Skip to main content
Log in

Modeling of unsaturated granular flows by a two-layer approach

  • Research Paper
  • Published:
Acta Geotechnica Aims and scope Submit manuscript

Abstract

Flows of partially saturated grain-fluid mixtures over complex curved topography are commonly observed in nature. However, comprehensive understanding of the physics behind them is to date out of reach. To investigate their dynamic process, a two-layer approach is proposed, in which the fluid-saturated granular layer is overlaid by the pure granular material. More specifically, the lower layer is described by a two-phase mixture theory of density preserving solid and fluid constituents. For the upper layer, the single-phase granular mass is treated as a frictional Coulomb-like continuum, and the dilation effect and the influence of the interstitial air are ignored. The capillarity effects and grains-size segregation are not considered in both the layers. The lower and upper layers interact at an interface which is a material surface for the fluid phase, but across which the mass exchange for the granular phase may take place. The granular mass exchange across the layer interface is parameterized by an entrainment type postulate. In addition, the classical jump conditions are employed to connect both layers at the interface dynamically. Furthermore, we perform the depth-averaged technique for the saturated grain-fluid mixture lower layer and the pure granular upper layer, respectively, to simplify the governing equations established. It is demonstrated that the resulting model equations can be reduced to most of the existing single-layer pure granular flow models and saturated two-phase single-layer debris flow models. Numerical solutions demonstrate that the present two-layer model can describe flows of partially saturated grain-fluid mixtures and the transition process of a saturated grain-fluid mixture into an under-saturated state.

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

Similar content being viewed by others

References

  1. Berzi D (2008) Approximate analytical solutions in a model for highly concentrated granular-fluid flows. Phy Rev E 78:011304

    Article  Google Scholar 

  2. Berzi D, Jenkins JT (2008) A theoretical analysis of free-surface flows of saturated granular-liquid mixtures. J Fluid Mech 608:393–410

    Article  MathSciNet  MATH  Google Scholar 

  3. Berzi D, Jenkins JT (2009) Steady inclined flows of granular-fluid mixtures. J Fluid Mech 641:359–387

    Article  MATH  Google Scholar 

  4. Bouchut F, Fernandez-Nieto ED, Mangeney A, Narbona-Reina G (2015) A two-phase shallow debris flow model with energy balance. Math Model Numer Anal 49:101–140

    Article  MathSciNet  MATH  Google Scholar 

  5. Chiou MC, Wang Y, Hutter K (2005) Influence of obstacles on rapid granular flows. Acta Mech 175:105–122

    Article  MATH  Google Scholar 

  6. Fernández-Nieto ED, Bouchut F, Bresch D, Castro Díaz MJ, Mangeney A (2008) A new Savage–Hutter type model for submarine avalanches and generated tsunami. J Comput Phys 227:7720–7754

    Article  MathSciNet  MATH  Google Scholar 

  7. Gray JMNT, Tai YC (1998) On the inclusion of a velocity-dependent basal drag in avalanche models. Ann Glaciol 26:277–280

    Article  Google Scholar 

  8. Gray JMNT, Wieland M, Hutter K (1999) Gravity-driven free surface flow of granular avalanches over complex basal topography. Proc R Soc A 445:1841–1874

    Article  MathSciNet  MATH  Google Scholar 

  9. Gray JMNT, Kokelaar BP (2010) Large particle segregation, transport and accumulation in granular free-surface flows. J Fluid Mech 652:105–137

    Article  MathSciNet  MATH  Google Scholar 

  10. Gray JMNT, Edwards AN (2014) A depth-averaged \(\mu (I)\)-rheology for shallow granular free-surface flows. J Fluid Mech 755:503–534

    Article  MathSciNet  MATH  Google Scholar 

  11. Gray JMNT, Ancey C (2011) Multi-component particle-size segregation in shallow granular avalanches. J Fluid Mech 678:535–588

    Article  MATH  Google Scholar 

  12. George DL, Iverson RM (2014) A depth-averaged debris-flow model that includes the effects of evolving dilatancy. II. Numerical predictions and experimental tests. Proc R Soc A 470:20130820

    Article  MathSciNet  Google Scholar 

  13. Hungr O (2003) Flow slides and flows in granular soils. In: Proceedings of international workshop flows

  14. Hutter K, Greve R (1993) Two-dimensional similarity solutions for finite-mass granular avalanches with Coulomb- and viscous-type frictional resistance. J Glaciol 39:357–372

    Article  Google Scholar 

  15. Hutter K, Jöhnk K, Svendsen B (1994) On interfacial transition conditions in two phase gravity flow. J Appl Math Phys 45:746–762

    Article  MathSciNet  MATH  Google Scholar 

  16. Hutter K, Luca I (2012) Two-layer debris mixture flows on arbitrary terrain with mass exchange at the base and the interface. Contin Mech Thermodyn 24:525–558

    Article  MathSciNet  MATH  Google Scholar 

  17. Hutter K, Siegel M, Savage SB, Nohguchi Y (1993) Two-dimensional spreading of a granular avalanche down an inclined plane Part I. Theory. Acta Mech 100:37–68

    Article  MathSciNet  MATH  Google Scholar 

  18. Iverson RM (1997) The physics of debris flows. Rev Geophys 35:245–296

    Article  Google Scholar 

  19. Iverson RM, George DL (2014) A depth-averaged debris-flow model that includes the effects of evolving dilatancy. I. Physical basis. Proc R Soc A 470:1–31

    Article  MathSciNet  Google Scholar 

  20. Iverson RM, Denlinger RP (2001) Flow of variable fluidized granular material across three dimensional terrain 1: Coulomb mixture theory. J Geophys Res 106(B1):537–552

    Article  Google Scholar 

  21. Iverson RM, Logan M, LaHusen RG LaHusen, Berti M (2010) The perfect debris flow? aggregated results from 28 large-scale experiments. J Geophys Res 115:F03005

    Article  Google Scholar 

  22. Denlinger RP, Iverson RM (2001) Flow of variable fluidized granular material across three dimensional terrain 2: numerical predictions and experimental tests. J Geophys Res 106(B1):553–566

    Article  Google Scholar 

  23. Johnson CG, Kokelaar BP, Iverson RM, Logan M, LaHusen RG, Gray JMNT (2012) Grain-size segregation and levee formation in geophysical mass flows. J Geophys Res 117(F1):2003–2012

    Article  Google Scholar 

  24. Kowalski J, McElwaine JN (2013) Shallow two-component gravity-driven flows with vertical variation. J Fluid Mech 714:434–462

    Article  MathSciNet  MATH  Google Scholar 

  25. Luca I, Kuo CY, Hutter K, Tai YC (2012) Modeling shallow over-saturated mixtures on arbitrary rigid topography. J Mech 28:523–541

    Article  Google Scholar 

  26. Meng X, Wang Y (2015a) Investigations of gravity-driven two-phase debris flows. In: Wu W (ed) Recent advances in modeling landslides and debris flows, Springer Series in Geomechanics and Geoengineering, pp 119–130

  27. Meng X, Wang Y (2015b) Modelling and numerical simulation of two-phase debris flows. Acta Geotech 11:1027–1045

    Article  Google Scholar 

  28. Morland LW, Sellers S (2001) Multiphase mixtures and singular surfaces. Int J Non-Linear Mech 36:131–146

    Article  MathSciNet  MATH  Google Scholar 

  29. Pouliquen OB, Forterre Y (2002) Friction law for dense granular flows: application to the motion of a mass down a rough inclined plane. J Fluid Mech 453:133–151

    Article  MATH  Google Scholar 

  30. Pitman EB, Le L (2005) A two-fluid model for avalanche and debris flows. Philos Trans R Soc A 363:1573–1601

    Article  MathSciNet  MATH  Google Scholar 

  31. Pudasaini SP (2012) A general two-phase debris flow model. J Geophys Res 117(F3):1–28

    Article  Google Scholar 

  32. Pudasaini SP, Hutter K (2007) Avalanche dynamics: dynamics of rapid flows of dense granular avalanches. Springer, Berlin

