Skip to main content
Log in

Air Entrainment Modeling in the SPH Method: A Two-Phase Mixture Formulation with Open Boundaries

  • Published:
Flow, Turbulence and Combustion Aims and scope Submit manuscript

Abstract

Air entrainment within water is a common feature of flows over hydraulic works – spill over a dam, wave breaking on a dike, etc. – and its accurate modeling is a key to better design such structures. The Smoothed Particle Hydrodynamics (SPH) method appears as a natural way to model such highly distorted flows. To avoid computationally prohibitive costs related to the full discretization of bubbles or drops, a mixture model for high density ratio flows relying on a volume-based formulation with relative velocity between phases has first been developed and validated in Fonty et al. (Proceedings of 13th international SPHERIC workshop. Galway, Ireland, 2018; Int J Multiph Flow 111:158–174, 2019. https://doi.org/10.1016/j.ijmultiphaseflow.2018.11.007). Instead of having a once and for all assigned phase as in multifluid SPH, each particle now carries both phases through their respective volume fractions. In the present work, in order to handle practical air entrainment application cases, the open boundary formulation described in Ferrand et al. (Comput Phys Commun 210:29–44, 2017. https://doi.org/10.1016/j.cpc.2016.09.009) is adapted to this mixture model. Then, after introducing turbulence through a \(k{-}\epsilon\) model, a specific closure on the air bubbles relative velocity is proposed including a Stokesian drag term and turbulent diffusion. This model is then applied to two cases of air entrainment: a stepped spillway for interfacial aeration and a plunging jet for local aeration. Finally a 3D industrial test case of discharge-control structure at the La Coche power plant (France) is considered. While valuable insights are obtained for the volume fraction field, further investigations are required to improve the modeling of the flow dynamics.

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
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24

Similar content being viewed by others

Notes

  1. The volume represents the SPH particle that will be introduced in the numerical section.

  2. \(\alpha\) and \(\beta\) denote the volume fractions together with the phase names, without any risk of confusion.

  3. In presence of surface tension, \(\varvec{M}^{\alpha }+\varvec{M}^{\beta }=-\,2H_{\beta \alpha }\sigma _{S}\) where \(H_{\beta \alpha }\) is the mean curvature between phases \(\alpha\) and \(\beta\), and \(\sigma _{S}\) is the surface tension coefficient (relation written considering that there are no extra interfacial momentum source accounting for local surface forces resulting from pressure and shear stress deviations from interfacial average values).

  4. The divergence-free property is relaxed for the continuity resolution but kept in the momentum equation resolution, as classically done in the weakly compressible SPH framework.

  5. In this work we are mainly concerned by what happens in the water phase and mixture region. The modelling of the air phase appears as a requirement of the mixture model to exchange phases between particles when air entrainment occurs but we do not intend here to be precise in the resolution of the air phase dynamics, that prove to be affected by numerical instabilities in the SPH framework.

  6. With the approximation of neglecting the viscous contribution.

  7. This choice is related to experimental measurements on the considered applications cases, (Chanson 2001; Bertola et al. 2017), for which peaks of bubble chord distributions were noticed around 1–2 mm.

  8. This approach proves to give correct values of the relative velocity without iteration: the error is below 1% for \(Re>104\) and below 0.2% for \(Re>218\). The peak of error is reached for very small Reynolds number, at \(Re\sim 2\) where it reaches 10%. However, in the practical air-water applications considered, we fall within the \(1\%\) error range.

  9. This test case aims at checking the correct implementation of the model, so that we chose a simplified formulation of the relative velocity to be able to derive an analytical formulation. These assumptions allow one to make the link between the general framework presented and this special verification test case.

  10. This velocity profile is an arbitrary choice to impose the desired volume flow rate as the “true” profile is not known in this region near the structure. It is applied across the water height from the first layer of particles near the bottom to the free surface.

  11. The code, still in development, is not optimized.

  12. The planar plunging jet test case suggested that one could get better results for that kind of configuration without the turbulence model considered in this work for local aeration.

References

  • Abramowitz, M., Stegun, I.: Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. U.S. Department of Commerce, Washington, DC (1964)

    MATH  Google Scholar 

  • Alvarado-Rodríguez et al., C.E., Klapp, J., Sigalotti, L.D.G., Domínguez, J.M., de la Cruz Sánchez, E.: Nonreflecting outlet boundary conditions for incompressible flows using SPH. Comput. Fluids 159, 177–188 (2017). https://doi.org/10.1016/j.compfluid.2017.09.020

    Article  MathSciNet  MATH  Google Scholar 

  • Bertevas, E., Tran-Duc, T., Khoo, B.C., Phan-Thien, N.: Smoothed particle hydrodynamics (SPH) applications in some sediment dispersion problems. In: Proceedings of the 7th International Conference on Computational Methods (ICCM2016). Berkeley, California, United States (2016)

  • Bertola, N.J., Wang, H., Chanson, H.: Air bubble entrainment at vertical plunging jets: a large-scale experimental study. Tech. rep., School of Civil Engineering, The University of Queensland (2017). Hydraulic Model Report CH, 104/17

  • Biń, A.: Gas entrainment by plunging liquid jets. Chem. Eng. Sci. 48(21), 3585–3630 (1993). https://doi.org/10.1016/0009-2509(93)81019-R

    Article  Google Scholar 

  • Blondel, F., Audebert, B., Pasutto, T., Stanciu, M.: Condensation models and boundary conditions for non-equilibrium wet steam flows. Int. J. Finite 10 (2013)

