} }}{12}\nabla \left( {\left| {\nabla \psi } \right|^{2} } \right) - \kappa \frac{{Gc^{4} }}{6}\left[ {\nabla \cdot \left( {\nabla \psi \nabla \psi } \right) - \frac{1}{2}\nabla \left| {\nabla \psi } \right|^{2} } \right] \hfill \\ \end{aligned} $$
(a6)

Therefore, the new discrete pressure tensor is simply given by:

$$ \to \varvec{P}_{d}^{new} = \left( {\rho c_{s}^{2} + \frac{{Gc^{2} }}{2}\psi^{2} } \right.\left. { + \left( { - \lambda - \kappa } \right)\frac{{Gc^{4} }}{12}\left| {\nabla \psi } \right|^{2} + \frac{{Gc^{4} }}{12}\psi \nabla^{2} \psi } \right)\varvec{I} + \frac{{Gc^{4} }}{6}\left( {\kappa \nabla \psi \nabla \psi + \psi \nabla \nabla \psi } \right) $$
(a7)

Notice that the added terms to the original interaction force in Eq. (a6) are in the gradient form just like the well-known \( \nabla \left( {\rho c_{s}^{2} } \right) \) term, thus they are directly absorbed in the discrete pressure tensor as well.

The normal component of the tensor is obtained as follows:

$$ P_{n}^{d} = \rho c_{s}^{2} + \frac{{Gc^{2} }}{2}\psi^{2} + \frac{{Gc^{4} }}{12}\left[ {\left( { - \lambda + \kappa } \right)\left( {\frac{d\psi }{dn}} \right)^{2} + 3\psi \frac{{d^{2} \psi }}{{dn^{2} }}} \right] $$
(a8)

Thus, \( \varepsilon_{d} = \frac{ - 2\alpha }{\beta } = - \frac{2}{3}\left( { - \lambda + \kappa } \right) \).

As observed, the same as the continuum analysis, defining \( \lambda = \lambda_{0} + \kappa \) leads to:

$$ \varepsilon_{d} = \frac{2}{3}\lambda_{0} $$
(a9)

Furthermore, considering a flat interface along y direction, the surface tension relation can be obtained as follows:

$$ P_{n}^{d} = P_{xx}^{d} = \rho c_{s}^{2} + \frac{{Gc^{2} }}{2}\psi^{2} + \frac{{Gc^{4} }}{12}\left[ {\left( { - \lambda + \kappa } \right)\left( {\frac{d\psi }{dx}} \right)^{2} + 3\psi \frac{{d^{2} \psi }}{{dx^{2} }}} \right] $$
(a10)
$$ P_{\tau }^{d} = P_{yy}^{d} = \rho c_{s}^{2} + \frac{{Gc^{2} }}{2}\psi^{2} + \frac{{Gc^{4} }}{12}\left[ {\left( { - \lambda - \kappa } \right)\left( {\frac{d\psi }{dx}} \right)^{2} + \psi \frac{{d^{2} \psi }}{{dx^{2} }}} \right] $$
(a11)
$$ \sigma_{d} = \int\limits_{ - \infty }^{ + \infty } {\left( {P_{xx} - P_{yy} } \right)} dx = \frac{{Gc^{4} }}{6}\int\limits_{ - \infty }^{ + \infty } {\left[ {\kappa \left( {\frac{d\psi }{dx}} \right)^{2} + \psi \frac{{d^{2} \psi }}{{dx^{2} }}} \right]} dx $$
(a12)
$$ \int_{ - \infty }^{ + \infty } {\psi \frac{{d^{2} \psi }}{{dx^{2} }}} dx = \left. {\psi \frac{d\psi }{dx}} \right|_{ - \infty }^{ + \infty } - \int_{ - \infty }^{ + \infty } {\frac{d\psi }{dx}} d\psi $$
(a13)

Knowing \( \frac{d\psi }{dx} \) is zero at \( \pm \infty \) and \( \frac{d\psi }{dx} = \frac{d\psi }{d\rho }\frac{d\rho }{dx} = \psi^{\prime}\frac{d\rho }{dx} \):

$$ \to \sigma_{d} = - \frac{{Gc^{4} }}{6}\left( {1 - \kappa } \right)\int\limits_{{\rho_{g} }}^{{\rho_{l} }} {\psi^{\prime 2} \frac{d\rho }{dx}} d\rho $$
(a14)

which is exactly the same relation as Eq. (36) derived from the continuum pressure tensor.

Appendix 2

In this appendix, the current model, the multirange potential [8] and the Li and Luo model [24] are further discussed and compared.