    Google Scholar 

  33. Pudasaini SP, Wang Y, Hutter K (2005) Modelling debris flows down general channels. Nat Hazards Earth Syst 5:799–819

    Article  Google Scholar 

  34. Savage SB, Hutter K (1989) The motion of a finite mass of granular material down a rough incline. J Fluid Mech 199:177–215

    Article  MathSciNet  MATH  Google Scholar 

  35. Schaeffer DG (1987) Instability in the evolution equations describing incompressible granular flow. J Differ Equ 66:19–50

    Article  MathSciNet  MATH  Google Scholar 

  36. Tai YC, Gray J (1998) Limiting stress states in granular avalanches. Ann Glaciol 26:272–276

    Article  Google Scholar 

  37. Tai YC, Noelle S, Gray J, Hutter K (2002) Shock-capturing and front-tracking methods for granular avalanches. J Comput Phys 175:269–301

    Article  MATH  Google Scholar 

  38. Turnbull B, Bowman ET, McElwaine JN (2015) Debris flows: experiments and modelling. C R Phys 16:86–96

    Article  Google Scholar 

  39. Wang Y, Hutter K (1999a) Shearing flows in a Goodman–Cowin type granular material—theory and numerical results. Particul. Sci. Technol. 17:97–124

    Article  Google Scholar 

  40. Wang Y, Hutter K (1999b) A constitutive model for multi-phase mixtures and its application in shearing flows of saturated soild-fluid mixtures. Granul Matter 1:163–181

    Article  Google Scholar 

  41. Wang Y, Hutter K (2001) Comparisons of numerical methods with respect to convectively dominated problems. Int J Numer Methods Fluids 37:721–745

    Article  MATH  Google Scholar 

  42. Wang Y, Hutter K, Pudasaini SP (2004) The Savage-Hutter theory: a system of partial differential equations for avalanche flows of snow, debris, and mud. J Appl Math Mech 84(8):507–527

    MathSciNet  MATH  Google Scholar 

  43. Wang Y, Oberlack M (2011) A thermodynamic model of multiphase flows with moving interfaces and contact line. Contin Mech Thermodyn 23:409–433

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

X. Meng, Y. Wang and J-T. Fischer thank the People Programme (Marie Curie Actions) of the European Union’s Seventh Framework Programme FP7/2007-2013 for the financial support through Grant No. 289911.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yongqi Wang.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (mp4 49 KB)

Supplementary material 2 (mp4 185 KB)

Appendices

Appendix 1: Treatment of boundary conditions

  1. 1.

    The kinematic boundary condition (9) at the top free surface can be expanded as

    $$\frac{\partial s}{\partial t} + u_g^s\frac{\partial s}{\partial x} + v_g^s\frac{\partial s}{\partial y} - w_g ^s= 0.$$
    (68)

    where downslope velocity is represented as \(u_g^s\), cross-slope velocity as \(v_g^s\), and normal velocity as \(w_g^s\).

    The down-slope, cross-slope and normal components of the dynamic boundary condition, (10), are respectively

    $$T_{g(xx)}^s\frac{\partial s}{\partial x}+T_{g(xy)}^s\frac{\partial s}{\partial y}-T_{g(xz)}^s =0,$$
    (69)
    $$T_{g(yx)}^s\frac{\partial s}{\partial x}+T_{g(yy)}^s\frac{\partial s}{\partial y}- T_{g(yz)}^s=0,$$
    (70)
    $${T}_{g(zx)}^s\frac{\partial s}{\partial x}+{T}_{g(zy)}^s\frac{\partial s}{\partial y}-{T}_{g(zz)}^s=0.$$
    (71)
  2. 2.

    Similarly, at the bottom, the kinematic boundary condition (11) for each constituent of the lower layer can be expanded as follows:

    $$w_s^b = 0, \qquad\qquad w_f^b = 0.$$
    (72)

    For the solid constituent of the lower-layer, the down-slope, and cross-slope components of the bed friction boundary condition (12) are

    $${T}_{e(xz)}^b =-\mu _s^b\frac{u^b_s}{\mid {{\varvec{u}}}^b_s\mid }{T}_{e(zz)}^b -(k_s^b\mu _s^b{\phi _{s}}^b)u_s^b,$$
    (73)
    $${T}_{e(yz)}^b=-\mu _s^b\frac{v^b_s}{\mid {{\varvec{u}}}^b_s\mid }{T}_{e(zz)}^b -(k_s^b\mu _s^b{\phi _{s}}^b)v_s^b,$$
    (74)

    respectively. Similarly, for the fluid constituent of the lower-layer, the down-slope, and cross-slope components of the bed friction condition (13) are

    $$\tau _{f(xz)}^b=k_f^b\,u_f^b,$$
    (75)
    $$\tau _{f(yz)}^b =k_f^b\,v_f^b,$$
    (76)
  3. 2.

    At the interface, the kinematic boundary condition (14) is expanded as

    $$\frac{\partial h_m}{\partial t}+u_f^i\frac{\partial h_m}{\partial x}+v_f^i\frac{\partial h_m}{\partial y}-w_f^i=0,$$
    (77)

    and the traction-free dynamic boundary condition (17) in the down-slope, cross-slope and normal directions are

    $$(-p^i+\tau _{f(xx)}^i)\frac{\partial s}{\partial x}+\tau _{f(xy)}^i\frac{\partial s}{\partial y}-\tau _{f(xz)}^i =0,$$
    (78)
    $$\tau _{f(yx)}^i\frac{\partial s}{\partial x}+(-p^i+\tau _{f(yy)}^i)\frac{\partial s}{\partial y}- \tau _{f(yz)}^i=0,$$
    (79)
    $$\tau _{f(zx)}^i\frac{\partial s}{\partial x}+\tau _{f(zy)}^i\frac{\partial s}{\partial y}-(-p^i+\tau _{f(zz)}^i)=0,$$
    (80)

    respectively. Relations (78)–(80) indicate that \(p^i=0\) and \(\tau _{f(jk)}^i=0, (j, k\in \{x, y, z\})\).

    The mass jump condition (15) (or (16)) is written as

    $$\begin{aligned} J&= -{\widetilde{\rho }}_s\phi _g\left[ \frac{\partial h_m}{\partial t}+u_s^i\frac{\partial h_m}{\partial x}+v_s^i\frac{\partial h_m}{\partial y}-w_s^i\right] \nonumber \\ &= -{\widetilde{\rho }}_s\phi _g\left[ \frac{\partial h_m}{\partial t}+u_g^i\frac{\partial h_m}{\partial x}+v_g^i\frac{\partial h_m}{\partial y} -w_g^i\right] . \end{aligned}$$
    (81)

    Similarly, the components of the momentum jump condition (18) in the down-slope, cross-slope and normal directions are

    $$-J(u_s^i-u_g^i)+T_{s(xx)}^i\frac{\partial h_m}{\partial x}+{T}_{s(xy)}^i\frac{\partial h_m}{\partial y}-{T}_{s(xz)}^i ={T}_{g(xx)}^i\frac{\partial h_m}{\partial x}+{T}_{g(xy)}^i\frac{\partial h_m}{\partial y}-{T}_{g(xz)}^i,$$
    (82)
    $$-J(v_s^i-v_g^i)+{T}_{s(yx)}^i\frac{\partial h_m}{\partial x}+{T}_{s(yy)}^i\frac{\partial h_m}{\partial y}-{T}_{s(yz)}^i ={T}_{g(yx)}^i\frac{\partial h_m}{\partial x}+{T}_{g(yy)}^i\frac{\partial h_m}{\partial y}-{T}_{g(yz)}^i,$$
    (83)
    $$-J(w_s^i-w_g^i)+{T}_{s(zx)}^i\frac{\partial h_m}{\partial x}+{T}_{s(zy)}^i\frac{\partial h_m}{\partial y}-{T}_{s(zz)}^i ={T}_{g(zx)}^i\frac{\partial h_m}{\partial x}+{T}_{g(zy)}^i\frac{\partial h_m}{\partial y}-{T}_{g(zz)}^i,$$
    (84)

    respectively, where \({T}_{s(jk)}^i={T}_{e(jk)}^i\) (\(j, k\in (x, y, z)\)) due to \(p^i=0\) indicated by (80).

