We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Sediment Transport in Alluvial Systems | SpringerLink

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Skip to main content

Sediment Transport in Alluvial Systems

  • Chapter
  • First Online:
Physics of Lakes

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

  • 925 Accesses

Abstract

Sediment transport arises in alluvial lake-river systems in two different forms: (i) as bed load, comprising the moving detritus of the river bed and of the shallow, often only near-shore regions, and (ii) the suspended sediment load of the finer fractions. In river hydraulics the latter are often neglected; so, the bed load transport is treated without back-coupling with the wash-load. This is justified on decadal time scales. In the deeper parts of lakes wind-induced shearing in the benthic boundary layer hardly mobilizes the bed material, which stays immobile for most time and may be set in motion only interruptedly. However, the particle laden fluid transports the suspended material, which is advected and may on longer time scales settle in deposition-prone regions. In general, the deposition to and erosion from the basal surface occur concurrently. This environmental interplay is studied in this article. The slurry—a mixture of the bearer fluid and particles of various sizes—is treated as a mixture of class I, in which mass, momentum and energy balances for the mixture as a whole are formulated to describe the geophysical fluid mechanical setting, whilst the suspended solid particles move through the bearer medium by diffusion. The governing equations of this problem are formulated, at first for a compressible, better non-density preserving, mixture. They thus embrace barotropic and baroclinic processes. These equations, generally known as Navier–Stokes–Fourier–Fick (NSFF) fluids, are subjected to turbulent filter operations and complemented by zeroth and first order closure schemes . Moreover, simplified versions, e.g. the (generalized) Boussinesq, shallow water and hydrostatic pressure assumptions are systematically derived and the corresponding equations presented in both conservative and non-conservative forms. Beyond the usual constitutive postulates of NSFF-fluids and turbulent closure schemes the non-buoyant suspended particles give rise to settling velocities; these depend on the particle size, expressed by a nominal particle diameter. A review of the recent hydraulic literature of terminal settling velocities is given. It shows that the settling velocity depends on the particle diameter and on the particle Reynolds number. A separate section is devoted to the kinematic and dynamic boundary conditions on material and non-material singular surfaces as preparation for the mathematical-physical description of the sediment transport model, which follows from an analysis of jump transition conditions at the bed. The simplest description of bed load transport does not use the concept of the motion of a thin layer of sediments. It treats it as a singular surface, which is equipped with surface grains of various grain size diameters. Such a simplified theoretical level is also used in this chapter; it implies that solid mass exchange, as erosion and deposition of different particle size fractions, is the only physical quantity relevant in the description of the sediment transport. It entails formulation of surface mass balances of an infinitely thin detritus layer for the sediment and surface momentum balance of the mixture. The deposition rate of the various grain fractions, expressed as grain classes, follows from a parameterization of the free fall velocity of isolated particles in still water, but is in general coupled with the local flow and then follows from the solution of the hydrodynamic equations and the processes at the basal surface. The erosion rate is governed by two statements, (a) a fracture criterion determining the threshold value of a stress tensor invariant at the basal surface, which separates existence and absence regimes of erosion, and (b) determination of the amount of erosion beyond the threshold value of the mentioned stress invariant.

This chapter is a slightly extended version of an article with the title ‘A global view of sediment transport in alluvial systems’ and co-authored by K.H. and Prof. Dr. Ioana Luca, Department of Mathematics II, University Politehnica of Bucharest, and published within a report on Sediment Transport of the Institut für Wasserbau at the Technische Universität, München in 2013 [18]. We thank Professor Dr. sc. tech. Peter Rutschmann for the permission of copyright in this extended form.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    It is assumed that no fragmentation of particles occurs.

  2. 2.

    Often these fields may even be dropped and simply assumed to be frozen to the fluid particles.

  3. 3.

    The focus is not on strong internal baroclinic motion but rather on the reproduction of the current near the basal surface (e.g. the benthic boundary layer).

  4. 4.

    In the theory of interfaces such sheets are called diffuse interfaces.

  5. 5.

    This is motivated by sieve experiments: one has a whole column of sieves, numbered \(0,\ldots , \alpha ,\ldots , N-1\), with the largest mesh size on top and the smallest at the bottom; class \(\alpha \) (\(\alpha =1,\ldots ,N\)) consists of those particles which are collected by sieve \(\alpha -1\). It is tacitly understood that the sieve with number ‘\(0\)’ is impermeable for all particles of sizes larger than a chosen minimum (say for clay and silt fractions which cannot pass very small holes simply because of cohesion coalescence).

  6. 6.

    Consider the term \(p\,\mathrm {div\,}{\varvec{v}}\) on the right-hand side of (32.3)\(_1\). With the aid of (32.1) this takes the form

    $$\begin{aligned} -p\,\mathrm {div\,}{\varvec{v}}= \frac{p}{\rho }\frac{{{\text {d}}}\rho }{{{\text {d}}}{t}} = -\rho \frac{{\text {d}}}{{{\text {d}}}{t}}\left( \frac{p}{\rho }\right) + \frac{{{\text {d}}}p}{{{\text {d}}}{t}}. \end{aligned}$$

    Therefore, the balance of mixture energy may also be written as

    $$\begin{aligned} \rho \frac{{\text {d}}}{{{\text {d}}}{t}}\left( \epsilon + \frac{p}{\rho }\right) - {\frac{{{\text {d}}}p}{{{\text {d}}}{t}}} = -\, \mathrm {div\,}{\varvec{q}}+ \mathrm {tr}\,({{\varvec{\sigma }}}_{\tiny {E}}{\varvec{D}}), \end{aligned}$$

    suggesting the definition of the mixture enthalpy (32.4). In almost density preserving materials the term \(p\,\mathrm {div\,}{\varvec{v}}\) in (32.3)\(_1\) and the term \({\text {d}}p/{\text {d}}t\) in (32.3)\(_2\) are generally ignored, which implies \({\text {d}}\epsilon /{{\text {d}}}{t} \approx {\text {d}}h/{{\text {d}}}{t}\), which is the reason why one can often see in the literature both formulations using \(\epsilon \) or \(h\).

  7. 7.

    For a non-buoyant particle \(\alpha \) falling in still water we have \({\varvec{w}}_{\alpha }^{s}\equiv -({{\varvec{v}}}_{\alpha }^{s}-{{\varvec{v}}})= w_{\alpha }^s{\varvec{e}}_z\); here \({{\varvec{v}}}_{\alpha }^{s}\) is the velocity of the solid particle, and \({\varvec{v}}\approx \varvec{ {0}}\) is the velocity of the surrounding fluid at rest. When the grain stops to decelerate it has attained the so-called terminal settling velocity or free fall velocity, see Fig. 32.6.

  8. 8.

    (a) Using (32.1) yields for the left-hand side of (32.2)

    $$\begin{aligned} \rho \frac{{\text {d}}{\varvec{v}}}{{{\text {d}}}{t}}= \frac{{{\text {d}}}\rho {\varvec{v}}}{{{\text {d}}}{t}}-\frac{{{\text {d}}}\rho }{{{\text {d}}}{t}}\,{\varvec{v}}\mathop {=}\limits ^{(32.1)} \frac{{{\text {d}}}\rho {\varvec{v}}}{{{\text {d}}}{t}} + (\rho \,\mathrm {div\,}{\varvec{v}}){\varvec{v}}= \frac{\partial \rho {\varvec{v}}}{\partial t} + \mathrm {div\,}(\rho {\varvec{v}}\otimes {\varvec{v}}), \end{aligned}$$

    whilst the right-hand side remains unchanged.

    (b) Using (32.1), for a scalar function \(f\) we obtain

    $$\begin{aligned} \rho \frac{{\text {d}}f}{{{\text {d}}}{t}}= \frac{{{\text {d}}}\rho f}{{{\text {d}}}{t}}-\frac{{{\text {d}}}\rho }{{{\text {d}}}{t}}\,f \mathop {=}\limits ^{(32.1)} \frac{{{\text {d}}}\rho f}{{{\text {d}}}{t}} + (\rho \,\mathrm {div\,}{\varvec{v}})f = \frac{\partial \rho f}{\partial t} + \mathrm {div\,}(\rho f{\varvec{v}}), \end{aligned}$$

    which turns (32.3) and (32.5) into (32.15) and (32.16), respectively.