The multirange interaction force proposed by Sbragaglia et al. is given by [8]:

$$ \varvec{F} = - \psi \left( {\varvec{x},t} \right)\mathop \sum \limits_{\alpha } w\left( {\left| {\varvec{e}_{\alpha } } \right|^{2} } \right)\left[ {G_{1} \psi \left( {\varvec{x} + \varvec{e}_{\alpha } ,t} \right)\varvec{ + }G_{2} \psi \left( {\varvec{x} + 2\varvec{e}_{\alpha } ,t} \right)} \right]\varvec{e}_{\alpha } $$
(b1)

This interaction force was claimed to be able to adjust the EOS through G1 and the surface tension through G2, independently. In part III of their paper [24], Li and Luo made an analysis of the Sbragaglia et al. model to obtain the corresponding relation for the ε parameter and found the following relation (see Eq. (16) in [24]):

$$ \varepsilon = \frac{{4G_{2} }}{{G_{1} + 6G_{2} }} $$
(b2)

where it is clear that the parameter ε is affected by both G1 and G2 coefficients implementing the multirange potential model of Sbragaglia et al. [8]. It should be mentioned that the interaction potential in the Li and Luo analysis of the multirange model is taken to be \( \psi = \sqrt {\frac{{2\left( {P_{EOS} - \rho c_{s}^{2} } \right)}}{{(G_{1} + 2G_{2} )c^{2} }}} \). Therefore, Li and Luo showed that the multirange interaction force is not capable of adjusting the EOS and surface tension independently, implementing real EOSs like the Carnahan-Starling.

In terms of achieving thermodynamic consistency, our approach is considerably different from the Li and Luo one. The parameter ε in Li and Luo model is given by [12]:

$$ \varepsilon = - 2\left( {\alpha + 24G\sigma } \right)/\beta $$
(b3)

where σ is a parameter adjusted to numerically achieve the thermodynamic consistency, achieving by applying the following model proposed earlier by Q. Li et al. [11]:

$$ \varvec{F}_{\alpha } = \omega_{\alpha } \delta t\left( {1 - \frac{1}{2\tau }} \right)\left[ {\frac{{\varvec{e}_{\alpha } - \varvec{V}^{\varvec{'}} }}{{c_{s}^{2} }} + \frac{{\left( {\varvec{e}_{\alpha } \cdot \varvec{V}^{\varvec{'}} } \right)}}{{c_{s}^{4} }}\varvec{e}_{\alpha } } \right] \cdot \varvec{F} $$
(b4)

In other words, they modified the velocity in the Guo force scheme as follows, using the SRT LB equation [11]:

$$ {\varvec{V}^{\prime}} = {\varvec{V}} + \sigma {\mathbf{F}}/\left( {\upsilon \psi ^{2} } \right) $$
(b5)

where \( \varvec{V} = \varvec{u} + {\mathbf{F}}\delta t/2\rho \) with \( \rho \varvec{u} = \mathop \sum \nolimits f_{\alpha } e_{\alpha } \) and in the denominator, \( \nu = (\tau - 0.5)/3 \) being the kinematic viscosity. Therefore, our approach for adjusting the mechanical stability condition ε, i.e., adding the term \( + \lambda \frac{{Gc^{4} }}{12}\nabla \left( {\left| {\nabla \psi } \right|^{2} } \right) \) to the original interaction force, is different and novel, although both models add a source term to the LB equation.

In addition, when developed to the MRT version, the Li and Luo model summarized in Eqs (b4) and (b5), is affected by other relaxation times, especially in Eq. (b5), thus some considerations and modifications are carried out on the force term in the momentum space and, as reported in Li and Luo [12], thus mathematical calculations are needed. Whereas, in the current model the source term in the momentum space is computed by a simple code multiplying the force term by the transformation matrix due to no modifications on the Guo force scheme.

In terms of the approach for adjusting the surface tension independently, again the source term should be calculated mathematically in the Li and Luo model, considering the discrete lattice effects on the source term whereas, computing a Laplacian term is the only needed task to make the surface tension adjustable in the current model. As mentioned before, the multirange interaction force by Sbragaglia et al. is based on modifying the interaction force as it is in the current model. Therefore, modification of the interaction force can be viewed at least a different mathematical modeling and approach. The superiority of the current model over the multirange model includes the independent adjustment of the EOS and surface tension, considering only the nearest neighbors interactions and, therefore, easier treatment of boundary conditions.

The value of ε for achieving thermodynamic consistency in the Li and Luo model is indeed a numerical finding as it is in our model, both models use the same interaction potential given by Eq. (13). However, it is not a hard task to find the consistent ε value and is found to be almost independent of the EOSs implemented. Similarly, the parameter λ for achieving thermodynamic consistency in the current model is found by performing several simulations, but when recognized, its value is almost the same implementing other EOSs. For example, we also implemented the Peng–Robinson (P-R) EOS and the consistent λ is found to be in the range of 2.6 to 2.65, which is almost the same value of 2.65 found for the C-S EOS. The consistent σ in the Li and Luo model is reported to be in the range of 0.11–0.12. Xu et al. [18] extended their model to the D3Q15 lattice and σ = 0.12 was reported for both C-S and P-R EOSs. Therefore, in this regard, the obtained thermodynamically consistent value of ε in the current model, like the Li and Luo one, is almost independent of the EOS and thus it is applicable in the general framework of the pseudopotential LBM. The thermodynamically consistent value of ε for the current model is, therefore, \( \varepsilon = 1 + \lambda_{0} = 1 + 2.65 = 3.65 \), almost independent of implementing different equation of states.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kharmiani, S.F., Niazmand, H. & Passandideh-Fard, M. An Alternative High-Density Ratio Pseudo-potential Lattice Boltzmann Model with Surface Tension Adjustment Capability. J Stat Phys 175, 47–70 (2019). https://doi.org/10.1007/s10955-019-02243-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-019-02243-1

Keywords

Navigation