    To rewrite (19) in a component form, the tangential velocity \({{\varvec{u}}}_k^{\tau }\) (\(k\in \{s, g\}\)) is rewritten as \({{\varvec{u}}}_k^{\tau }={{\varvec{u}}}_k^i -({{\varvec{u}}}_k^i\cdot {{\varvec{n}}}^i){{\varvec{n}}}^i\) that yields \({{\varvec{u}}}_s^{\tau }-{{\varvec{u}}}_g^{\tau }={{\varvec{u}}}_s^i-{{\varvec{u}}}_g^i\) due to \({{\varvec{u}}}_s^i\cdot {{\varvec{n}}}^i={{\varvec{u}}}_g^i\cdot {{\varvec{n}}}^i\) from the jump condition (16) together with the continuity assumption of the solid volume fraction across the interface. It follows that the components of relation (19) in the down-slope, cross-slope and normal directions are,

    $$\begin{aligned}&{T}_{e(xx)}^i\frac{\partial h_m}{\partial x}+{T}_{e(xy)}^i\frac{\partial h_m}{\partial y}-{T}_{e(xz)}^i \nonumber \\ &\quad =({{\varvec{n}}}^i\cdot {{{\varvec{T}}}}_e^i{{\varvec{n}}}^i) \left( \frac{\partial h_m}{\partial x}-\frac{u^i_s-u^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i\right) -k_s^i\mu _s^i\phi _g(u^i_s-u^i_g) ,\end{aligned}$$
    (85)
    $$\begin{aligned}&{T}_{e(yx)}^i\frac{\partial h_m}{\partial x}+{T}_{e(yy)}^i\frac{\partial h_m}{\partial y}-{T}_{e(yz)}^i \nonumber \\ &\quad = ({{\varvec{n}}}^i\cdot {{{\varvec{T}}}}_e^i{{\varvec{n}}}^i)\left( \frac{\partial h_m}{\partial y}-\frac{v^i_s-v^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i\right) -k_s^i\mu _s^i\phi _g(v^i_s-v^i_g) , \end{aligned}$$
    (86)
    $$\begin{aligned}&{T}_{e(zx)}^i\frac{\partial h_m}{\partial x}+{T}_{e(zy)}^i\frac{\partial h_m}{\partial y}-{T}_{e(zz)}^i\nonumber \\ &\quad = ({{\varvec{n}}}^i\cdot {{{\varvec{T}}}}_e^i{{\varvec{n}}}^i)\left( -1- \frac{w^i_s-w^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i\right) - k_s^i\mu _s^i\phi _g(w^i_s-w^i_g), \end{aligned}$$
    (87)

    respectively.

Appendix 2: Depth-averaged procedure

1.1 Conservation equations for mass

Integrating the mass equation (1) of the upper-layer granular phase by using Leibniz rule to swap the orders of integration and differentiation, one can obtain

$$\begin{aligned}&\frac{\partial }{\partial t}(h_g{\widetilde{\rho }}_s\phi _g)+\frac{\partial }{\partial x}( h_g{\widetilde{\rho }}_s\phi _g{\overline{u}}_g)+\frac{\partial }{\partial y}( h_g{\widetilde{\rho }}_s\phi _g{{\overline{v}}}_g)\nonumber \\ &\quad -\left[ {\widetilde{\rho }}_s\phi _g\left( \frac{\partial z}{\partial t}+u_g\psi \frac{\partial z}{\partial x}+v_g\frac{\partial z}{\partial y}-w_g\right) \right] _i^s=0, \end{aligned}$$
(88)

where the square bracket denotes the difference of the top and interfacial values of a function q, i.e., \([q]_i^s=q^s-q^i\). The term within \([\cdot ]\) can be further handled by the kinematic boundary condition (68) at the top surface, and jump condition (81), which leads to

$$\frac{\partial }{\partial t}(h_g{\widetilde{\rho }}_s\phi _g)+\frac{\partial }{\partial x}( h_g{\widetilde{\rho }}_s\phi _g{\overline{u}}_g)+\frac{\partial }{\partial y}( h_g{\widetilde{\rho }}_s\phi _g{{\overline{v}}}_g) =J.$$
(89)

Similarly, integrating (3) and (4) can yield the depth-averaged mass-conservation equations for the solid and fluid constituents of the lower layer. They are given by

$$\frac{\partial }{\partial t}({h}_m{\overline{\rho }}_s)+\frac{\partial }{\partial x}({h}_m\overline{\rho _su_s})+\frac{\partial }{\partial y}({h}_m\overline{\rho _sv_s}) =-J,$$
(90)
$$\frac{\partial }{\partial t}({h}_m{\overline{\rho }}_f) + \frac{\partial }{\partial x}({h}_m\overline{\rho _fu_f})+\frac{\partial }{\partial y}({h}_m\overline{\rho _fv_f})=0,$$
(91)

where \({\overline{\rho }}_s={\widetilde{\rho }}_s{\overline{\phi }}_s\) and \({\overline{\rho }}_f={\widetilde{\rho }}_f{\overline{\phi }}_f\). We follow Pitman and Le [30] and Iverson and Denlinger [20] to assume that volume fractions are uniformly distributed along the depth, such that equations (90) and (91) can be transformed into

$$\frac{\partial }{\partial t}({h}_m{\overline{\rho }}_s)+\frac{\partial }{\partial x}({h}_m{\overline{\rho }}_s{\overline{u}}_s)+\frac{\partial }{\partial y}({h}_m{\overline{\rho }}_s{{\overline{v}}}_s) =-J,$$
(92)
$$\frac{\partial }{\partial t}({h}_m{\overline{\rho }}_f) + \frac{\partial }{\partial x}({h}_m{\overline{\rho }}_f{\overline{u}}_f)+\frac{\partial }{\partial y}({h}_m{\overline{\rho }}_f{{\overline{v}}}_f)=0.$$
(93)

1.2 Normal component of momentum equations

Since the Mohr–Coulomb friction criterion, \(\mid {\varvec{S}}\mid =N\tan \varphi\), does not explicitly show connections between stress and strain rate tensor, we therefore use shallow-water approximation to simplify the conservation equations. The shallow-water approximation indicates that the typical depth \({\mathcal {H}}\) is much smaller than the typical length \({\mathcal {L}}\), i.e., the geometrical aspect ratio \(\epsilon ={\mathcal {H}}/{\mathcal {L}}\ll 1\) is a small quantity. In classical pure granular flow model (see [34]), the normal momentum balance usually reduces to hydrostatic force balance, when shallow-water approximation is employed. Consequently, the upper-layer granular normal momentum equation reduces to

$$\frac{\partial T_{g(zz)}}{\partial z}={\widetilde{\rho }}_s\phi _gg_z,$$
(94)

where \(g_z=g\cos \zeta\). Relation (94) can be vertically integrated from any position z to the top free surface, which yields

$$T_{g(zz)}={\widetilde{\rho }}_s\phi _gg_z(s-z)$$
(95)

by using traction-free condition at the free surface.