  9. 9.

    For Reynolds filters one has \(\langle \,f^{{\prime }}\,\rangle =0\). Other filters may not enjoy this property. They then allow subgrid scale modelling, which e.g., is employed in so-called Large Eddy Simulation (LES).

  10. 10.

    If for the velocity the so-called Favre averaging operator is employed,

    $$\begin{aligned} \{{\varvec{v}}\} \equiv \frac{\langle \rho {\varvec{v}}\rangle }{\langle \rho \rangle }, \end{aligned}$$
    (32.19)

    then the averaged mass balance takes the form

    $$\begin{aligned} \frac{\partial \langle \rho \rangle }{\partial t} + \mathrm {div\,}(\langle \,\rho \rangle \{{\varvec{v}}\}) = 0. \end{aligned}$$
    (32.20)

    So, Favre averaging would preserve the invariance of the balance of mass under filtering. However, this would also imply consequences in the remaining balance laws. A complete derivation using Favre averaging is e.g. given in Luca et al. (2004) [27]. We prefer to stay with (32.18).

  11. 11.

    For these formulae we employ the symbol \(\langle \cdot \rangle \) of filter operation to emphasize the role of the averaged laminar quantities and averages of turbulent correlation quantities. Note also that \({\varvec{Q}}_{\epsilon }\) and \({\varvec{Q}}_{h}\) differ by their turbulent correlation terms.

  12. 12.

    We shall refer to specific pages of [19] as (I, \(\ldots \) ).

  13. 13.

    (a) Parameterization (32.50)\(_{4}\) does not account for cross dependences of the form

    $$\begin{aligned} - \sum _{\beta =1}^{N} \lambda _{\alpha \beta } \left( \chi _{\ell }^{(c_{\beta })}+\frac{\nu _{t}}{\sigma _{c_{\beta }}}\right) {\mathrm {grad}}\,\,c_{\beta },\quad \alpha = 1,\dots , N, \end{aligned}$$

    with \(\lambda _{\alpha \beta } < 1.\) Our selection in (32.50)\(_{4}\) is \(\lambda _{\alpha \beta } = \delta _{\alpha \beta }\). In principle the more general case is possible.

  14. 14.

    For a solenoidal velocity field it is often customary to incorporate the contribution of the turbulent kinetic energy \(k\) in relation (32.50)\(_1\) into the pressure term, or to ignore it.

  15. 15.

    Such a representative element in class \(\alpha \) has already been used when defining the advected mass flux \(\rho c_\alpha {{\varvec{w}}}_\alpha ^s\). To simplify the notation, we use \({\mathfrak {d}_{\alpha }}\) for the diameter of this grain particle; note that \({\mathfrak {d}_{\alpha }}\in [d_{\alpha -1}, d_\alpha )\), so that \({\mathfrak {d}_{\alpha }}\) should not be confused with \(d_{\alpha }\).

  16. 16.

    McGauhey [28], Zanke [57], Concha and Almendra [9], Turton and Clark [46], Zhang [59], Julien [23], Soulsby [44], Cheng [8], Ahrens [1], Guo [13], Jimenez and Madsen [21], Brown and Lawler [6], She et al. [39], Camenen [7].

  17. 17.

    \(N\) is the number of experimental points where values for \((w^{s}_{\alpha })^{\text {exp}}\) have been measured.

  18. 18.

    This equation holds with \({\varvec{w}}_{\alpha }^{s} = w_{\alpha }^{s}{\varvec{e}}_{z}\). If \({\varvec{w}}_{\alpha }^{s}\) is parameterized as in Eq. (32.11), then

    $$\begin{aligned} \frac{\partial }{\partial x}(\rho \,c_{\alpha }w_{\alpha }^{s}\tan \theta \cos \xi ) + \frac{\partial }{\partial y}(\rho \,c_{\alpha }w_{\alpha }^{s}\tan \theta \sin \xi ) \end{aligned}$$

    must be added to the right-hand side of Eq. (32.80). Here, \(\xi \) is the angle between the \(x\)-axis and \({\varvec{v}}_{H}\).

  19. 19.

    The symbol \([f]\) denotes an order of magnitude for the quantity \(f\) within the range of values which \(f\) may assume (in the physical dimensions in which it is expressed) in the processes under consideration.

  20. 20.

    We neglect the contribution of the turbulent kinetic energy in (32.52)\(_1\).

  21. 21.

    \(\phi ^{(c_{\alpha })}\) could consist of fragmentation and abrasion of suspended particles, which, however, are unlikely processes.

  22. 22.

    The proof of this formula requires the definition (32.112)\(_1\) and expressions of the form

    $$\begin{aligned} \frac{\partial }{\partial \xi ^{\mathfrak b}}\left( {\varvec{\tau }}_{{\mathfrak a}}\cdot {\varvec{\tau }}_{{\mathfrak c}}\right) = \frac{\partial {\varvec{\tau }}_{{\mathfrak a}}}{\partial \xi ^{{\mathfrak b}}}\cdot {\varvec{\tau }}_{{\mathfrak c}} + {\varvec{\tau }}_{{\mathfrak a}}\cdot \frac{\partial {\varvec{\tau }}_{{\mathfrak c}}}{\partial \xi ^{{\mathfrak b}}} = \frac{\partial g_{{\mathfrak a\mathfrak c}}}{\partial \xi ^{{\mathfrak b}}}. \end{aligned}$$
    (32.113)
  23. 23.

    It states that, if \({\varvec{\varPhi }}^{f_{{\mathcal S}}}\) is a differentiable surface tensor field, then, see Fig. 32.9,

    $$\begin{aligned} \int _{\mathcal S}\text {Div}({\varvec{\varPhi }}^{f_{{\mathcal S}}}){\text {d}}a = \int _{{\mathcal C}}{\varvec{\varPhi }}^{f_{{\mathcal S}}}\cdot {\varvec{h}}{\text {d}}s, \end{aligned}$$

    in which \({\varvec{h}}\) is a unit vector normal to \({\mathcal C}\) and tangential to \({\mathcal S}\), see e.g., [41] p. 669–670.

  24. 24.

    If barotropic surface waves are ignored, i.e., the rigid lid approximation is imposed, then (32.144) is replaced by \(z = 0\), where the origin of the coordinate system is at the undeformed free surface and the \(x\) and \(y\) axes are horizontal.

  25. 25.

    The right-hand side of (32.147)\(_2\) should involve the difference \(({\varvec{v}}_{H}^{\text {wind}} - {\varvec{v}}_{H}^{\text {water}})_{{\mathcal S}}\), but the water velocity is very much smaller than the wind velocity, which justifies the approximation.

  26. 26.

    If \({{\mathfrak d}_{\alpha }}\) and \({{\mathfrak d}_{\alpha _{\text {thres}}}}\) are the nominal grain diameters of the grain size classes \(\alpha \) and \(\alpha _{\text {thres}}\), respectively, then all grains with \({{\mathfrak d}_{\alpha }}<{{\mathfrak d}_{\alpha _{\text {thres}}}}\) are mobilized, whilst those with \({{\mathfrak d}_{\alpha }} > {{\mathfrak d}_{\alpha _{\text {thres}}}}\) are still at rest.

  27. 27.

    Experiments on erosion inception and inferences thereof are performed and deduced for sediments with a single particle diameter. This is why ensuing formulae are written without identifying a specific particle class, even though later, they are applied to such classes.

  28. 28.

    More generally, a criterion marking the onset of erosion is an equation of the form

    $$\begin{aligned} f(I_{\varvec{\sigma }},\, I\!I_{{\varvec{\sigma _{\!\tiny {E}}}}},\, I\!I\!I_{{\varvec{\sigma _{\!\tiny {E}}}}}) = 0 \quad (*) \end{aligned}$$

    between the first stress invariant and the second and third stress deviator invariants at the basal surface. A dependence on \(I_{\varvec{\sigma }}\) describes a possible influence of the (mean) pressure; that on \(I\!I_{ {\varvec{\sigma _{\!\tiny {E}}}}} \) accounts for the significance of shearing, but the role of \(I\!I\!I_{{\varvec{\sigma _{\!\tiny {E}}}}}\) is presently not clear. In the form \((*)\) the erosion inception is very much reminiscent of the onset criterion of yield in the theories of plasticity.