  • Brattberg, T., Chanson, H.: Air entrapment and air bubble dispersion at two-dimensional plunging water jets. Chem. Eng. Sci. 53(24), 4113–4127 (1998). https://doi.org/10.1016/S0009-2509(98)80004-3

    Article  Google Scholar 

  • Brattberg, T., Chanson, H., Toombes, L.: Experimental investigations of free-surface aeration in the developing flow of two-dimensional water jets. J. Fluids Eng. 120(4), 738–744 (1998). https://doi.org/10.1115/1.2820731

    Article  Google Scholar 

  • Chanson, H.: Air bubble diffusion in supercritical open channel flows. In: Proceedings of the 12th Australasian fluid mechanics conference. Sydney, Australia (1995)

  • Chanson, H.: Air Bubble Entrainment in Free-Surface Turbulent Shear Flows. Academic Press, Cambridge (1996). https://doi.org/10.1016/B978-0-12-168110-4.X5000-0

    Book  Google Scholar 

  • Chanson, H.: Caractéristiques diphasiques des écoulements sur les coursiers en marches d’escalier. La Houille Blanche 8, 16–28 (2001). https://doi.org/10.1051/lhb/2001084

    Article  Google Scholar 

  • Chanson, H., Brattberg, T.: Air entrainment by two-dimensional plunging jets: the impingement region and the very-near flow field. In: Proceedings of the 1998 ASME Fluids Engineering Division Summer Meeting. Washington DC, United States (1998)

  • Chanson, H., Toombes, L.: Experimental Investigations of Air Entrainment in Transition and Skimming Flows down a Stepped Chute. Application to Embankment Overflow Stepped Spillways. Tech. Rep. CE158, Department of Civil Engineering, The University of Queensland, Brisbane, Australia (2001)

  • Clift, R., Grace, J.R., Weber, M.E.: Bubbles, Drops and Particles. Dover, Downers (1978)

    Google Scholar 

  • Cueille, P.V.: Modélisation par Smoothed Particle Hydrodynamic des phénomènes de diffusion présents dans un écoulement. Ph.D. thesis, INSA de Toulouse (2005)

  • Denèfle, R.: Modélisation locale diphasique eau-vapeur des écoulements dans les générateurs de vapeur. Ph.D. thesis, Université Bordeaux 1 (2013)

  • Douillet-Grellier, T., Vuyst, F.D., Calandra, H., Ricoux, P.: Simulations of intermittent two-phase flows in pipes using smoothed particle hydrodynamics. Comput. Fluids 177, 101–122 (2018). https://doi.org/10.1016/j.compfluid.2018.10.004

    Article  MathSciNet  MATH  Google Scholar 

  • Español, P., Revenga, M.: Smoothed dissipative particle dynamics. Phys. Rev. E (2003). https://doi.org/10.1103/PhysRevE.67.026705

  • Ferrand, M., Joly, A., Kassiotis, C., Violeau, D., Leroy, A., Morel, F.X., Rogers, B.D.: Unsteady open boundaries for SPH using semi-analytical conditions and Riemann solver in 2D. Comput. Phys. Commun. 210, 29–44 (2017). https://doi.org/10.1016/j.cpc.2016.09.009

    Article  MathSciNet  MATH  Google Scholar 

  • Ferrand, M., Laurence, D.R., Rogers, B.D., Violeau, D., Kassiotis, C.: Unified semi-analytical wall boundary conditions for inviscid, laminar or turbulent flows in the meshless SPH method. Int. J. Numer. Meth. Fluids 71, 446–472 (2013). https://doi.org/10.1002/fld.3666

    Article  MathSciNet  MATH  Google Scholar 

  • Fonty, T., Ferrand, M., Leroy, A., Joly, A., Violeau, D.: An upwind scheme for conservative, realizable two-phase mixture SPH model with high density ratios. In: Proceedings of 13th international SPHERIC workshop. Galway, Ireland (2018)

  • Fonty, T., Ferrand, M., Leroy, A., Joly, A., Violeau, D.: Mixture model for two-phase flows with high density ratios: a conservative and realizable SPH formulation. Int. J. Multiph. Flow 111, 158–174 (2019). https://doi.org/10.1016/j.ijmultiphaseflow.2018.11.007

    Article  MathSciNet  Google Scholar 

  • Ghaïtanellis, A., Violeau, D., Ferrand, M., Abderrezzak, K.E.K., Leroy, A., Joly, A.: A SPH elastic-viscoplastic model for granular flows and bed-load transport. Adv. Water Resour. 111, 156–173 (2018). https://doi.org/10.1016/j.advwatres.2017.11.007

    Article  Google Scholar 

  • Grenier, N., Antuono, M., Colagrossi, A., Touzé, D.L., Alessandrini, B.: An Hamiltonian interface SPH formulation for multi-fluid and free surface flows. J. Comput. Phys. 228(22), 8380–8393 (2009). https://doi.org/10.1016/j.jcp.2009.08.009

    Article  MathSciNet  MATH  Google Scholar 

  • Grenier, N., Touzé, D.L., Colagrossi, A., Colicchio, G., Antuono, M.: SPH multiphase simulation of bubbly flows. Towards oil and water separation. In: Proceedings of the ASME 2013 32nd international conference on ocean, offshore and arctic engineering OMAE 2013, vol. 7. Nantes, France (June 9–14, 2013). https://doi.org/10.1115/OMAE2013-11610