Similarly, in most of the previous two-phase grain-fluid mixture models, the fluid normal momentum balance implies that the pore pressure is hydrostatic, and the solid normal momentum balance indicates a balance of gravity, internal stress, and buoyancy force, when shallow-water approximation is employed to streamline the normal momentum balances. Consequently, the lower-layer fluid and solid normal momentum balances reduce to

$$\phi _f\frac{\partial p}{\partial z} + {\widetilde{\rho }}_f\phi _fg_z= 0,$$
(96)
$${\widetilde{\rho }}_s{\phi _{s}}g_z + \frac{\partial T_{e(zz)}}{\partial z} +{\phi _{s}}\frac{\partial p}{\partial z}= 0,$$
(97)

respectively. Integrating relation (96) from any position z in the lower-layer mixture to the interface can yield

$$p(z) = {\widetilde{\rho }}_fg_z (h_m-z),$$
(98)

by using the traction-free condition \(p^i=0\) at the interface. Similarly, vertical integration of (97) leads to

$$T_{e(zz)} = ({\widetilde{\rho }}_s-{\widetilde{\rho }}_f){\overline{\phi }}_sg_z (h_m-z)+T_{e(zz)}^i,$$
(99)

where the stress \(T_{e(zz)}^i\) at the interface remains to be determined. To evaluate \(T_{e(zz)}^i\), we appeal to relation (84). In shallow granular flow models, the derivatives \(\partial h_m/\partial x\) and \(\partial h_m/\partial y\) are small quantities, and the magnitudes of the normal velocities, \(w_\eta\) (\(\eta \in \{g, s, f\}\)), are smaller than the magnitudes of the horizontal velocities, \(u_\eta\) and \(v_\eta\). With these descriptions, relation (84) reduces to

$$T_{e(zz)}^i=T_{s(zz)}^i = T_{g(zz)}^i = {\widetilde{\rho }}_s\phi _gh_gg_z.$$
(100)

Substitution of relation (100) into (99) leads to

$$T_{e(zz)} = ({\widetilde{\rho }}_s-{\widetilde{\rho }}_f){\overline{\phi }}_sg_z (h_m-z)+ {\widetilde{\rho }}_s\phi _gh_gg_z.$$
(101)

1.3 Horizontal momentum-conservation equations

  1. 1.

    For the pure granular phase in the upper layer, integrating downslope component of equation (2) throughout the upper layer by using Leibniz rule to interchange the orders of integration and differentiation can yield the depth-averaged form. The left-hand side terms turn into

    $$\begin{aligned} \int _{h_m}^s\hbox{LHS}\,\hbox{d}z &=\frac{\partial }{\partial t}(h_g{\widetilde{\rho }}_s\phi _g{\overline{u}}_g)+\frac{\partial }{\partial x}(h_g {\widetilde{\rho }}_s\phi _g{\overline{u}}_g^2)+\frac{\partial }{\partial y}(h_g{\widetilde{\rho }}_s\phi _g{\overline{u}}_g{{\overline{v}}}_g)-Ju_g^i , \end{aligned}$$
    (102)

    where we used plug-flow assumption as previous studies did (see [30]). The right-hand side terms turn into

    $$\begin{aligned} \int _{h_m}^s \hbox{RHS}\,\hbox{d}z&= {\widetilde{\rho }}_g\phi _gh_gg_x -\frac{\partial (h_g{\overline{T}}_{g(xy)})}{\partial y}-\frac{\partial ( h_g{\overline{T}}_{g(xx)})}{\partial x}\nonumber \\ &+\left[ {T}_{g(xx)}\frac{\partial z}{\partial x} +T_{g(xy)}\frac{\partial z}{\partial y}-{T}_{g(xz)}\right] _{h_m}^s, \end{aligned}$$
    (103)

    where \(g_x=g\sin \zeta\), and the term \(-\partial ( h_g{\overline{T}}_{g(xy)})/{\partial y}\) is usually ignored, since it is a high-order term with respect to aspect ratio \(\epsilon\) (see [8]). The gradient of lateral stress, \(-\partial (h_g{\overline{T}}_{g(xx)})/{\partial x}\), can be determined through the following [34]. We postulate that \(T_{g(xx)}\) is proportional to \(T_{g(zz)}\) throughout the upper-layer depth, i.e., \(T_{g(xx)}=K_x^gT_{g(zz)}\), where \(K_x^g\) is an earth pressure coefficient and defined in (28). Then, \({\overline{T}}_{g(xx)}\) can be derived by vertically integrating relation \(T_{g(xx)}=K_x^gT_{g(zz)}\). They result in

    $$\begin{aligned} \frac{\partial (h_g{\overline{T}}_{g(xx)})}{\partial x}=\frac{\partial }{\partial x}\left( \frac{1}{2}K^g_x h_g^2{\widetilde{\rho }}_s\phi _gg_z\right) . \end{aligned}$$
    (104)

    Moreover, the terms in \([\cdot ]\) of relation (103) can reduce to

    $$\begin{aligned} \left[ {T}_{g(xx)}\frac{\partial z}{\partial x} +{T}_{g(xy)}\frac{\partial z}{\partial y}-{T}_{g(xz)}\right] _{h_m}^s&=-({{\varvec{n}}}^i\cdot {{{\varvec{T}}}}_e^i{{\varvec{n}}}^i)\left( \frac{\partial h_m}{\partial x}-\frac{u^i_s-u^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i\right) +J(u_s^i-u_g^i)\nonumber \\ &\quad +k_s^i\mu _s^i\phi _g(u^i_s-u^i_g), \end{aligned}$$
    (105)

    by using boundary conditions (69), (82), and (85), where the normal stress \(({{\varvec{n}}}^i\cdot {{{\varvec{T}}}}_e^i{{\varvec{n}}}^i)\) at the interface remains to be determined. In shallow flows, the normal stress \(({{\varvec{n}}}^i\cdot {{{\varvec{T}}}}_e^i{{\varvec{n}}}^i)\) can be approximated as normal traction, \(T_{e(zz)}^i\), plus a centrifugal force term caused by curvature, see [8]. We assume the same curvature between the interface and the bottom that indicates that the centrifugal force is \(h_g{\widetilde{\rho }}_s\phi _g\kappa {\overline{u}}_g^2\). These descriptions yield

    $$\begin{aligned} ({{\varvec{n}}}^i\cdot {{{\varvec{T}}}}_e^i{{\varvec{n}}}^i)= h_g{\widetilde{\rho }}_s\phi _gg_z+ h_g{\widetilde{\rho }}_s\phi _g\kappa {\overline{u}}_g^2. \end{aligned}$$
    (106)

    Similarly, integrating the cross-slope component of equation (2) throughout the upper layer can yield the depth-averaged form. The left-hand side terms turn into

    $$\begin{aligned} \int _{h_m}^s\hbox{LHS}\,\hbox{d}z &=\frac{\partial }{\partial t}(h_g{\widetilde{\rho }}_s\phi _g{{\overline{v}}}_g)+\frac{\partial }{\partial x}(h_g{\widetilde{\rho }}_s \phi _g{\overline{u}}_g{{\overline{v}}}_g)+\frac{\partial }{\partial y}(h_g{\widetilde{\rho }}_s\phi _g{{\overline{v}}}_g^2)-Jv_g^i , \end{aligned}$$
    (107)

    and the right-hand side terms are transformed into

    $$\begin{aligned} \int _{h_m}^s\hbox{RHS}\,\hbox{d}z &= -\frac{\partial }{\partial y}\left( \frac{1}{2} K_y^g{\widetilde{\rho }}_sh_g^2\phi _gg_z\right) -{\widetilde{\rho }}_s\phi _gh_gg_z\frac{\partial h_m}{\partial y}+J(v_s^i-v_g^i) \nonumber \\ & \quad +k_s^i\mu _s^i\phi _g(v^i_s-v^i_g) +\frac{v^i_s-v^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i{\widetilde{\rho }}_s h_g\phi _g \left( g_z+\kappa {\overline{u}}_g^2\right) , \end{aligned}$$
    (108)

    where the first term on the right-hand side represents gradient of internal stress; the second term represents the effect of lower-layer height, the third term denote the momentum exchange, and the fourth terms and the fifth term represent viscous and Coulomb friction at the interface.

  2. 2.