  29. 29.

    To prove this, we write

    $$\begin{aligned} {\varvec{w}} = \alpha {\varvec{\tau }}_1 + \beta {\varvec{\tau }}_2 + {\mathcal U}_{b} {\varvec{n}}_{b} = \frac{\partial b}{\partial t}{\varvec{e}}_{3}. \end{aligned}$$

    If the above expressions for \({\varvec{\tau }}_1\) and \({\varvec{\tau }}_2\) are substituted this yields

    $$\begin{aligned} {\varvec{w}} = \left( \alpha - c{\mathcal U}_{b} \frac{\partial b}{\partial x}\right) {\varvec{e}}_1 + \left( \beta - c{\mathcal U}_{b} \frac{\partial b}{\partial y}\right) {\varvec{e}}_2 + \left( \alpha \frac{\partial b}{\partial x} + \beta \frac{\partial b}{\partial y} + c{\mathcal U}_{b}\right) {\varvec{e}}_3 = \frac{\partial b}{\partial t}{\varvec{e}}_3, \end{aligned}$$

    implying

    $$\begin{aligned} \alpha = c{\mathcal U}_{b}\frac{\partial b}{\partial x},\quad \beta = c{\mathcal U}_{b}\frac{\partial b}{\partial y}. \end{aligned}$$
  30. 30.

    Developing a model with the consideration of (32.183) requires further assumptions on \(S\), \({\varvec{(}}\!\!-p\varvec{I}+{\varvec{R}})^+{\varvec{n}}_b\cdot {\varvec{n}}_b\), \((-p\varvec{I}+{\varvec{R}})^-{\varvec{n}}_b\cdot {\varvec{n}}_b\). We prefer to give an erosion/deposition law and so omit (32.183).

  31. 31.

    If we assume \(\mu _{\alpha } = \rho _s\nu _{\text {mean}}^{\alpha }h\), then the mean volume fractions \(\nu _{\text {mean}}^{\alpha }\) are known once the height \(h\) is known: \(\nu _{\text {mean}}^{\alpha }=\mu _{\alpha }/(\rho _s h)\); or, equivalently, if \(\nu _{\text {mean}}\) is known, see (32.205): \(\nu _{\text {mean}}^{\alpha }=(\mu _\alpha /\mu ) \nu _{\text {mean}}\). For the detritus layer the mean volume fractions \(\nu _{\text {mean}}^{\alpha }\) are practically better quantities than the surface densities \(\mu _\alpha \).

  32. 32.

    The determination of the velocity of a particle in a moving and perhaps accelerating fluid field relative to the velocity of the fluid at the same position before the latter was inserted in the fluid, is a complex topic of fluid dynamics which does not, in general, agree with the free fall velocity.

  33. 33.

    Such a filter need not to fulfil the condition \(\langle \langle \cdot \rangle \rangle = \langle \cdot \rangle \), where \(\langle \cdot \rangle \) is the filter operation.

References

  1. Ahrens, J.P. A fall velocity equation. J. Waterways, Port, Coastal and, Ocean Engineering 126(2), 99–102 (2000)

    Google Scholar 

  2. Alts, T. and Hutter, K. Continuum description of the dynamics and thermodynamics of phase boundaries between ice and water. Part I: Surface balance laws and their interpretation in terms of three dimensional balance laws averaged over the phase change boundary layer. J. Non-Equilibrium Thermodyn. 13, 221–257 (1988)

    Google Scholar 

  3. Alts, T. and Hutter, K. Continuum description of the dynamics and thermodynamics of phase boundaries between ice and water. Part II: Thermodynamics. J. Non-Equilibrium Thermodyn. 13, 259–280 (1988)

    Google Scholar 

  4. Alts, T. and Hutter, K. Continuum description of the dynamics and thermodynamics of phase boundaries between ice and water. Part III: Thermostatics and its consequences. J. Non-Equilibrium Thermodyn. 13, 301–329 (1988)

    Google Scholar 

  5. Alts, T. and Hutter, K. Continuum description of the dynamics and thermodynamics of phase boundaries between ice and water. Part IV: On thermostatic stability and well posedness. J. Non-Equilibrium Thermodyn. 14, 1–22 (1989)

    Google Scholar 

  6. Brown, P.P. and Lawler, D.F. Sphere drag and settling velocity revisited. J. Environmental Engineering, 129(3), 222–231 (2003)

    Google Scholar 

  7. Caamenen, B. Simple and general formula for the settling velocity of particles. J. Hydraulic Engineering, 133(2), 229–233 (2007)

    Google Scholar 

  8. Cheng, N.S. Simplified settling velocity formula for sediment particle. J. Hydraul. Engineering, 123(2), 149–152 (1997)

    Google Scholar 

  9. Concha, F. and Almendra, E.R. Settling velocities of particle systems: I Settling velocities of individual spherical partiles. Int. J. Mineral Processing. 5(4), 349–367 (1979)

    Google Scholar 

  10. Eglund, F. and Hansen, E. A monograph on sediment transport in alluvial streams, third Ed. Technical Press, Copenhagen (1972)

    Google Scholar 

  11. Einstein, H.A. The bedload function for sediment transportation in open channel flow. Technical Bulletin Nr 1026 U.S. Dept. Agriculture, Washington DC. (1950)

    Google Scholar 

  12. Germano, M., Piolelli, U., Moin, P. and Cabiot, W.H. A dynamic subgrid-scale eddy viscosity model. Phys. Fluids A 3(7), 1760–1765 (1991)

    Google Scholar 

  13. Guo, J. Logarithmic matching and its application in computational hydraulics and sediment transport. J. Hydraul. Research, 40(5), 555–565 (2002)

    Google Scholar 

  14. Hanjalic, K. and Launder, B.E. A Reynolds stress model of turbulence in its application to thin shear flows. J. Fluid Mech. 52, 609–638 (1972)

    Google Scholar 

  15. Hutter, K. The physics of ice-water phase change surfaces In: CISM Lectures: Modelling macroscopic phenomena at liquid boundaries, Murdoch, I. and Kosinski, W., eds., Springer Verlag New York, Vienna (1992)

    Google Scholar 

  16. Hutter, K. Fluid und Thermodynamik, eine Einführung. Springer Verlag Berlin etc. (2003)

    Google Scholar 

  17. Hutter, K. and Jöhnk, K. Continuum Methods of Physical Modeling. Springer, Berlin, 635p (2004)

    Google Scholar 

  18. Hutter, K. and Luca, I.: A global view of sediment transport in alluvial systems. In: Contributions on Sediment Transport. Bericht des Lehrstuhls und der Versuchsanstalt für Wasserbau und Wasserwirtschaft der TU München Nr. 127, Herausgeber: Prof. P. Rutschmann (2013)

    Google Scholar 

  19. Hutter, K., Wang, Y. and Chubarenko, I. Physics of Lakes - Foundation of the Mathematical and Physical Background Vol. I. Springer, Berlin, 434p (2011)

    Google Scholar 

  20. Hutter, K., Wang, Y. and Chubarenko, I. Physics of Lakes - Lakes as Oscillators Vol. II, Springer, Berlin, 646p (2011)

    Google Scholar 

  21. Jimenez, J.A. and Madsen, O.S. A simple formula to estimate settling velocity of natural sediments. J. Waterways, Port, Coastal and, Ocean Engineering 129(2), 70–78 (2003)

    Google Scholar 

  22. Jones, W.P and Launder, B.E. The prediction of laminarisation with a two-equation model of turbulence. J. Heat Mass Tranf. 15, 301–314 (1972)

    Google Scholar 

  23. Julien, Y.P. Erosion and Sedimentation. Cambridge University Press, Cambridge (1995)

    Google Scholar 

  24. Kraft, S. Wang, Y. and Oberlack, M. Large eddy simulation of sediment deformation in a turbulent flow by means of level set method. J. Hydr. Eng. ASCE, November 2011, 1–12 (2011)

    Google Scholar 

  25. Launder, B.E and Spalding, D.B. The numerical computation of turbulent flow. Comp. Meth. Appl. Mech. Eng. 3, 269–288 (1974)