  • Guyot, G., Rodriguez, M.: La Coche Pelton enhancement project scale model. In: Proceedings of the 37th IAHR World Congress. Kuala Lumpur, Malaysia (2017)

  • Hammani, I., Oger, G., Touzé, D.L., Colagrossi, A., Marrone, S.: How to derive the multi-fluid delta-SPH model. In: Proceedings of the 13th international SPHERIC workshop. Galway, Ireland (2018)

  • Hirt, C.W.: Modeling turbulent air entrainment of air at a free surface. Tech. rep., Flow Science, Inc. (2003)

  • Hu, X.Y., Adams, N.A.: A multi-phase SPH method for macroscopic and mesoscopic flows. J. Comput. Phys. 213, 844–861 (2006). https://doi.org/10.1016/j.jcp.2005.09.001

    Article  MathSciNet  MATH  Google Scholar 

  • Hu, X.Y., Adams, N.A.: An incompressible multi-phase SPH method. J. Comput. Phys. 227, 264–278 (2007). https://doi.org/10.1016/j.jcp.2007.07.013

    Article  MATH  Google Scholar 

  • Ishii, M., Hibiki, T.: Thermo-Fluid Dynamics of Two-Phase Flow, 2nd edn. Springer, Berlin (2011)

    Book  Google Scholar 

  • Kiger, K., Duncan, J.H.: Air-entrainment mechanisms in plunging jets and breaking waves. Annu. Rev. Fluid Mech. 44(06), 563–596 (2012). https://doi.org/10.1146/annurev-fluid-122109-160724

    Article  MathSciNet  MATH  Google Scholar 

  • Kobus, H.: Local air entrainment and detrainment. In: Symposium on Scale Effects in Modelling Hydraulic Structures. Stuttgart, Germany (1984). https://doi.org/10.18419/opus-559

  • Kobus, H., Westrich, B.: An example of a combined disharge-control and aeration structure. In: Proceedings of the 20th Congress of the IAHR. Moscow, USSR (1983)

  • Kruisbrink, A.C.H., Pearce, F.R., Yue, T., Morvan, H.P.: An SPH multi-fluid model based on quasi buoyancy for interface stabilization up to high density ratios and realistic wave speed ratios. Int. J. Numer. Meth. Fluids 87(10), 487–507 (2018). https://doi.org/10.1002/fld.4498

    Article  MathSciNet  Google Scholar 

  • Kunz, P., Hirschler, M., Huber, M., Nieken, U.: Inflow/outflow with Dirichlet boundary conditions for pressure in ISPH. J. Comput. Phys. 326, 171–187 (2016). https://doi.org/10.1016/j.jcp.2016.08.046

    Article  MathSciNet  Google Scholar 

  • Labois, M.: Modélisation des déséquilibres mécaniques dans les écoulements diphasiques : approches par relaxation et par modèle réduit. Ph.D. thesis, Université de Provence, Marseille (2008). pp. 24–27

  • Lastiwka, M., Basa, M., Quinlan, N.J.: Permeable and non-reflecting boundary conditions in SPH. Int. J. Numer. Meth. Fluids 61(7), 709–724 (2009). https://doi.org/10.1002/fld.1971

    Article  MathSciNet  MATH  Google Scholar 

  • Launder, B., Spalding, D.B.: The numerical computation of turbulent flow. Comput. Methods Appl. Mech. Eng. 3(2), 269–289 (1974). https://doi.org/10.1016/0045-7825(74)90029-2

    Article  MATH  Google Scholar 

  • Leroy, A.: Un nouveau modèle SPH incompressible : vers l’application à des cas industriels. Ph.D. thesis, Université Paris Est (2014)

  • Leroy, A., Violeau, D., Ferrand, M., Kassiotis, C.: Unified semi-analytical wall boundary conditions applied to 2-d incompressible SPH. J. Comput. Phys. 261, 106–129 (2014). https://doi.org/10.1016/j.jcp.2013.12.035

    Article  MathSciNet  MATH  Google Scholar 

  • Li, S., Zhang, J., Xu, W.: Numerical investigation of air-water flow properties over steep float and pooled stepped spillways. J. Hydraul. Res. 56(1), 1–14 (2018). https://doi.org/10.1080/00221686.2017.1286393

    Article  Google Scholar 

  • Manninen, M., Taivassalo, V.: On the mixture model for multiphase flow, vol. 288. Technical Research Center of Finland (1996)

  • Morris, J.P., Fow, P.J., Zhu, Y.: Modeling low Reynolds number incompressible flows using SPH. J. Comput. Phys. 136(1), 214–226 (1997). https://doi.org/10.1006/jcph.1997.5776

    Article  MATH  Google Scholar 

  • Nikseresht, A.H., Talebbeydokhti, N., Rezaei, M.J.: Numerical simulation of two-phase flow on step-pool spillways. Sci. Iran. 20(2), 222–230 (2013). https://doi.org/10.1016/j.scient.2012.11.013

    Article  Google Scholar 

  • Oger, G.: Aspects théoriques de la méthode SPH et applications à l’hydrodynamique à surface libre. Ph.D. thesis, École Centrale de Nantes (2006)

  • Pope, S.B.: Turbulent Flows. Cambridge University Press, Cambridge (2000). https://doi.org/10.1017/CBO9780511840531

    Book  MATH  Google Scholar 

  • Price, D., Laibe, G.: Two-phase mixtures in SPH–A new approach. In: Proceedings of 10th International SPHERIC Workshop, Parma, Italy, pp. 68–75 (2015a)