    For the solid constituent in the lower layer, integrating downslope component of equation (7) can yield the depth-averaged form. The left-hand side terms turn into

    $$\begin{aligned} \int ^{h_m}_0 \hbox{LHS}\,\hbox{d}z&=\frac{\partial }{\partial t}(h_m{\overline{\rho }}_s{\overline{u}}_s)+\frac{\partial }{\partial x}(h_m {\overline{\rho }}_s{\overline{u}}_s^2)+\frac{\partial }{\partial y}(h_m{\overline{\rho }}_s{\overline{u}}_s{{\overline{v}}}_s)+Ju_s^i, \end{aligned}$$
    (109)

    and the right-hand side terms are transformed into

    $$\begin{aligned} \int ^{h_m}_0 \hbox{RHS}\,\hbox{d}z&= -\frac{\partial (h_m\overline{{T}}_{e(xy)})}{\partial y} -\frac{\partial (h_m{\overline{T}}_{e(xx)})}{\partial x}-\int _0^{h_m}{\phi _{s}}\frac{\partial p}{\partial x}\hbox{d}z \nonumber \\ &+\left[ {T}_{e(xx)}\frac{\partial z}{\partial x} +T_{e(xy)}\frac{\partial z}{\partial y}-T_{e(xz)}\right] _0^{h_m}\nonumber \\ &+C_d h_m{\overline{\phi }}_s{\overline{\phi }}_f({\overline{u}}_f-{\overline{u}}_s) +h_m{\overline{\rho }}_sg_x , \end{aligned}$$
    (110)

    where the lateral stress gradient, \(\partial (h_m\overline{{T}}_{e(xy)})/\partial y\), can be ignored as we mentioned above. The other stress gradient, \(\partial (h_m{\overline{T}}_{e(xx)})/\partial x\), can be determined by introducing earth pressure coefficient. Referring to Savage and Hutter [34] and its modifications (see, e.g., Hutter et al. [17]; Gray et al. [8]), the following relation

    $$\begin{aligned} {T}_{e(xx)}^b=K^s_x{T}_{e(zz)}^b \end{aligned}$$
    (111)

    holds at the base, where the bed earth pressure coefficient \(K^s_x\) is defined in (36).

    Moreover, relation \({T}_{e(xx)}^i={T}_{g(xx)}^i\) is postulated, which does not violate the momentum jump condition (82). Since relation \({T}_{g(xx)}^i=K^g_x{T}_{g(zz)}^i\) holds at the interface, relation \({T}_{e(xx)}^i=K^g_x{T}_{e(zz)}^i\) can be obtained. To proceed, we postulate that horizontal stresses vary linearly with normal stress throughout the depth of the lower layer. However, the factors are linear with z-coordinate instead of constant value used previously (see, e.g., Savage and Hutter [34]). This more flexible approach leads to

    $$\begin{aligned}&{T}_{e(xx)}=\left( \frac{K^g_x-K^s_x}{{h}_m}z+K^s_x\right) {T}_{e(zz)}, \end{aligned}$$
    (112)

    which is a generalization of Savage–Hutter theory. When \(K^g_x=K^s_x\), relation (112) reproduces the form used in the Savage–Hutter theory. Additionally, relation (112) will produce \({T}_{e(xx)}^i=K^g_x{T}_{e(zz)}^i\) at the interface, and (111) at the base.

    Exploiting relations (101) and (112), one can yield

    $$\begin{aligned} \frac{\partial }{\partial x}(h_m{\overline{T}}_{e(xx)})&= \frac{\partial }{\partial x}\left[ \frac{1}{2} h_gh_m{\widetilde{\rho }}_s\phi _gg_z(K^s_x+K^g_x)\right. \nonumber \\ &\quad \left. +\frac{1}{6}({\widetilde{\rho }}_s-{\widetilde{\rho }}_f) g_zh_m^2{\overline{\phi }}_s(2K^s_x+K^g_x)\right] . \end{aligned}$$
    (113)

    Moreover, the term involving fluid pressure gradient, arising in relation (110), can be formulated as

    $$\begin{aligned} -\int _0^{h_m}{\phi _{s}}\frac{\partial p}{\partial x}\hbox{d}z=-{\widetilde{\rho }}_f{\overline{\phi }}_s\frac{\partial }{\partial x}\left( \frac{1}{2}h_m^2g_z\right) , \end{aligned}$$
    (114)

    by using relation (98). The terms in \([\cdot ]\), arising in relation (110), can be simplified as

    $$\begin{aligned}&\left[ {T}_{e(xx)}\frac{\partial z}{\partial x} +T_{e(xy)}\frac{\partial z}{\partial y}-T_{e(xz)}\right] _0^{h_m} \nonumber \\ &\quad =({{\varvec{n}}}^i\cdot {{{\varvec{T}}}}_e^i{{\varvec{n}}}^i) \left( \frac{\partial h_m}{\partial x}-\frac{u^i_s-u^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i\right) -k_s^i\mu _s^i\phi _g(u^i_s-u^i_g)\nonumber \\ &\qquad -\mu _s^b\frac{u_s^b}{\mid {{\varvec{u}}}_s^b\mid }T_{e(zz)}^b -(k_s^b\mu _s^b{\phi _{s}}^b)u_s^b, \nonumber \\ &\quad =-\frac{u^i_s-u^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^ih_g{\widetilde{\rho }}_s\phi _g (g_z+\kappa {\overline{u}}_g^2)+h_g{\widetilde{\rho }}_s\phi _gg_z\frac{\partial h_m}{\partial x}\nonumber \\ &\qquad -k_s^i\mu _s^i\phi _g(u^i_s-u^i_g)-(k_s^b\mu _s^b{\phi _{s}}^b)u_s^b \nonumber \\ &\quad \quad -\frac{v_s^b}{\mid {{\varvec{u}}}_s^b\mid }\mu _s^b \left[ {\widetilde{\rho }}_sg_z\left( h_m{\overline{\phi }}_s (1-\gamma )+h_g\phi _g\right) \right. \nonumber \\ &\qquad \left. +\kappa \left( h_m{\overline{\phi }}_s({\widetilde{\rho }}_s {\overline{u}}_s^2 -{\widetilde{\rho }}_f{\overline{u}}_f^2 )+{\widetilde{\rho }}_sh_g\phi _g{\overline{u}}_g^2\right) \right] , \end{aligned}$$
    (115)

    by substituting boundary conditions (73) and (85). In (115), the overburden pressure \(T_{e(zz)}^b\) at the bottom is written as a normal traction plus a curvature-dependency term, i.e., \(T_{e(zz)}^b={\widetilde{\rho }}_sg_z\left( h_m{\overline{\phi }}_s (1-\gamma )+h_g\phi _g\right) +\kappa \left( h_m{\overline{\phi }}_s({\widetilde{\rho }}_s {\overline{u}}_s^2 -{\widetilde{\rho }}_f{\overline{u}}_f^2 )+{\widetilde{\rho }}_sh_g\phi _g{\overline{u}}_g^2\right)\), since the bed curvature also contributes a part of pressure and furthermore affects the friction at the bottom, see [8].

    Similarly, integrating the cross-slope component of equation (7) yields that the left-hand side terms turn into

    $$\begin{aligned} \int _0^{h_m}\hbox{LHS}\,\hbox{d}z =\frac{\partial }{\partial t}(h_m{\overline{\rho }}_s{{\overline{v}}}_s)+\frac{\partial }{\partial x}(h_m {\overline{\rho }}_s{\overline{u}}_s{{\overline{v}}}_s)+\frac{\partial }{\partial y}(h_m{\overline{\rho }}_s{{\overline{v}}}_s^2)+Jv_s^i, \end{aligned}$$
    (116)