    Google Scholar 

  26. Lilly, D. K. The representation of small-scale turbulence in numerical simulation experiments. Proc. IBM Scientific Computing Symp. Environ. Sci., 195–210 (1967)

    Google Scholar 

  27. Luca, I., Fang, Ch. and Hutter, K. A thermodynamic model of turbulent motion in a granular material. Continuum Mech. Thermodyn. 16, 363–390 (2004)

    Google Scholar 

  28. McGauhay, P.H. Theory of Sedimentation. J. of the American Water Works Assoc. 48, 437–448 (1956)

    Google Scholar 

  29. Miles, J. W. On the stability of heterogeneous shear flows. J. Fluid Mech. 10, 496–508 (1961)

    Google Scholar 

  30. Moin, P. and Kim, J. Numerical investigation of turbulent channel flow. J. Fluid Mech., 118(1), 341–377 (1982)

    Google Scholar 

  31. Müller, I. Thermodynamics, Pitman, London (1985)

    Google Scholar 

  32. Munk, W. H. On the wind-driven ocean circulation J. Meteorology, 7, 79 (1950)

    Google Scholar 

  33. Piomelli, U. High Reynolds number calculations using the dynamic subgri-scale stress model. Phys. Fluids A 5(6), 1484 (1993)

    Google Scholar 

  34. Piomelli, U. and Zang, T. A. Large eddy simulation of transitional channel flow. Comput. Phys. Commun. 65(1–3), 224–230 (1991)

    Google Scholar 

  35. Prandtl, L. Neuere Ergebnisse der Turbulenzforschung. Zeitschr. VDI 77, 105–113 (1933)

    Google Scholar 

  36. Rodi, W. Examples of calculation methods for flow and mixing in stratified fluids. J. Geophys. Res. (C5), 92, 5305–5328 (1987)

    Google Scholar 

  37. Rodi, W. Turbulence Models and Their Application in Hydraulics, IAHR Monograph Series. A.A. Balkema, Rotterdam/Brookfield (1993)

    Google Scholar 

  38. Rotta, J.C. Turbulente Strömungen, eine Einführung in die Theorie und ihre Anwendung, Teubner, Stuttgart, 267 p (1972)

    Google Scholar 

  39. She, K., Trim, L. and Pope, D. Fall velocities of natural sediment particles: A simple mathematical presentation of the fall velocity law. J. Hydraul. Research, 43(2), 189–195 (2005)

    Google Scholar 

  40. Shields, A. Anwendung der Ähnlichkeitsmechanik und der Turbulenzforschung. Mitt. der Preussischen Versuchsanstalt für Wasser- und Schiffsbau, Heft 26, Berlin (1936)

    Google Scholar 

  41. Slattery, J. C., Sagis, L. and Oh, E.-S. Interfacial transport phenomena. 2\(^{\text{ nd }}\) Ed., Springer (2007)

    Google Scholar 

  42. Smagorinsky, J. General circulation experiments with the primitive equations I. The basic experiment. Mon. Weather Rev. 91(3), 99–165 (1963)

    Google Scholar 

  43. Song, Z., Wu, T., Xu, F. and Li, R. A simple formula for predicting settling velocity of sediment particles. Water Science and, Engineering, 1(1) 37–43 (2008) DOI:10.3882/j.issn.1674-2370.2008.01.005

    Google Scholar 

  44. Soulsby, R.L. Dynamics of Marine Sands. Thomas Telford, London (1997)

    Google Scholar 

  45. Thomas, T.Y. Plastic Flow and Fracture in Solids. Academic Press, New York, London (1961)

    Google Scholar 

  46. Turton, R. and Clark, N.N. An explicit relationship to predict spherical particle terminal velocity. Power Technol. 53(2), 1127–1129 (1987).

    Google Scholar 

  47. Van Rijn LC (1984) Sediment pick-up functions. J. Hydraul. Eng. ASCE 110(10):1494–1502

    Article  Google Scholar 

  48. Van Rijn, L.C. Sediment transport, Part II: Suspended load transport. J. Hydraul. Eng. ASCE 110(11), 1613–1641 (1984).

    Google Scholar 

  49. Van Rijn, L.C. Unified view of sediment transport by currents and waves. I: Initiation of motion, bed roughness, and bed-load transport. J. Hydraul. Eng. ASCE 133(6), 649–667 (2007)

    Google Scholar 

  50. Vetsch, D. Numerical Simulation of Sediment Transport with Meshfree Methods. Doctoral Dissertation, Laboratory of Hydraulics, Hydrology and Glaciology, Swiss Federal Institute of Technology, Zurich p. 200 (2012)

    Google Scholar 

  51. Wang, Y. Windgetriebene Strömungen in einem Rechteckbecken und im Bodensee. PhD thesis, TH Darmstadt, Germany, Shaker Verlag, Aachen, ISBN 3-8265-1381-9. p. 432 (1996).

    Google Scholar 

  52. Wilcox, D.C. Reassessment of the scale determining equation for advanced turbulence models. AIAA J. 26(11), 1299–1310 (1988)

    Google Scholar 

  53. Wilcox, D.C. Turbulence modeling for CFD. DWC Industries, Inc., La Ca\(\tilde{\text{ n }}\)ada, California, 2nd Edition (1998)

    Google Scholar 

  54. Yalin, M.S. Mechanics of sediment transport. Second Edition Pergamon Press, Oxford, (1977)

    Google Scholar 

  55. Yalin, M.S. On the determination of ripple geometry. J. Hydraul. Eng. 111(8), 1148–1155 (1985)

    Google Scholar 

  56. Yalin, M.S. and da Silva, A.M.F. Fluvial processes. IAHR International Association of Hydraulic Engineering and Research (2001)

    Google Scholar 

  57. Zanke, U. Berechnung der Sinkgeschwindigkeiten von Sedimenten. Mitt. des Franzius-Instituts für Wasserbau 46, 243 p (1977)

    Google Scholar 

  58. Zanke, U.C.E. Zum Einfluss der Turbulenz auf den Beginn der Sedimentbewegung Mitt. Institut für Wasserbau und Wasserwirtschaft. Techn. Hochschule Darmstadt p 120 (2001)

    Google Scholar 

  59. Zhang, R.J. Sediment Dynamics in Rivers. Water Resources Press (in Chinese) (1989)

    Google Scholar 

  60. Zhu, L.J. and Cheng, N.S. Settlement of settling particles. River and Harbour Engineering Department, Nanjing Hydraulic Research Institute, Nanjing (in Chinese) (1993)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kolumban Hutter .

Appendices

Appendix A: Implications from the Second Law of Thermodynamics

This appendix gives a justification for the approximation (32.44). The results which are presented can be taken from any book on thermodynamics, e.g. Hutter (2003) [16]. The basis of the considerations is the so-called Gibbs relation of a heat conducting fluid,

$$\begin{aligned} {\text {d}}\eta = \frac{1}{T}\left( {\text {d}}\epsilon - \frac{p}{\rho ^{2}}\,{\text {d}}\rho \right) , \end{aligned}$$
(32.210)

in which \(\eta \) is the entropy, \(T\) the Kelvin temperature, \(\epsilon \) the internal energy, \(p\) the pressure and \(\rho \) the fluid density; (32.210) is a consequence of the second law of thermodynamics. Solving (32.210) for \({\text {d}}\epsilon \),

$$\begin{aligned} {\text {d}}\epsilon = T{\text {d}}\eta + \frac{p}{\rho ^{2}}\,{\text {d}}\rho , \end{aligned}$$
(32.211)

identifies \(\epsilon \) as a function of \(\eta \) and \(\rho \), so that, alternatively and with \(\epsilon = \hat{\epsilon }(\eta , \rho )\),

$$\begin{aligned} {\text {d}}\epsilon = \frac{\partial \hat{\epsilon }}{\partial \eta }\,{\text {d}}\eta + \frac{\partial \hat{\epsilon }}{\partial \rho }\,{\text {d}}\rho . \end{aligned}$$
(32.212)

Comparison of (32.211) and (32.212) implies

$$\begin{aligned} T = \frac{\partial \hat{\epsilon }}{\partial \eta }, \quad p = \rho ^{2}\,\frac{\partial \hat{\epsilon }}{\partial \rho }. \end{aligned}$$
(32.213)

The internal energy, interpreted as a function of entropy \(\eta \) and density \(\rho \), is a thermodynamic potential for the absolute temperature and the pressure.