  • Price, D.J., Laibe, G.: A fast and explicit algorithm for simulating the dynamics of small dust grains with smoothed particle hydrodynamics. Monthly Notices of the Royal Astronomical Society pp. 1–15 (2015b). https://doi.org/10.1093/mnras/stv996

  • Qu, X.L., Khezzar, L., Danciu, D., Labois, M., Lakehal, D.: Characterization of plunging liquid jets: a combined experimental and numerical investigation. Int. J. Multiph. Flow 37(7), 722–731 (2011). https://doi.org/10.1016/j.ijmultiphaseflow.2011.02.006

    Article  Google Scholar 

  • Ren, B., Li, C., Yan, X., Ling, M.C., Bonet, J., Hu, S.M.: Multiple-fluid SPH simulation using a mixture model. ACM Trans. Graph. 33, 171:1–171:11 (2014). https://doi.org/10.1145/2645703

    Article  MATH  Google Scholar 

  • Rezavand, M., Zhang, C., Hu, X.: Multi-phase simulation of highly violent flows using an SPH method based on a Riemann solver. In: Proceedings of the 14th International SPHERIC Workshop. Exeter, UK (2019)

  • Schiller, L., Naumann, Z.: A Drag Coefficient Correlation. Zeitschrift des Vereins Deutscher Ingenieure Zeitung 77, 318–320 (1935)

    Google Scholar 

  • Shih, T.H., Liou, W.W., Shabbir, A., Yang, Z., Zhu, J.: A new \(k-\epsilon\) eddy-viscosity model for high Reynolds number turbulent flows-model development and validation. Comput. Fluids 24(3), 227–238 (1995). https://doi.org/10.1016/0045-7930(94)00032-T

    Article  MATH  Google Scholar 

  • Simonin, O.: Prediction of the dispersed phase turbulence in particle-laden jets. ASME FED 121, 197–206 (1994)

    Google Scholar 

  • Smoller, J.: Shock Waves and Reaction-Diffusion Equations, 2nd edn. Springer, Berlin (1994)

    Book  Google Scholar 

  • Tafuni, A., Domínguez, J.M., Vacondio, R., Crespo, A.J.C.: A versatile algorithm for the treatment of open boundary conditions in Smoothed particle hydrodynamics GPU models. Comput. Methods Appl. Mech. Eng. 342, 604–624 (2018). https://doi.org/10.1016/j.cma.2018.08.004

    Article  MathSciNet  MATH  Google Scholar 

  • Tran-Duc, T., Phan-Thien, N., Khoo, B.C.: A smoothed particle hydrodynamics (SPH) study of sediment dispersion on the seafloor. Phys. Fluids 29(8), 083302 (2017). https://doi.org/10.1063/1.4993474

    Article  Google Scholar 

  • Vacondio, R., Rogers, B.D., Stansby, P.K., Mignosa, P.: SPH modelling of shallow flow with open boundaries for practical flood simulation. J. Hydraul. Eng. 138(6), 530–541 (2012). https://doi.org/10.1061/(ASCE)HY.1943-7900.0000543

    Article  Google Scholar 

  • Valero, D., Bung, D.B.: Development of the interfacial air layer in the non-aerated region of high-velocity spillway flows. Instabilities growth, entrapped air and influence on the self-aeration onset. Int. J. Multiph. Flow 84, 66–74 (2016). https://doi.org/10.1016/j.ijmultiphaseflow.2016.04.012

    Article  MathSciNet  Google Scholar 

  • Violeau, D.: Dissipative forces for Lagrangian models in computational fluid dynamics and application to smoothed-particle hydrodynamics. Phys. Rev. E 80, (2009). https://doi.org/10.1103/PhysRevE.80.036705

  • Violeau, D., Leroy, A.: On the maximum time step in weakly compressible SPH. J. Comput. Phys. 256, 388–415 (2014). https://doi.org/10.1016/j.jcp.2013.09.001

    Article  MathSciNet  MATH  Google Scholar 

  • Wendland, H.: Piecewise polynomial, positive definite and compactly supported radial functions of minimal degree. Adv. Comput. Math. 4(1), 389–396 (1995). https://doi.org/10.1007/BF02123482

    Article  MathSciNet  MATH  Google Scholar 

  • Yan, X., Jiang, Y.T., Li, C.F., Martin, R.R., Hu, S.M.: Multiphase SPH simulation for interactive fluids and solids. ACM Trans. Gr. 35(4) (2016). https://doi.org/10.1145/2897824.2925897

Download references

Funding

This study was funded by EDF R&D and the French Research Agency (CIFRE Grant Number #2016-0362).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas Fonty.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Appendices

Appendices

1.1 Time Integration of the Continuity Equation

The following derivation follows faithfully (Ferrand et al. 2017), but replacing the density \(\rho\) by the inverse of the volume noted \(\sigma\), making some adjustments when needed. From the continuity Eq. (5), one has:

$$\begin{aligned} \displaystyle {\frac{d \sigma _{a}}{d t}}=-\,\sigma _{a}D_{a}^{\gamma }\{\varvec{j}\} \end{aligned}$$
(54)

with:

$$\begin{aligned} D_{a}^{\gamma }\{\varvec{j}\}=-\,\frac{1}{\gamma _{a}\sigma _{a}}\sum _{b\in \left( \mathcal {F}\cup \mathcal {V}\right) }\theta _{b}\left( \varvec{j}_{a}-\varvec{j}_{b}\right) \cdot \varvec{\nabla }w_{ab}+\frac{1}{\gamma _{a}\sigma _{a}}\sum _{s\in \mathcal {S}}\sigma _{s}\left( \varvec{j}_{a}-\varvec{j}_{s}\right) \cdot \varvec{\nabla }\gamma _{as} \end{aligned}$$
(55)