    and the right-hand side terms turn into

    $$\begin{aligned}&\int _0^{h_m}\hbox{RHS}\,\hbox{d}z \nonumber \\ &\quad =-\frac{\partial }{\partial y}\left[ \frac{1}{2} h_gh_m{\widetilde{\rho }}_s\phi _gg_z(K^s_y+K^g_y)+\frac{1}{6}({\widetilde{\rho }}_s -{\widetilde{\rho }}_f) g_zh_m^2{\overline{\phi }}_s(2K^s_y+K^g_y)\right] \nonumber \\ &\qquad -\frac{v_s^b}{\mid {{\varvec{u}}}_s^b\mid }\mu _s^b \left[ {\widetilde{\rho }}_s\cos \zeta \left( h_m{\overline{\phi }}_s (1-\gamma )+h_g\phi _g\right) \right. \nonumber \\ &\qquad \left. +\kappa \left( h_m{\overline{\phi }}_s({\widetilde{\rho }}_s {\overline{u}}_s^2 -{\widetilde{\rho }}_f{\overline{u}}_f^2 )+{\widetilde{\rho }}_sh_g\phi _g{\overline{u}}_g^2\right) \right] \nonumber \\ &\qquad -\frac{v^i_s-v^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^ih_g{\widetilde{\rho }}_s\phi _g (g_z+\kappa {\overline{u}}_g^2) -k_s^i\mu _s^i\phi _g(v^i_s-v^i_g)\nonumber \\ &\qquad +C_d h_m{\overline{\phi }}_s{\overline{\phi }}_f({{\overline{v}}}_f-{{\overline{v}}}_s), \end{aligned}$$
    (117)

    where the first term of the right-hand side represents the cross-slope gradient of effective stress; the second term accounts for the bed Coulomb friction; and the rest terms describe interfacial Coulomb friction, interfacial viscous friction, and viscous drag between the two phases of the lower mixture consecutively.

  3. 3.

    For the fluid phase of the lower layer, vertically integrating the downslope component of equation (6) allows to derive the depth-averaged form. The left-hand side terms turn into

    $$\begin{aligned} \int ^{h_m}_0\hbox{LHS}\,\hbox{d}z =\frac{\partial }{\partial t}({h}_m{\overline{\rho }}_f{\overline{u}}_f)+\frac{\partial }{\partial x}(h_m {\overline{\rho }}_f{\overline{u}}_f^2)+\frac{\partial }{\partial y}(h_m{\overline{\rho }}_f{\overline{u}}_f{{\overline{v}}}_f). \end{aligned}$$
    (118)

    The right-hand side terms turn into

    $$\begin{aligned} \int ^{h_m}_0\hbox{RHS}\,\hbox{d}z &= -{\overline{\rho }}_f\frac{\partial }{\partial x}\left( \frac{1}{2}h_m^2g_z\right) + \frac{\partial }{\partial x}(h_m{\overline{\phi }}_f{\overline{\tau }}_{f(xx)}) +\frac{\partial }{\partial y}(h_m{\overline{\phi }}_f{\overline{\tau }}_{f(xy)}) \nonumber \\ &-\left[ -\phi _fp\frac{\partial z}{\partial x}+\phi _f\tau _{f(xx)}\frac{\partial z}{\partial x} +\phi _f\tau _{f(xy)}\frac{\partial z}{\partial y}-\phi _f\tau _{f(xz)}\right] _0^{h_m} \nonumber \\ &+{\overline{\rho }}_fg_x\sin \zeta - C_dh_m{\overline{\phi }}_s{\overline{\phi }}_f({\overline{u}}_f-{\overline{u}}_s) \end{aligned}$$
    (119)

    where the first term on the right-hand side, \(-{\overline{\rho }}_f\partial (h_m^2g_z)/2\partial x\), accounts for the contribution of fluid pressure. The second and third terms represent the contribution of diffusion terms; however, they are often ignored, since they have a negligible effect on dynamics compared to the bed friction (terms in \([\cdot ]\)), see [19]. The terms within \([\cdot ]\) appear due to mathematical operation (Leibniz rule), and they can be simplified as follows

    $$\begin{aligned} \left[ -\phi _fp\frac{\partial z}{\partial x}+\phi _f\tau _{f(xx)}\frac{\partial z}{\partial x} +\phi _f\tau _{f(xy)}\frac{\partial z}{\partial y}-\phi _f\tau _{f(xz)}\right] _0^{h_m} =k_f^bu_f^b \end{aligned}$$
    (120)

    by using boundary conditions (75) and (78).

    Similarly, vertically integrating the cross-slope component of equation (6) can yield the depth-averaged form. The left-hand side terms turn into

    $$\begin{aligned} \int ^{h_m}_0\hbox{LHS}\,\hbox{d}z =\frac{\partial }{\partial t}({h}_m{\overline{\rho }}_f{{\overline{v}}}_f)+\frac{\partial }{\partial x}(h_m {\overline{\rho }}_f{\overline{u}}_f{{\overline{v}}}_f)+\frac{\partial }{\partial y}(h_m{\overline{\rho }}_f{{\overline{v}}}_f^2). \end{aligned}$$
    (121)

    The right-hand side terms turn into

    $$\begin{aligned} \int ^{h_m}_0 \hbox{RHS}\,\hbox{d}z&=-{\overline{\rho }}_f\frac{\partial }{\partial y}\left( \frac{1}{2}h_m^2g_z\right) - k_f^bv_f^b -C_dh_m{\overline{\phi }}_s{\overline{\phi }}_f({{\overline{v}}}_f-{{\overline{v}}}_s), \end{aligned}$$
    (122)

    where the first term on the right-hand side represents the contribution of fluid pressure; the second term stems from the bed friction; and the last term stands for viscous drag between the fluid and the solid phases.

1.4 Depth-averaged dimensional model equations

Combining relation (102) with (103) and using (104)–(106) can yield the downslope component of upper-layer granular momentum equations. Similarly, combining relation (107) with (108) can yield the cross-slope component. They are given by

$$\begin{aligned}&\frac{\partial }{\partial t}( h_g{\widetilde{\rho }}_s\phi _g)+\frac{\partial }{\partial x}( h_g{\widetilde{\rho }}_s\phi _g{\overline{u}}_g)+\frac{\partial }{\partial y}( h_g{\widetilde{\rho }}_s\phi _g{{\overline{v}}}_g) =J, \end{aligned}$$
(123)
$$\begin{aligned}&\frac{\partial }{\partial t}(h_g{\widetilde{\rho }}_s\phi _g{\overline{u}}_g)+\frac{\partial }{\partial x}(h_g {\widetilde{\rho }}_s\phi _g{\overline{u}}_g^2)+\frac{\partial }{\partial y}(h_g{\widetilde{\rho }}_s\phi _g{\overline{u}}_g{{\overline{v}}}_g)+\frac{\partial }{\partial x}\left( \frac{1}{2}K_x^gh_g^2{\widetilde{\rho }}_s\phi _gg_z\right) \nonumber \\ &\quad =Ju_g^i -{\widetilde{\rho }}_sh_g\phi _gg_z\frac{\partial h_m}{\partial x} + s_{x(g)}, \end{aligned}$$
(124)
$$\begin{aligned}&\frac{\partial }{\partial t}(h_g{\widetilde{\rho }}_s\phi _g{{\overline{v}}}_g)+\frac{\partial }{\partial x}(h_g {\widetilde{\rho }}_s\phi _g{\overline{u}}_g{{\overline{v}}}_g)+\frac{\partial }{\partial y}(h_g{\widetilde{\rho }}_s\phi _g{{\overline{v}}}_g^2)+\frac{\partial}{\partial y}\left( \frac{1}{2}K_y^gh_g^2{\widetilde{\rho }}_s\phi _gg_z\right) \nonumber \\ &\quad =Jv_g^i -{\widetilde{\rho }}_sh_g\phi _gg_z\frac{\partial h_m}{\partial y} + s_{y(g)} \end{aligned}$$
(125)

together with mass-conservation equation, where the source terms \(s_{x(g)}\) and \(s_{y(g)}\) are

$$\begin{aligned}&s_{x(g)} = {\widetilde{\rho }}_sh_g\phi _gg_x + k_s^i\mu _s^i\phi _g(u^i_s-u^i_g) +\frac{u^i_s-u^i_g}{\mid {{\varvec{u}}}^ i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i {\widetilde{\rho }}_sh_g\phi _g\left( g_z+\kappa {\overline{u}}_g^2\right) , \end{aligned}$$
(126)
$$\begin{aligned}&s_{y(g)} = k_s^i\mu _s^i\phi _g(v^i_s-v^i_g) +\frac{v^i_s-v^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i {\widetilde{\rho }}_sh_g\phi _g\left( g_z+\kappa {\overline{u}}_g^2\right) . \end{aligned}$$
(127)