With the functions

$$\begin{aligned} \begin{array}{ll} \displaystyle \psi = \epsilon - T\eta &{} {\textsc {Helmholtz}}\,\text {free energy}, \\ &{} \\ \displaystyle h = \epsilon + \frac{p}{\rho } &{} \text {enthalpy}, \\ &{} \\ \displaystyle g = h - T\eta &{} {\textsc {Gibbs}}\,\text {free energy}, \end{array} \end{aligned}$$
(32.214)

(these are Legendre transformations) the Gibbs relation (32.211) takes the alternative forms

$$\begin{aligned} \begin{array}{ccc} \displaystyle {\text {d}}\psi = - \eta {\text {d}}T + \frac{p}{\rho ^{2}}\,{\text {d}}\rho &{} \longrightarrow &{} \displaystyle \psi = \hat{\psi }(T, \rho ), \\ \displaystyle {\text {d}}h = - T{\text {d}}\eta + \frac{1}{\rho }\,{\text {d}}p &{} \longrightarrow &{} \displaystyle h = \hat{h}(\eta , p), \\ \displaystyle {\text {d}}g = - \eta {\text {d}}T + \frac{1}{\rho }\,{\text {d}}p &{} \longrightarrow &{} \displaystyle g = \hat{g}(T, p). \end{array} \end{aligned}$$
(32.215)

With the indicated different dependencies and the obvious potential properties, analogous to (32.213), we have

$$\begin{aligned} \begin{array}{ll} \displaystyle \eta = - \frac{\partial \hat{\psi }}{\partial T}, &{} \displaystyle p = \rho ^{2}\frac{\partial \hat{\psi }}{\partial \rho } , \\ \displaystyle T = - \frac{\partial \hat{h}}{\partial \eta } , &{} \displaystyle \frac{1}{\rho } = \frac{\partial \hat{h}}{\partial p}, \\ \displaystyle \eta = - \frac{\partial \,\hat{g}}{\partial T}, &{}\displaystyle \frac{1}{\rho } = \frac{\partial \hat{g}}{\partial p}, \end{array} \end{aligned}$$
(32.216)

and the integrability conditions

$$\begin{aligned} \begin{array}{ll} \displaystyle -\frac{\partial \eta }{\partial \rho } \equiv \frac{\partial }{\partial T}\left( \frac{p}{\rho ^{2}}\right) &{} \displaystyle \text {for}\quad \hat{\psi }(T, \rho ), \\ \displaystyle -\frac{\partial T}{\partial p} \equiv \frac{\partial }{\partial T}\left( \frac{1}{\rho }\right) &{} \displaystyle \displaystyle \text {for}\quad \hat{h}(T, p),\\ \displaystyle -\frac{\partial \eta }{\partial p} \equiv \frac{\partial }{\partial T}\left( \frac{1}{\rho }\right) &{} \displaystyle \text {for}\quad \hat{g}(T, p). \end{array} \end{aligned}$$
(32.217)

Internal energy formulation. If we regard \(T\) and \(\rho \) as the independent thermodynamic variables, then according to (32.214)\(_1\) we have

$$\begin{aligned} \epsilon = \psi - T\frac{\partial \psi }{\partial T} = - T^{2}\frac{\partial }{\partial T}\left( \frac{\psi }{T}\right) , \end{aligned}$$
(32.218)

and therefore,

$$\begin{aligned} \begin{array}{c} \displaystyle \rho \frac{{\text {d}}\epsilon }{{{\text {d}}}{t}} = \rho \,c_{v}\frac{{\text {d}}T}{{{\text {d}}}{t}} + \rho \,c_{T\rho }\frac{{\text {d}}\rho }{{{\text {d}}}{t}},\\ \displaystyle c_{v}: = - \frac{\partial }{\partial T}\left( T^{2}\frac{\partial }{\partial T} \left( \frac{\hat{\psi }}{T}\right) \right) = \frac{\partial \hat{\epsilon }}{\partial T},\\ \displaystyle c_{T\rho } := - T^{2}\frac{\partial }{\partial T}\left( \frac{\partial \hat{\psi }/\partial \rho }{T}\right) = \frac{\partial \hat{\epsilon }}{\partial \rho }. \end{array} \end{aligned}$$
(32.219)

With the separation assumption

$$\begin{aligned} \psi = \hat{\psi }_{T}(T) + \hat{\psi }_{\rho }(\rho ), \end{aligned}$$
(32.220)

\(c_{v} = \hat{c}_{v}(T)\) and \(c_{T\rho } = \hat{c}_{T\rho }(\rho ) = {\text {d}}\hat{\psi }_{\rho }/{{\text {d}}}\rho \). Therefore, (32.219)\(_1\) can be written as

$$\begin{aligned} \rho \frac{{\text {d}}\epsilon }{{{\text {d}}}{t}} = \rho \,\hat{c_{v}}(T)\frac{{{\text {d}}}T}{{{\text {d}}}{t}} + \underbrace{\rho \frac{{\text {d}}\hat{\psi }_{\rho }}{{{\text {d}}}\rho } \frac{{{\text {d}}}\rho }{{{\text {d}}}{t}}}_{\text {nearly}\; 0} \approx \rho \hat{c}_{v}(T)\frac{{{\text {d}}}T}{{{\text {d}}}{t}}. \end{aligned}$$
(32.221)

The second term on the right-hand side of (32.221) can be ignored since density variations in a nearly incompressible fluid are minute.

Enthalpy formulation. If we regard \(T\) and \(p\) as the independent thermodynamic variables, the Gibbs free energy is the thermodynamic potential and the enthalpy the adequate internal energy function. In view of (32.215) we now have

$$\begin{aligned} h = g - T\frac{\partial g}{\partial T }= - T^{2}\frac{\partial }{\partial T}\left( \frac{g}{T}\right) , \end{aligned}$$
(32.222)

and therefore,

$$\begin{aligned} \begin{array}{c} \displaystyle \rho \frac{{\text {d}}h}{{{\text {d}}}{t}} =\rho \,c_{p}\frac{{\text {d}}t}{{{\text {d}}}{t}} + \rho \,c_{Tp}\frac{{\text {d}}p}{{{\text {d}}}{t}}, \\ \displaystyle c_{p}: = - \frac{\partial }{\partial T}\left( T^{2}\frac{\partial }{\partial T} \left( \frac{\hat{g}}{T}\right) \right) = \frac{\partial \hat{h}}{\partial T},\\ \displaystyle c_{Tp} := - T^{2}\frac{\partial }{\partial T}\left( \frac{1}{T}\frac{\partial g}{\partial p}\right) = \frac{\partial \hat{h}}{\partial p}. \end{array} \end{aligned}$$
(32.223)

With the separation assumption

$$\begin{aligned} h = \hat{g}_{T}(T) + \hat{g}_{p}(p), \end{aligned}$$
(32.224)

\(c_{p} = \hat{c}_{p}(T)\) and \(c_{Tp} = \hat{c}_{Tp}(p) = {\text {d}}\hat{g}_{p}/{{\text {d}}}p\). Therefore, (32.210)\(_1\) can be written as

$$\begin{aligned} \rho \frac{{\text {d}}h}{{{\text {d}}}{t}} = \rho \,\hat{c_{p}}(T)\frac{{{\text {d}}}T}{{{\text {d}}}{t}} + \underbrace{\rho \frac{{\text {d}}\hat{g}_{p}}{{{\text {d}}}p}\frac{{{\text {d}}}p}{{{\text {d}}}{t}}}_{\text {nearly}\; 0} \, \approx \, \rho \hat{c}_{p}(T)\frac{{\text {d}}t}{{{\text {d}}}{t}}. \end{aligned}$$
(32.225)

Here the second term on the right-hand side can be ignored, since \({\text {d}}\hat{g}_{p}/{\text {d}}p\) must be very small, the growth of the enthalpy due to a pressure rise cannot be large as its working is due to dilatational deformations, which are small.