Let us make the distinction between the Eulerian fluid velocity \(\varvec{J}_{b}\) and the Lagrangian particle velocity \(\varvec{j}_{b}\). Additional terms then appear in the divergence operator computation and lead to:

$$\begin{aligned} \displaystyle {\frac{d \sigma _{a}}{d t}}=\displaystyle {\frac{1}{\gamma _{a}}}\sum _{b\in \left( \mathcal {F}\cup \mathcal {V}\right) }\theta _{b}\left( \varvec{j}_{a}-\varvec{j}_{b}\right) \cdot \varvec{\nabla }w_{ab}-\delta \sigma _{a}^{i/o}-\displaystyle {\frac{1}{\gamma _{a}}}\sum _{s}\sigma _{s}\left( \varvec{j}_{a}-\varvec{j}_{s}\right) \cdot \varvec{\nabla }\gamma _{as}+\frac{\sigma _{a}}{\gamma _{a}}\delta \gamma _{a}^{i/o} \end{aligned}$$
(56)

With the definitions:

$$\begin{aligned} \delta \sigma _{a}^{i/o}&= \frac{1}{\gamma _{a}}\sum _{v\in \mathcal {V}^{i/o}}\theta _{v}\left( \varvec{J}_{v}-\varvec{j}_{v}\right) \cdot \varvec{\nabla }w_{av} \end{aligned}$$
(57)
$$\begin{aligned} \delta \gamma _{a}^{i/o}&= \sum _{s\in \mathcal {S}^{i/o}}\frac{\sigma _{s}}{\sigma _{a}}\left( \varvec{J}_{s}-\varvec{j}_{s}\right) \cdot \varvec{\nabla }\gamma _{as} \end{aligned}$$
(58)

where \(\mathcal {V}^{i/o}\) and \(\mathcal {S}^{i/o}\) are respectively the sets of vertex particles and segments belonging to the open boundaries. In a Lagrangian frame, (29) leads to:

$$\begin{aligned} \displaystyle {\frac{d w_{ab}}{d t}}&= \left( \varvec{j}_{a}-\varvec{j}_{b}\right) \cdot \varvec{\nabla }w_{ab} \end{aligned}$$
(59)
$$\begin{aligned} \displaystyle {\frac{d \gamma _{a}}{d t}}&= \sum _{s\in \mathcal {S}}\left( \varvec{j}_{a}-\varvec{j}_{s}\right) \cdot \varvec{\nabla }\gamma _{as} \end{aligned}$$
(60)

If one makes the approximation:

$$\begin{aligned} \displaystyle {\frac{d \gamma _{a}}{d t}}\approx \sum _{s\in \mathcal {S}}\frac{\sigma _{s}}{\sigma _{a}}\left( \varvec{j}_{a}-\varvec{j}_{s}\right) \cdot \varvec{\nabla }\gamma _{as} \end{aligned}$$
(61)

as volume variations remain limited, (56) now writes:

$$\begin{aligned} \displaystyle {\frac{d \sigma _{a}}{d t}}=\displaystyle {\frac{1}{\gamma _{a}}}\displaystyle {\frac{d }{d t}}\left( \sum _{b\in \left( \mathcal {F}\cup \mathcal {V}\right) }\theta _{b}w_{ab}\right) -\delta \sigma _{a}^{i/o}-\displaystyle {\frac{\sigma _{a}}{\gamma _{a}}}\displaystyle {\frac{d \gamma _{a}}{d t}}+\frac{\sigma _{a}}{\gamma _{a}}\delta \gamma _{a}^{i/o} \end{aligned}$$
(62)

Hence:

$$\begin{aligned} \displaystyle {\frac{d }{d t}}\left( \gamma _{a}\sigma _{a}\right) =\gamma _{a}\displaystyle {\frac{d \sigma _{a}}{d t}}+\sigma _{a}\displaystyle {\frac{d \gamma _{a}}{d t}}=\displaystyle {\displaystyle {\frac{d }{d t}}}\left( \sum _{b\in \left( \mathcal {F}\cup \mathcal {V}\right) }\theta _{b}w_{ab}\right) -\gamma _{a}\delta \sigma _{a}^{i/o}+\sigma _{a}\delta \gamma _{a}^{i/o} \end{aligned}$$
(63)

The temporal integration of the continuity equation between \(t^{n}\) and \(t^{n+1}\) leads to:

$$\begin{aligned} \left( \gamma _{a}\sigma _{a}\right) ^{n+1}-\left( \gamma _{a}\sigma _{a}\right) ^{n}=\sum _{b\in \left( \mathcal {F}\cup \mathcal {V}\right) }\left( \theta _{b}^{n+1}w_{ab}^{n+1}-\theta _{b}^{n}w_{ab}^{n}\right) -\int _{t^{n}}^{t^{n+1}}\gamma _{a}\delta \sigma _{a}^{i/o}+\int _{t^{n}}^{t^{n+1}}\sigma _{a}\delta \gamma _{a}^{i/o} \end{aligned}$$
(64)