Similarly, combining relation (109) with (110) and using relations (113)–(115) can yield the downslope component of the lower-layer solid momentum equations. Combining relation (116) with (117) can yield the cross-slope component of the lower-layer solid momentum equations. They are given by

$$\begin{aligned}&\frac{\partial }{\partial t}( h_m{\overline{\rho }}_s)+\frac{\partial }{\partial x}( h_m{\overline{\rho }}_s{\overline{u}}_s)+\frac{\partial }{\partial y}( h_m{\overline{\rho }}_s{{\overline{v}}}_s) =-J, \end{aligned}$$
(128)
$$\begin{aligned}&\frac{\partial }{\partial t}(h_m{\overline{\rho }}_s{\overline{u}}_s)+\frac{\partial }{\partial x}(h_m{\overline{\rho }}_s{\overline{u}}_s^2)+\frac{\partial }{\partial y}(h_m{\overline{\rho }}_s{\overline{u}}_s{{\overline{v}}}_s) \nonumber \\ &\qquad \qquad +\frac{\partial }{\partial x}\left( \frac{1}{2}h_mh_g{\overline{\rho }}_sg_z(K_y^s+K_y^g)\right) +\frac{\partial }{\partial x}\left( \frac{1}{6}({\widetilde{\rho }}_s-{\widetilde{\rho }}_f)g_zh_m^2{\overline{\phi }}_s (2K_y^s+K_y^g)\right) \nonumber \\ &\qquad \qquad \qquad \qquad \qquad =-Ju_s^i-\frac{\partial }{\partial x}\left( \frac{1}{2}{\widetilde{\rho }}_f{\overline{\phi }}_sh_m^2g_z \right) +\frac{1}{2}{\widetilde{\rho }}_fh_m^2g_z\frac{\partial {\overline{\phi }}_s}{\partial x} +{\widetilde{\rho }}_sh_g\phi _gg_z\frac{\partial h_m}{\partial x} + s_{x(s)}, \end{aligned}$$
(129)
$$\begin{aligned}&\frac{\partial }{\partial t}(h_m{\overline{\rho }}_s{{\overline{v}}}_s)+\frac{\partial }{\partial x}(h_m {\overline{\rho }}_s{\overline{u}}_s{{\overline{v}}}_s)+\frac{\partial }{\partial y}(h_m{\overline{\rho }}_s{{\overline{v}}}_s^2) \nonumber \\ &\qquad \qquad +\frac{\partial }{\partial y}\left( \frac{1}{2}h_mh_g{\overline{\rho }}_sg_z(K_y^s+K_y^g)\right) +\frac{\partial }{\partial y}\left( \frac{1}{6}({\widetilde{\rho }}_s-{\widetilde{\rho }}_f)g_zh_m^2{\overline{\phi }}_s (2K_y^s+K_y^g)\right) \nonumber \\ &\qquad \qquad \qquad \qquad \qquad =-Jv_s^i-\frac{\partial }{\partial y}\left( \frac{1}{2}{\widetilde{\rho }}_f{\overline{\phi }}_sh_m^2g_z \right) +\frac{1}{2}{\widetilde{\rho }}_fh_m^2g_z\frac{\partial {\overline{\phi }}_s}{\partial y} +{\widetilde{\rho }}_sh_g\phi _gg_z\frac{\partial h_m}{\partial y} + s_{y(s)}, \end{aligned}$$
(130)

together with mass equation, where the source terms \(s_{x(s)}\) and \(s_{y(s)}\) are given by

$$\begin{aligned} s_{x(s)} &={h}_m{\overline{\rho }}_sg_x +C_d{h}_m{\overline{\phi }}_s{\overline{\phi }}_f({\overline{u}}_f-{\overline{u}}_s) -k_s^b\mu _s^b{\phi _{s}}^b\,u_s^b\nonumber \\ &-\frac{u^b_s}{\mid {{\varvec{u}}}^b_s\mid }\mu _s^b\left[ g_z\left( {h}_m{\overline{\phi }}_s({\widetilde{\rho }}_s -{\widetilde{\rho }}_f) +{\widetilde{\rho }}_s h_g\phi _g\right) +\kappa \left( {h}_m{\overline{\phi }}_s({\widetilde{\rho }}_s{\overline{u}}_s^2 -{\widetilde{\rho }}_f{\overline{u}}_f^2) + {\widetilde{\rho }}_sh_g\phi _g{\overline{u}}_g^2\right) \right] \nonumber \\ &-\frac{u^i_s-u^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^i h_g\phi _g{\widetilde{\rho }}_s\left( g_z+ \kappa {\overline{u}}_g^2\right) -k_s^i\mu _s^i\phi _g(u_s^i-u_g^i) , \end{aligned}$$
(131)
$$\begin{aligned} s_{y(s)} &=C_d{h}_m{\overline{\phi }}_s{\overline{\phi }}_f({{\overline{v}}}_f-{{\overline{v}}}_s) -k_s^b\mu _s^b{\phi _{s}}^b\,v_s^b \nonumber \\ &-\frac{v^b_s}{\mid {{\varvec{u}}}^b_s\mid }\mu _s^b\left[ g_z\left( {h}_m{\overline{\phi }}_s({\widetilde{\rho }}_s -{\widetilde{\rho }}_f) +{\widetilde{\rho }}_s h_g\phi _g\right) +\kappa \left( {h}_m{\overline{\phi }}_s({\widetilde{\rho }}_s{\overline{u}}_s^2 -{\widetilde{\rho }}_f{\overline{u}}_f^2) + {\widetilde{\rho }}_sh_g\phi _g{\overline{u}}_g^2\right) \right] \nonumber \\ &-\frac{v^i_s-v^i_g}{\mid {{\varvec{u}}}^i_s-{{\varvec{u}}}^i_g\mid }\mu _s^ih_g\phi _g{\widetilde{\rho }}_s\left( g_z+ \kappa {\overline{u}}_g^2\right) -k_s^i\mu _s^i\phi _g(v_s^i-v_g^i). \end{aligned}$$
(132)

Combining relation (118) with (119) and using relation (120) leads to the downslope component of the fluid momentum equations. Similarly, combining relation (121) with (122) can yield the cross-slope component. The mass-conservation and momentum-conservation equations are given by

$$\begin{aligned}&\frac{\partial }{\partial t}( h_m{\overline{\rho }}_f)+\frac{\partial }{\partial x}( h_m{\overline{\rho }}_f{\overline{u}}_f)+\frac{\partial }{\partial y}( h_m{\overline{\rho }}_f{{\overline{v}}}_f) =0, \end{aligned}$$
(133)
$$\begin{aligned}&\frac{\partial }{\partial t}(h_m{\overline{\rho }}_f{\overline{u}}_f)+\frac{\partial }{\partial x}(h_m{\overline{\rho }}_f{\overline{u}}_f^2)+\frac{\partial }{\partial y}(h_m{\overline{\rho }}_f{\overline{u}}_f{{\overline{v}}}_f) \nonumber \\ &\quad +\frac{\partial }{\partial x}\left( \frac{1}{2}{\overline{\rho }}_fg_zh_m^2\right) =\frac{1}{2}h_m^2g_z\frac{\partial {\overline{\rho }}_f}{\partial x} + s_{x(f)}, \end{aligned}$$
(134)
$$\begin{aligned}&\frac{\partial }{\partial t}(h_m{\overline{\rho }}_f{\overline{u}}_f)+\frac{\partial }{\partial x}(h_m{\overline{\rho }}_f{\overline{u}}_f{{\overline{v}}}_f)+\frac{\partial }{\partial y}(h_m{\overline{\rho }}_f{{\overline{v}}}_f^2) \nonumber \\ &\quad + \frac{\partial }{\partial y}\left( \frac{1}{2}{\overline{\rho }}_fg_zh_m^2\right) =\frac{1}{2}h_m^2g_z\frac{\partial {\overline{\rho }}_f}{\partial y} + s_{y(f)}. \end{aligned}$$
(135)