Parameterizations. Because the temperature range of lake or ocean water is small, \(0\,^\circ \)C\(\, \leqslant T \leqslant 50\,^\circ \)C, the coefficients \(c_v\) and \(c_p\) exhibit a constrained variability and may well be assumed to be constant or linear functions of \(T\). This then suggests to use

  • for constant specific heats,

    $$\begin{aligned} \begin{array}{c} \displaystyle \epsilon = \int _{T_0}^{T}c_{v}(\bar{T})\,{\text {d}}\bar{T} = c_{v}^{0}(T - T_0) + \epsilon _0,\quad \\ h = \displaystyle \int _{T_0}^{T}c_{p}(\bar{T})\,{\text {d}}\bar{T} = c_{p}^{0}(T - T_0) + h_0, \end{array} \end{aligned}$$
    (32.226)
  • for specific heats as linear functions of T:

    $$\begin{aligned} \displaystyle \epsilon&= \int _{T_0}^{T}[c_{v}^0 + c_{v}^{\prime }(\bar{T} - T_0)]\,{\text {d}}\bar{T} = c_{v}^{0}(T - T_0) + \frac{1}{2}c_{v}^{\prime }(T - T_0)^{2} + \epsilon _0,\nonumber \\ \displaystyle h&= \int _{T_0}^{T}[c_{p}^0 + c_{p}^{\prime }(\bar{T} - T_0)]\,{\text {d}}\bar{T} = c_{p}^{0}(T - T_0) + \frac{1}{2}c_{p}^{\prime }(T - T_0)^{2} + h_0. \end{aligned}$$
    (32.227)

The expressions (32.221), (32.225), (32.226), (32.227) provide a thermodynamic justification of relations (32.44).

Appendix B: Turbulent Closure by Large Eddy Simulation

Large Eddy Simulation (LES) is another popular approach for simulating turbulent flows. In this technique the large, geometry-dependent eddies are explicitly accounted for by using a subgrid-scale (SGS) model. Equations (32.76)–(32.80) are now interpreted as resolved field equations obtained by applying a non-statistical filter to the Navier–Stokes equations. Footnote 33

The effect of the small eddies on the resolved filtered field is included in the SGS-parameterization of the stress \({\varvec{R}}\), given by (32.52) but by

$$\begin{aligned} {\varvec{R}}= 2\rho \nu _{SGS}{\varvec{D}}, \quad \mathrm {tr}\,{\varvec{D}}= 0, \end{aligned}$$
(32.228)

where \(\nu _{SGS}\) is the SGS-turbulent viscosity,

$$\begin{aligned} \nu _{SGS} \equiv (C_{s}\varDelta )^{2}\left( \mathrm {tr}\,(2{\varvec{D}}^{2})\right) ^{1/2}. \end{aligned}$$
(32.229)

This parameterization is due to Smagorinsky (1963) [42]. \(C_{s}\) is a dimensionless coefficient, called Smagorinsky constant, and \(\varDelta \) is a length scale, equal to the local grid spacing. Thus, (32.228) with (32.229) is the classical viscous power law relating stress and stretching. According to Kraft et al. [24], the above ‘model is found to give acceptable results in LES of homogeneous and isotropic turbulence. With \(C_{s} \approx 0.17\) according to Lilly (1967) [26], it is too dissipative [\(\ldots \)] in the near wall region because of the excessive eddy-viscosity arising from the mean shear (Moin and Kim (1982) [30]). The eddy viscosity predicted by Smagorinsky is nonzero in laminar flow regions; the model introduces spurious dissipation which damps the growth of small perturbations and thus restrains the transition to turbulence (Piomelli and Zang (1991) [34]).

The limitations of the Smagorinsky model have led to the formulation of more general SGS models. The best known of these newer models may be the dynamic SGS (DSGS) model of Germano et al. (1991) [12]. In this model \(C_{s}\) is not a fixed constant but is calculated as a function of position and time, \(C_{s}(\mathbf{x}, t)\), which vanishes near the boundary with the correct behaviour (Piomelli (1993) [24, 33]).

The parameterisations for the energy flux, \({\varvec{Q}}_{\epsilon }\) and constituent mass fluxes, \({\varvec{J}}_{\alpha }\), are the same as stated in (32.52)\(_{2,3}\), however, with \(\nu _{SGS}\) evaluated as given in (32.229). It is also evident from this presentation that the \((k-\varepsilon )\)-equations are not needed.

Appendix C: Justification for (32.157)

In this appendix we provide a derivation of formula (32.157) for erosion inception on the basis of dimensional analysis. We consider sediment transport at a lake basal surface. It is rather intuitive that the erosion inception will likely depend on a stress (the shear stress) on the lake side of the basal surface, \(\tau _c\), the true densities, \(\rho _s\), \(\rho _f\), of the sediment grains and the fluid, the solid concentration, \(c_s\), gravity acceleration, \(g\), mixture kinematic viscosity, \(\nu \), and the nominal diameter, \({\mathfrak d}\), of the sediment corn, all evaluated at the base. So, inception of sediment transport can likely be described by an equation of the form

$$\begin{aligned} f(\tau _c, \rho _s, \rho _f, g, {\mathfrak d}, \nu , c_s) = 0. \end{aligned}$$
(32.230)

The dimensional matrix of the above 7 variables has rank 3; so, there are 4 dimensionless \(\pi \)-products, which we choose as follows:

$$\begin{aligned} \pi _1 = \frac{\tau _c}{\varDelta \rho \,g\,{\mathfrak d}}, \quad \pi _2 = \frac{\rho _s}{\rho _f}, \quad \pi _3 = c_s, \quad \pi _4 = \left( \frac{g}{\varDelta \nu ^{2}}\right) ^{1/3}{\mathfrak d}, \end{aligned}$$
(32.231)

where \(\rho \) is the mixture density and \(\varDelta \equiv (\rho _s/\rho - 1)\). Here, \(\tau _c\) has been scaled with the ‘submerged’ density \((\rho _s - \rho )\). Furthermore, it is not difficult to see that for small \(c_s\) the mixture density in (32.231) may approximately be replaced by \(\rho _f\). We may thus write

$$\begin{aligned} f(\pi _1, \pi _2, \pi _3, \pi _4) = 0 \quad \text {or} \quad \frac{\tau _c}{\varDelta \,\rho \,g{\mathfrak d}} = \tilde{f}(\pi _2, \pi _3, \pi _4). \end{aligned}$$
(32.232)

The number of variables is now reduced from 7 to 4, a dramatic reduction. However, even further reduction is possible. For sediment transport in the geophysical environment \(\pi _2\) is very nearly a constant on the entire Globe, and \(\pi _3\) is very small (\(\leqslant \!\!10^{-2}\)); so, the \(\pi _3\)-dependence may be dropped (i.e. expressed in a Taylor series expansion of \(\pi _3\) and restricted to the term \(\tilde{f}(\pi _2, 0, \pi _4))\). Thus, we may assume

$$\begin{aligned} \theta _c \equiv \frac{\tau _c}{\varDelta \,\rho \,g{\mathfrak d}} = \tilde{f}(Re_c^{*}) = \tilde{f}({\mathfrak d}^{*}),\quad \pi _4 = Re_{c}^{*} = {\mathfrak d}^*\equiv \left( \frac{g}{\varDelta \,\nu ^{2}}\right) ^{1/3}{\mathfrak d} . \end{aligned}$$
(32.233)

This derivation assumes that only a single sediment fraction is present. It is important to note that the viscosity \(\nu \) of the mixture is present in the variables describing the erosion inception. If it is dropped, then \(\tilde{f}\) in (32.233) reduces to a constant and

$$\begin{aligned} \tau _c = \text {const.}\times \varDelta \,\rho \,g{\mathfrak d}^*, \end{aligned}$$

which is not supported by experiments. Omitting \(g\) as a governing parameter is disastrous, because \(\pi _1\) and \(\pi _4\) are then missing as \(\pi \)-products. In this case \(\tilde{f}(\pi _2, \pi _3) = 0\) is simply meaningless.