With the virtual displacement \(\delta r_{a}^{i/o}=\delta t\left( \varvec{J}_{a}^{n}-\varvec{j}_{a}^{n}\right)\), the virtual variations terms write:

$$\begin{aligned} \int _{t^{n}}^{t^{n+1}}\gamma _{a}\delta \sigma _{a}^{i/o}&= \sum _{v\in \mathcal {V}^{I/O}}\theta _{v}^{n}\left( w\left( \varvec{r}_{av}^{n}+\delta \varvec{r}_{v}^{i/o}\right) -w\left( \varvec{r}_{av}^{n}\right) \right) \end{aligned}$$
(65)
$$\begin{aligned} \int _{t^{n}}^{t^{n+1}}\sigma _{a}\delta \gamma _{a}^{i/o}&= \frac{1}{2}\sum _{s\in \mathcal {S}^{I/O}}\sigma _{s}^{n}\left( \varvec{\nabla }\gamma _{as}\left( \varvec{r}_{as}^{n}+\delta \varvec{r}_{s}^{i/o}\right) +\varvec{\nabla }\gamma _{as}\left( \varvec{r}_{as}^{n}\right) \right) \cdot \delta \varvec{r}_{s}^{i/o} \end{aligned}$$
(66)

where \(\delta r_{s}^{i/o}=\delta t\left( \varvec{J}_{s}^{n}-\varvec{j}_{s}^{n}\right)\) and \(\delta r_{v}^{i/o}=\delta t\left( \varvec{J}_{v}^{n}-\varvec{j}_{v}^{n}\right)\). In the present work, the factor \(\sigma _{s}^{n}\) was approximated by \(\sigma _{a}^{n}\), consistently with the approximation (61). For an analytical computation of \(\gamma _{a}\):

$$\begin{aligned} \int _{t^{n}}^{t^{n+1}}\sigma _{a}\delta \gamma _{a}^{i/o}=\sum _{s\in \mathcal {S}^{I/O}}\sigma _{s}^{n}\left( \gamma _{as}\left( \varvec{r}_{as}^{n}+\delta \varvec{r}_{s}^{i/o}\right) -\gamma _{as}\left( \varvec{r}_{as}^{n}\right) \right) \end{aligned}$$
(67)

1.2 Riemann Problem Formulation

We follow here a similar reasoning to Ferrand et al. (2017). Neglecting the relative velocity and the viscous efforts at the boundary, the homogeneous (i.e. no relative velocity) mixture model projected along the wall normal (n index refers to the normal component and \(\tau\) index to the tangential component) writes in non-conservative form:

$$\begin{aligned} \displaystyle {\frac{\partial }{\partial t}}\varvec{W}+\varvec{B}\left( \varvec{W}\right) \displaystyle {\frac{\partial \varvec{W}}{\partial \varvec{n}}}=0 \end{aligned}$$
(68)

where

$$\begin{aligned} \varvec{W}=\left( \begin{array}{c} \sigma \\ \alpha \\ j_{n} \\ j_{\tau } \end{array}\right) \text { and } \varvec{B}\left( \varvec{W}\right) =\left( \begin{array}{cccc} j_{n} &{} 0 &{} \sigma &{} 0 \\ 0 &{} j_{n} &{} 0 &{} 0 \\ \displaystyle {}\frac{c^{2}}{\sigma } &{} D &{} j_{n} &{} 0 \\ 0 &{} 0 &{} 0 &{} j_{n} \end{array}\right) \end{aligned}$$

where \(D=\frac{1}{\rho }\left( \rho ^{\alpha }\left( c^{\alpha }\right) ^{2}-\rho ^{\beta }\left( c^{\beta }\right) ^{2}\right) \left( \frac{\sigma }{\sigma _{0}}-1\right)\) is deduced from the linearized state law (28). The eigenvalues of \(\varvec{B}\) are therefore \(j_{n}\) (of multiplicity 2) and \(j_{n}\pm c\) where the sound speed writes:

$$\begin{aligned} c=\sqrt{\frac{\alpha \rho ^{\alpha }\left( c^{\alpha }\right) ^{2}+\beta \rho ^{\beta }\left( c^{\beta }\right) ^{2}}{\alpha \rho ^{\alpha }+\beta \rho ^{\beta }}} \end{aligned}$$
(69)

Following Smoller (1994), one can compute the associated k-Riemann invariants that are detailed in Table 10 where

$$\begin{aligned} \psi \left( \alpha ,\sigma \right) =c\left( \alpha \right) \ln \left( \frac{\sigma }{\sigma _{0}}\right) \end{aligned}$$
(70)

As illustrated on Fig. 25, depending on the sign of \(\lambda _{0}\), the state of the boundary segment is determined by the first or second state. As flows considered are subsonic, \(\lambda _{-1}<0\) and \(\lambda _{+1}>0\). \(\lambda _{-1}\) is considered as a ghost wave so that data at exterior state and state 1 are considered to be equal. \(\lambda _{0}\) is a contact discontinuity so that the associated k-Riemann invariants hold and simplify in the relations \(j_{n,1}=j_{n,2}\) and \(c\left( \alpha _{1}\right) \ln \left( \frac{\sigma _{1}}{\sigma _{0}}\right) =c\left( \alpha _{2}\right) \ln \left( \frac{\sigma _{2}}{\sigma _{0}}\right)\). To relate the fluid velocity along the normal of the segment and the pressure, one needs to find the relation between state 2 and the interior state (determined through a SPH interpolation with a renormalization by a Shepard filter) across the wave \(\lambda _{+1}\). The characteristic wave \(\lambda _{+1}\) can correspond to two possible discontinuities:

  • Expansion wave: characteristics are diverging, a smooth transition connects both states and:

    $$\begin{aligned} \lambda _{+1,2}<\lambda _{+1,int}\rightarrow \text { Riemann Invariant } \end{aligned}$$
    (71)
  • Shock wave: characteristics are diverging, a smooth transition connects both states and:

    $$\begin{aligned} \lambda _{+1,2}>\lambda _{+1,int}\rightarrow \text { Rankine-Hugoniot } \end{aligned}$$
    (72)

For the shock wave, the Rankine-Hugoniot jump conditions expressed here in terms of conservative variables write:

$$\begin{aligned} S\left( \rho _{2}-\rho _{int}\right)&= \rho _{2}j_{n,2}-\rho _{int}j_{n,int} \end{aligned}$$
(73)
$$\begin{aligned} S\left( \rho _{2}\alpha _{2}-\rho _{int}\alpha _{int}\right)&= \rho _{2}\alpha _{2}j_{n,2}-\rho _{int}\alpha _{int}j_{n,int} \end{aligned}$$
(74)
$$\begin{aligned} S\left( \rho _{2}j_{n,2}-\rho _{int}j_{n,int}\right)&= p_{2}+\rho _{2}j_{n,2}^{2}-p_{int}-\rho _{int}j_{n,int}^{2} \end{aligned}$$
(75)
$$\begin{aligned} S\left( \rho _{2}j_{\tau ,2}-\rho _{int}j_{\tau ,int}\right)&= \rho _{2}j_{n,2}j_{\tau ,2}-\rho _{int}j_{n,int}j_{\tau ,int} \end{aligned}$$
(76)

where S is the speed of the shock. Unknown pressure terms are computed thanks to the state Eq. (28). For the density, we use:

$$\begin{aligned} \frac{\rho }{\alpha \rho ^{\alpha }+\beta \rho ^{\beta }}=\frac{\sigma }{\sigma _{0}} \end{aligned}$$
(77)

Combining (73) and (76) with (74), one gets \(\alpha _{2}=\alpha _{int}\) and \(j_{\tau ,2}=j_{\tau ,int}\). Equations (73) and (75) can be combined to determine \(\rho _{2}\) and then deduce \(u_{n,2}\):

$$\begin{aligned} \rho _{2}\rho _{int}\left( j_{n,2}-j_{n,int}\right) ^{2}=\left( \rho _{2}-\rho _{int}\right) \left( p_{2}-p_{int}\right) \end{aligned}$$
(78)

This relation is explicit if \(\rho _{2}\) is known but implicit if \(j_{n,2}\) is known. One should then iterate. Assuming that \(\rho\) variations are small, we deduced a first guess from the linearization of the relation that proved to be sufficiently accurate. For the other quantities to compute, at an inlet (i.e. \(\lambda _{0}>0\)), tangential velocity and volume fraction are defined by the user. They are determined from the interior state at an outlet (i.e. \(\lambda _{0}<0\)). In pseudo-code, it leads to the schemes:

figure a
Fig. 25
figure 25

Riemann problems configurations (Ferrand et al. 2017). The state imposed at the boundary is underlined

Table 10 k-Riemann invariants for each eigenvalue

Analytical Solution for the Two-Phase Mixture Poiseuille Flow

At steady state, the volume fraction equation of (6) becomes:

$$\begin{aligned} \varvec{\nabla }\cdot \left( \alpha \beta \varvec{v}^{r}\right) =0 \end{aligned}$$
(79)

Under the longitudinal periodicity condition, it writes:

$$\begin{aligned} \displaystyle {\frac{d }{d z}}\left( \alpha \beta \varvec{v}^{r}\cdot \varvec{e}_{z}\right) =0 \end{aligned}$$
(80)

The no-flux conditions at the upper and lower walls imply that at steady state the Eq. (80) becomes after integration:

$$\begin{aligned} \varvec{v}^{r}\cdot \varvec{e}_{z}=0 \end{aligned}$$
(81)

The volume fraction equation will therefore depend on the chosen closure on the relative velocity. Starting with the closure \(\varvec{v}^{r}=\varvec{v}_{0}-K\varvec{\nabla }\alpha /\alpha\), the momentum equation of (3) becomes in this framework:

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle {\frac{d p}{d z}}=-\,\rho g \\ \displaystyle {\frac{d }{d z}}\left( \rho \nu \displaystyle {\frac{d }{d z}}\varvec{j}\cdot \varvec{e}_{x}\right) +\rho \varvec{F}\cdot \varvec{e}_{x}=0 \end{array}\right. \end{aligned}$$
(82)

In the simplified momentum Eq. (82), the dynamic viscosity is variable and depends on the volume fraction solution of Eq. (81). Let us nondimensionalize the system using \(z_{\star }=z/e\), \(j_{\star }=\varvec{j}\cdot \varvec{e}_{x}/U_{m0}\) with \(U_{m0}=\rho ^{\beta }Fe^{2}/\left( 3\mu ^{\beta }\right)\) (discharge for the usual single-fluid Poiseuille flow), \(p_{\star }=p/\left( \rho ^{\beta }ge\right)\) and introduce the Péclet number \(\textsf {Pe}=e|\varvec{v}_{0}\cdot \varvec{e}_{z}|/K\) as the ratio of convective and diffusive transports. Noting the density ratio \(R_{\rho }=\left( \rho ^{\alpha }-\rho ^{\beta }\right) /\rho ^{\beta }\) and viscosity ratio \(R_{\mu }=\left( \mu ^{\alpha }-\mu ^{\beta }\right) /\mu ^{\beta }\), the system becomes:

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle {\frac{d \alpha }{d z_{\star }}}=\textsf {Pe}\,\alpha \\ \displaystyle {\frac{d p_{\star }}{d z_{\star }}}=-\,\left( 1+R_{\rho }\alpha \right) \\ \displaystyle {\frac{d }{d z_{\star }}}\left[ \left( 1+R_{\mu }\alpha \right) \displaystyle {\frac{d j_{\star }}{d z_{\star }}}\right] =-\,3\left( 1+R_{\rho }\alpha \right) \end{array}\right. \end{aligned}$$
(83)

Constant dynamic viscosity The nondimensionalized solution for a constant dynamic viscosity \(\mu =\mu ^{\alpha }=\mu ^{\beta }\) writes:

  • Volume fraction

    $$\begin{aligned} \alpha \left( z_{\star }\right) =\alpha _{1}\exp \left( \textsf {Pe}\,z_{\star }\right) \end{aligned}$$
    (84)
  • Pressure profile

    $$\begin{aligned} p_{\star }\left( z_{\star }\right) =p_{B\star }+1-z_{\star }+\alpha _{1}R_{\rho }\frac{\exp \left( \textsf {Pe}\right) }{\textsf {Pe}}\left[ 1-\exp \left( -\textsf {Pe}\left( 1-z_{\star }\right) \right) \right] \end{aligned}$$
    (85)

    where \(p_{B\star }\) is the background pressure nondimensionalized as the total pressure.

  • Longitudinal velocity

    $$\begin{aligned} j_{\star }(z_{\star })&= \displaystyle {}\frac{3}{2}\left( 1-z_{\star }^{2}\right) +\frac{3}{\textsf {Pe}^{2}}\left[ \textsf {Li}_{2,R}\left( z_{\star }\right) +\textsf {Pe}\,z_{\star }\ln _{R}\left( z_{\star }\right) \right. \nonumber \\&\displaystyle {}\left. -\frac{r}{R}\ln _{R}\left( z_{\star }\right) +C_{1}\left( \ln _{R}\left( z_{\star }\right) -\textsf {Pe}\,z_{\star }\right) -C_{2}\right] \end{aligned}$$
    (86)

    where \(\textsf {Li}_{2}\) is the dilogarithm function that writes \(\textsf {Li}_{2}(x)=-\,\int _{0}^{x}\ln (1-t)/t\,dt\)Abramowitz and Stegun (1964). If \(|x|\le 1\), one can write a series expression \(\textsf {Li}_{2}(x)=\sum _{n}x^{n}/n^{2}\). We introduced the notations \(\ln _{R}\left( z_{\star }\right) =\ln \left( 1+\alpha _{1}R_{\mu }\exp \left( \textsf {Pe}\,z_{\star }\right) \right)\) and \(\textsf {Li}_{2,R}\left( z_{\star }\right) =\textsf {Li}_{2}\left( -\alpha _{1}R_{\mu }\exp \left( \textsf {Pe}\,z_{\star }\right) \right)\). \(C_{1}\) and \(C_{2}\) are deduced from the no-slip condition at walls:

    $$\begin{aligned} C_{1}&= \displaystyle {}\frac{\textsf {Li}_{2,R}(1)-\textsf {Li}_{2,R}(-1)+\textsf {Pe}\left[ \ln _{R}\left( 1\right) +\ln _{R}\left( -1\right) \right] -R_{\rho }/R_{\mu }\left[ \ln _{R}\left( 1\right) -\ln _{R}\left( -1\right) \right] }{2\textsf {Pe}-\ln _{R}\left( 1\right) +\ln _{R}\left( -1\right) } \end{aligned}$$
    (87)
    $$\begin{aligned} C_{2}&= \textsf {Li}_{2,R}(1)+\textsf {Pe}\ln _{R}(1)-\frac{R_{\rho }}{R_{\mu }}\ln _{R}(1)+C_{1}\left( \ln _{R}(1)-\textsf {Pe}\right) \end{aligned}$$
    (88)

\(\alpha _{1}\) is computed thanks to the conservation of volume (integrating over the height of the channel) for a given initial uniform profile of \(\alpha \left( z\right) =\alpha _{0}\):

$$\begin{aligned} \frac{\alpha _{1}}{\alpha _{0}}=\frac{\textsf {Pe}}{\sinh \left( \textsf {Pe}\right) } \end{aligned}$$
(89)

To avoid complete separation of phases, (84) gives a condition on \(\alpha _{1}\):

$$\begin{aligned} 0\le \alpha _{1}\le \exp \left( -\textsf {Pe}\right) \end{aligned}$$
(90)

And therefore a condition on the initial uniform volume fraction \(\alpha _{0}\) using (89):

$$\begin{aligned} 0\le \alpha _{0}\le \frac{1-\exp \left( -2\textsf {Pe}\right) }{2\textsf {Pe}} \end{aligned}$$
(91)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fonty, T., Ferrand, M., Leroy, A. et al. Air Entrainment Modeling in the SPH Method: A Two-Phase Mixture Formulation with Open Boundaries. Flow Turbulence Combust 105, 1149–1195 (2020). https://doi.org/10.1007/s10494-020-00165-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10494-020-00165-7

Keywords

Navigation