Appendix 3: Fluxes and non-conservative products

To derive the fluxes, we need to explicitly express the mass-exchange rate. To this end, we eliminate the temporal derivatives from (56) by using mass equations (22), (29) and (37), so that we obtain

$$\begin{aligned} J&= -c_m\left[ h_m\left( \frac{\partial u_g}{\partial x}+\frac{\partial v_g}{\partial y}\right) -\frac{\partial }{\partial x}(h_m{\phi _{s}}u_s+h_m\phi _fu_f)-\frac{\partial }{\partial y} (h_m{\phi _{s}}v_s+h_m\phi _fv_f)\right] \nonumber \\ &- \frac{c(h_m+h_g)\phi _g}{h_g(1-\phi _g)}(\phi _m-{\phi _{s}}), \end{aligned}$$
(136)

where \(c_m=\phi _g/(1-\phi _g)\). Furthermore, the vector system (57) is a compact form of the model equations. For this system, the fluxes can be determined directly from the left-hand sides of Eqs. (22), (23), (29), (30), (37) and (38), which yields

$$\begin{aligned}&\displaystyle {\varvec{F}}= \left( m_g^x-c_m(m^x_{s}+m_f^x), \,\,\,\, m^x_{s}+c_m(m^x_{s}+m_f^x), \,\,\,\, m_f^x, \,\,\,\, (m_g^x)^2/ {\hat{h}}_g+\frac{1}{2}\beta ^x_g {\hat{h}}_g^2, \right. \nonumber \\ &\quad \quad \, \left. (m^x_{s})^2/ {\hat{h}}_s +\frac{1}{2}\beta ^{x_1}_{s} {\hat{h}}_g({\hat{h}}_s+{\hat{h}}_f)+\frac{1}{2}\beta ^{x_2}_{s}{\hat{h}}_s({\hat{h}}_s+{\hat{h}}_f), \,\,\,\, (m_f^x)^2/{\hat{h}}_f+\frac{1}{2}\beta ^x_f{\hat{h}}_f({\hat{h}}_s+{\hat{h}}_f) \right) ^T. \end{aligned}$$
(137)

Eq. (137) indicates a flux Jacobian

$$\begin{aligned} \frac{\partial {\varvec{F}}}{\partial \varvec{\Omega }} = \left[ {\begin{array}{cccccc} 0 &{} 0 &{} 0 &{} 1 &{} -c_m &{} -c_m\\ 0 &{} 0 &{} 0 &{} 0 &{} 1+c_m &{} c_m\\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 1\\ \displaystyle -\frac{(m_g^x)^2}{{\hat{h}}_g^2}+\beta _g^x{\hat{h}}_g &{} 0 &{} 0 &{} \displaystyle 2\frac{m_g^x}{{\hat{h}}_g} &{} 0 &{} 0 \\ \displaystyle \frac{\beta ^{x_1}_s({\hat{h}}_s+{\hat{h}}_f)}{2} &{} B^*&{} C^*&{} 0 &{} \displaystyle 2\frac{m_s^x}{{\hat{h}}_s} &{} 0\\ 0 &{} \displaystyle \frac{\beta _f^x{\hat{h}}_f}{2} &{} \displaystyle -\frac{(m_f^x)^2}{{\hat{h}}_f^2}+\frac{\beta _f^x(2{\hat{h}}_f+{\hat{h}}_s)}{2}&{} 0 &{} 0 &{} \displaystyle 2\frac{m_f^x}{{\hat{h}}_f} \end{array} } \right] , \, \end{aligned}$$
(138)

where \(B^*= {-(m_s^x)^2}/{{\hat{h}}_s^2}+{\beta _s^{x_1}{\hat{h}}_g/2+\beta _s^{x_2}(2{\hat{h}}_s+{\hat{h}}_f)}/2\) and \(C^*= \beta _s^{x_1}{\hat{h}}_g/2 + \beta _s^{x_2}{\hat{h}}_s/2\).

Inspect the non-conservative terms of Eqs. (22), (23), (29), (30), (37) and (38) can yield the form of \({\varvec{W}}_x\). It should be noted that in the process of formulating non-conservative terms of Eq.(30), we combine the terms \(-\partial (\epsilon \gamma {\overline{\phi }}_sh_m^2\cos \zeta /2)/\partial x\) and \((\epsilon \gamma h_m^2\cos \zeta /2)\partial {\overline{\phi }}_s/\partial x\) on the right-hand side of (30) as \(-\epsilon \gamma {\overline{\phi }}_s\partial (h_m^2\cos \zeta /2)/\partial x\), which can be further approximated as \(-(\epsilon \gamma {\overline{\phi }}_sh_m\cos \zeta )\partial h_m/\partial x\) due to \(\partial \zeta /\partial x \approx 0\) justified by \(\lambda<< 1\). With the above instructions, the matrix-valued function \({\varvec{W}}_x\) can be directly formulated as

$$\begin{aligned} {{\varvec{W}}}_x= \left[ {\begin{array}{cccccc} \displaystyle -\frac{c_mh_mu_g}{{\hat{h}}_g} &{} 0 &{} 0 &{} \displaystyle \frac{c_mh_m}{{\hat{h}}_g} &{} 0 &{} 0 \\ \displaystyle \frac{c_mh_mu_g}{{\hat{h}}_g} &{} 0 &{} 0 &{} \displaystyle -\frac{c_mh_m}{{\hat{h}}_g} &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 \\ \displaystyle -\frac{c_mh_mu_su_g}{{\hat{h}}_g} &{} \epsilon {\hat{h}}_g\cos \zeta &{} \epsilon {\hat{h}}_g\cos \zeta &{} \displaystyle \frac{c_m \displaystyle h_mu_s}{{\hat{h}}_g} &{} \displaystyle -c_mu_s &{} \displaystyle -c_mu_s\\ \displaystyle \frac{c_mh_mu_su_g}{{\hat{h}}_g}&{} -\epsilon ({\hat{h}}_g-\gamma {\hat{h}}_s)\cos \zeta &{} -\epsilon ({\hat{h}}_g-\gamma {\hat{h}}_s)\cos \zeta &{} \displaystyle -\frac{c_mh_mu_s}{{\hat{h}}_g} &{} c_mu_s &{} \displaystyle c_mu_s\\ \displaystyle 0 &{} \displaystyle \frac{\epsilon {\hat{h}}_f\cos \zeta }{2} &{} \displaystyle -\frac{\epsilon {\hat{h}}_s\cos \zeta }{2} &{} 0 &{} 0 &{} 0 \\ \end{array} } \right] , \, \end{aligned}$$
(139)

where \(h_m={\hat{h}}_s+{\hat{h}}_f\), and \(u_g=m_g^x/{\hat{h}}_g\) and \(u_s=m_s^x/{\hat{h}}_s\).

To formulate the source terms we extract the non-derivative term \({\mathcal {G}}=-c\phi _g(h_m+h_g)(\phi _m-{\phi _{s}})/(h_g(1-\phi _g))\) from mass-exchange rate (136), so that the vector of source terms can be written as

$$\begin{aligned} {\varvec{S}} &=\left( {\mathcal {G}},\,\,\, -{\mathcal {G}},\,\,\, 0, \,\,\, {\mathcal {G}}m_s^x/{\hat{h}}_s + s_{x(g)}, \,\,\, -{\mathcal {G}}m_s^x/{\hat{h}}_s + s_{x(s)}, \,\,\, s_{x(f)} \right) ^T. \end{aligned}$$
(140)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Meng, X., Wang, Y., Wang, C. et al. Modeling of unsaturated granular flows by a two-layer approach. Acta Geotech. 12, 677–701 (2017). https://doi.org/10.1007/s11440-016-0509-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11440-016-0509-x

Keywords

Navigation