Appendix D: Justification for (32.177), (32.178) and (32.182), (32.183)

Justification for (32.177), (32.178). For the constituent masses, noting that

$$\begin{aligned} {\rho _{\alpha }({\varvec{v}}_{\alpha } - {\varvec{w}})}=\underbrace{\rho _{\alpha }({\varvec{v}}_{\alpha } - {\varvec{v}})}_{\equiv \,{\displaystyle {{\mathfrak {J}}}}_{\alpha },\;\, \text {see eq. (8)}}{ \displaystyle +\rho _{\alpha }({\varvec{v}} - {\varvec{w}})}, \end{aligned}$$

the non-averaged balance (32.140), in which \(f_{\mathcal {S}}=\mu _\alpha \), \({{\varvec{\phi }}}^{f_{\mathcal {S}}}=\varvec{ {0}}\), \(f=\rho _\alpha \), \({\varvec{v}}={\varvec{v}}_\alpha \), \({{\varvec{\phi }}}^{f}=\varvec{ {0}}\), can be written as

$$\begin{aligned} \begin{array}{c} \displaystyle \frac{\partial \mu _\alpha }{\partial t}+\left( \mu _{\alpha } {\varvec{v}}_{{\mathcal {S}}\alpha }\right) ^{\mathfrak a}_{\;;\mathfrak a} -\frac{\partial \mu _\alpha }{\partial \xi ^{\mathfrak a}}\,w^{\mathfrak a}- 2\mu _{\alpha }\,{\mathcal U}_{b} K= - [\![{\varvec{{\mathfrak {J}}}}_{\alpha } + \rho _{\alpha }({\varvec{v}} - {\varvec{w}}) ]\!]\cdot {\varvec{n}}_b. \end{array} \end{aligned}$$
(32.234)

Analogously, for the fluid we deduce

$$\begin{aligned} \displaystyle \frac{\partial \mu _f}{\partial t}+\left( \mu _{f} {\varvec{v}}_{{{\mathcal {S}}}f}\right) ^{\mathfrak a}_{\;;\mathfrak a} -\frac{\partial \mu _f}{\partial \xi ^{\mathfrak a}}\,w^{\mathfrak a}- 2\mu _{f}\,{\mathcal U}_{b} K =- [\![{\varvec{{\mathfrak {J}}}}_{f} + \tilde{\rho }_{f}({\varvec{v}} - {\varvec{w}}) ]\!]\cdot {\varvec{n}}_b, \end{aligned}$$
(32.235)

where \({\varvec{{\mathfrak {J}}}}_{f}\equiv \tilde{\rho }_f({\varvec{v}}_{f} - {\varvec{v}})\), with \(\tilde{\rho }_f\) and \({\varvec{v}}_f\) the mass density and velocity of the fluid (\(\tilde{\rho }_f=\rho -\sum _\alpha \rho _\alpha \)). Now we sum Eqs. (32.234) and (32.235) over all constituents. Using relation

$$\begin{aligned} \sum _{\alpha }{\varvec{{\mathfrak {J}}}}_{\alpha } + {\varvec{{\mathfrak {J}}}}_{f} = \varvec{ {0}}, \end{aligned}$$
(32.236)

and definitions

$$\begin{aligned} \mu \equiv \sum _{\alpha }\mu _{\alpha } + \mu _{f},\quad \mu {\varvec{v}}_{{\mathcal {S}}} \equiv \sum _{\alpha }\mu _{\alpha }{\varvec{v}}_{{{\mathcal {S}}}\alpha } + \mu _{f}{\varvec{v}}_{{{\mathcal {S}}}f} \end{aligned}$$
(32.237)

for the mixture surface density \(\mu \) and mixture velocity \({\varvec{v}}_{{\mathcal {S}}}\), we obtain the mass balance for the mixture by summation of (32.234) and (32.235):

$$\begin{aligned} \displaystyle \frac{\partial \mu }{\partial t}+\left( \mu {\varvec{v}}_{{\mathcal {S}}}\right) ^{\mathfrak a}_{\;;\mathfrak a} -\frac{\partial \mu }{\partial \xi ^{\mathfrak a}}\,w^{\mathfrak a}- 2\mu \,{\mathcal U}_{b} K = - [\![\rho ({\varvec{v}} - {\varvec{w}}) ]\!]\cdot {\varvec{n}}_b . \end{aligned}$$
(32.238)

We now average Eqs. (32.234) and (32.235). In so doing we assume that the interface does not perform any fluctuations, whence necessarily \( \langle \,{\varvec{n}}_b\,\rangle = {\varvec{n}}_b\), \( \langle \,K\,\rangle = K\), \(\langle \,{\varvec{w}}\,\rangle = {\varvec{w}}\) and \(\langle {\mathcal U}_{b}\rangle = {\mathcal U}_{b}\). Thus, for the averaged equations we get

$$\begin{aligned} \begin{array}{r} \displaystyle \frac{\partial \langle \mu _\alpha \rangle }{\partial t}+\left( \langle \mu _{\alpha }\rangle \, \langle \,{\varvec{v}}_{{\mathcal {S}}\alpha }\,\rangle \right) ^{\mathfrak a}_{\;;\mathfrak a} +\left( \langle \,\mu _{\alpha }^{\prime } \,\left( {\varvec{v}}_{{{\mathcal {S}}}\alpha }\right) ^{\prime }\,\rangle \right) ^{\mathfrak a}_{\;;\mathfrak a}-\frac{\partial \langle \mu _\alpha \rangle }{\partial \xi ^{\mathfrak a}}\,w^{\mathfrak a}- 2\langle \mu _{\alpha }\rangle \,{\mathcal U}_{b} K\\ = - [\![\langle \,{\varvec{{\mathfrak {J}}}}_{\alpha }\rangle +\langle \,\rho _{\alpha }^{\prime }{\varvec{v}} ^{\prime }\,\rangle + \langle \,\rho _{\alpha }\,\rangle ({\langle \,\varvec{v}}\,\rangle - {\varvec{w}}) ]\!]\cdot {\varvec{n}}_b, \end{array} \end{aligned}$$
(32.239)
$$\begin{aligned} \begin{array}{r} \displaystyle \frac{\partial \langle \mu _f\rangle }{\partial t}+\left( \langle \mu _{f}\rangle \, \langle \,{\varvec{v}}_{{{\mathcal {S}}}f}\,\rangle \right) ^{\mathfrak a}_{\;;\mathfrak a} +\left( \langle \,\mu _{f}^{\prime } \,\left( {\varvec{v}}_{{{\mathcal {S}}}f}\right) ^{\prime }\,\rangle \right) ^{\mathfrak a}_{\;;\mathfrak a} -\frac{\partial \langle \mu _f\rangle }{\partial \xi ^{\mathfrak a}}\,w^{\mathfrak a}- 2\langle \mu _{f}\rangle \,{\mathcal U}_{b} K \;\,\\ = - [\![\langle \,{\varvec{{\mathfrak {J}}}}_{f}\rangle + \langle \,\tilde{\rho }_{f}^{\prime }{\varvec{v}} ^{\prime }\,\rangle + \langle \,\tilde{\rho }_{f}\,\rangle ({\langle \,\varvec{v}}\,\rangle - {\varvec{w}}) ]\!]\cdot {\varvec{n}}_b. \end{array} \end{aligned}$$
(32.240)

If we sum (32.239) and (32.240), because of (32.236), (32.237) we obtain

$$\begin{aligned} \displaystyle \frac{\partial \langle \mu \rangle }{\partial t}&+\left( \langle \mu \rangle \, \langle \,{\varvec{v}}_{{\mathcal {S}}}\,\rangle \right) ^{\mathfrak a}_{\;;\mathfrak a} +\left( \langle \,\mu ^{\prime } \,\left( {\varvec{v}}_{{\mathcal {S}}}\right) ^{\prime }\,\rangle \right) ^{\mathfrak a}_{\;;\mathfrak a} -\frac{\partial \langle \mu \rangle }{\partial \xi ^{\mathfrak a}}\,w^{\mathfrak a}- 2\langle \mu \rangle \,{\mathcal U}_{b} K \nonumber \\&\qquad \quad = - [\![\qquad \qquad \underbrace{\langle \,\rho ^{\prime }{\varvec{v}}^{\prime }\,\rangle }\qquad +\, \langle \,\rho \,\rangle (\langle {\varvec{v}}\,\rangle - {\varvec{w}})]\!]\cdot {\varvec{n}}_b.\\&\qquad \qquad \quad \quad \equiv {{\varvec{\phi }}}^{\rho } z {\text {in Table 32.6}}\nonumber \end{aligned}$$
(32.241)

Of course, (32.241) is the average of (32.238), and only two of (32.239)–(32.241) are independent. For computations of initial boundary value problems we recommend to use (32.239) and (32.241) and to infer \(\langle \mu _{f}\rangle \) a posteriori from \(\langle \mu _{f}\rangle =\langle \mu \rangle - \sum \nolimits _{\alpha }\langle \mu _{\alpha }\rangle \).

It follows: with Reynolds averaging we have a non-vanishing mass flux in the mass balance (32.241). A Favre-type averaging would have to be performed. However, if \(\rho ^{\prime }\) is small on both sides of the basal surface we can drop \(\langle \,\rho ^{\prime }{\varvec{v}}^{\prime }\,\rangle \) in (32.241). Moreover, with \(\rho ^{\prime }\approx 0\), \(\rho _\alpha =\rho c_\alpha \), decomposition (32.9) and definition of \({\varvec{J}}_\alpha \) [see (32.43)], for the constituent class \(\alpha \) the mass flux \(\langle \,{\varvec{{\mathfrak {J}}}}_{\alpha }\rangle +\langle \,\rho _{\alpha }^{\prime }{\varvec{v}} ^{\prime }\,\rangle \) takes the form

$$\begin{aligned} \langle \,{\varvec{{\mathfrak {J}}}}_{\alpha }\rangle +\langle \,\rho _{\alpha }^{\prime }{\varvec{v}} ^{\prime }\,\rangle ={\varvec{J}}_{\alpha }-\rho \,\langle \,c_{\alpha }\rangle \langle {\varvec{w}}^{s}_\alpha \rangle , \end{aligned}$$

which explains Table 32.6 for Model 2. The main text, formulae (32.177), (32.178) [as deduced from (32.239), (32.241)] and Table 32.6 show the averaged fields without the averaging operator \(\langle \cdot \rangle \) and with negligible correlations

$$\begin{aligned} \langle \,\mu _{\alpha }^{\prime } \,\left( {\varvec{v}}_{{{\mathcal {S}}}\alpha }\right) ^{\prime }\,\rangle ,\quad \langle \,\mu ^{\prime } \,\left( {\varvec{v}}_{{\mathcal {S}}}\right) ^{\prime }\,\rangle . \end{aligned}$$

Justification for (32.182) and (32.183). Now we consider (32.139), in which \(f_{\mathcal {S}}=\mu _\alpha {\varvec{v}}_{{\mathcal {S}}\alpha }\), \({{\varvec{\phi }}}^{f_{\mathcal {S}}}=-{{\varvec{\sigma }}}_{{\mathcal {S}}\alpha }\), \(\pi ^{f_{\mathcal {S}}}=0\), \(s^{f_{\mathcal {S}}}=\mu _\alpha {{\varvec{g}}}\), \(f=\rho _\alpha {\varvec{v}}_\alpha \), \({\varvec{v}}={\varvec{v}}_\alpha \), \({{\varvec{\phi }}}^{f}=-{{\varvec{\sigma }}}_\alpha \), for each \(\alpha =1,\ldots ,N\):

$$\begin{aligned}&\displaystyle \quad \frac{\partial }{\partial t}\left( \mu _\alpha {\varvec{v}}_{{\mathcal {S}}\alpha }\right) +\text {Div}\,\left( \mu _\alpha {\varvec{v}}_{{\mathcal {S}}\alpha }\otimes {\varvec{v}}_{{\mathcal {S}}\alpha }-{{\varvec{\sigma }}}_{{\mathcal {S}}\alpha }\right) -\frac{\partial }{\partial \xi ^{\mathfrak a}}(\mu _\alpha {\varvec{v}}_{{\mathcal {S}}\alpha })\,w^{\mathfrak a} \nonumber \\ \displaystyle&= -[\![\rho _\alpha {\varvec{v}}_\alpha \otimes ({\varvec{v}}_\alpha -{\varvec{w}})-{\varvec{\sigma }}_\alpha ]\!]\,{\varvec{n}}_b+\mu _\alpha {{\varvec{g}}}. \end{aligned}$$
(32.242)

A similar equation holds for the interstitial fluid:

$$\begin{aligned}&\displaystyle \frac{\partial }{\partial t}\left( \mu _f{\varvec{v}}_{{\mathcal {S}}f}\right) +\text {Div}\left( \mu _f{\varvec{v}}_{{\mathcal {S}}f}\otimes {\varvec{v}}_{{\mathcal {S}}f }-{{\varvec{\sigma }}}_{{\mathcal {S}}f}\right) - \frac{\partial }{\partial \xi ^{\mathfrak a}}(\mu _f{\varvec{v}}_{{\mathcal {S}}f})\,w^{\mathfrak a} \nonumber \\&=\displaystyle -[\![\tilde{\rho }_f {\varvec{v}}_f\otimes ({\varvec{v}}_f-{\varvec{w}})-{\varvec{\sigma }}_f]\!]\,{\varvec{n}}_b+\mu _f{{\varvec{g}}}. \end{aligned}$$
(32.243)

Summing (32.242), (32.243) and using definitions (32.237) we obtain

$$\begin{aligned}&\displaystyle \frac{\partial }{\partial t}\left( \mu {\varvec{v}}_{{\mathcal {S}}}\right) +\text {Div}\left( \mu {\varvec{v}}_{{\mathcal {S}}}\otimes {\varvec{v}}_{{\mathcal {S}}}-{{\varvec{\sigma }}}_{{\mathcal {S}}}\right) - \frac{\partial }{\partial \xi ^{\mathfrak a}}(\mu {\varvec{v}}_{{\mathcal {S}}})\,w^{\mathfrak a} \nonumber \\&\quad = \displaystyle -[\![\rho {\varvec{v}}\otimes ({\varvec{v}}-{\varvec{w}})-{\varvec{\sigma }}]\!]\,{\varvec{n}}_b+\mu {{\varvec{g}}}, \end{aligned}$$
(32.244)

where the bulk, \({\varvec{\sigma }}\), and surface, \({\varvec{\sigma }}_{\mathcal {S}}\), mixture stress tensors are defined by

$$\begin{aligned} \rho {\varvec{v}}\otimes {\varvec{v}}-{{\varvec{\sigma }}} \equiv \sum _\alpha \left( \rho _\alpha {\varvec{v}}_{\alpha }\otimes {\varvec{v}}_{\alpha }-{{\varvec{\sigma }}}_{\alpha }\right) +\tilde{\rho }_f {\varvec{v}}_{ f}\otimes {\varvec{v}}_{{\mathcal {S}}f}-{{\varvec{\sigma }}}_{ f}, \end{aligned}$$
(32.245)
$$\begin{aligned} \mu {\varvec{v}}_{{\mathcal {S}}}\otimes {\varvec{v}}_{{\mathcal {S}}} -{{\varvec{\sigma }}}_{\mathcal {S}}\equiv \sum _\alpha \left( \mu _\alpha {\varvec{v}}_{{\mathcal {S}}\alpha }\otimes {\varvec{v}}_{{\mathcal {S}}\alpha }-{{\varvec{\sigma }}}_{{\mathcal {S}}\alpha }\right) +\mu _f {\varvec{v}}_{{\mathcal {S}}f}\otimes {\varvec{v}}_{{\mathcal {S}}f}-{{\varvec{\sigma }}}_{{\mathcal {S}}f}. \end{aligned}$$
(32.246)

Averaging (32.244) under the assumptions \(\mu ^\prime \approx 0\), \(\rho ^{\prime }\approx 0\), recalling definition (32.43)\(_1\) of the Reynolds stress tensor \({\varvec{R}}\) and introducing the laminar and turbulent surface mixture stress tensor \({{\varvec{R}}}_{{\mathcal {S}}}\) according to

$$\begin{aligned} {{\varvec{R}}}_{{\mathcal {S}}}\equiv \langle {{\varvec{\sigma }}}_{{\mathcal {S}}}\rangle - \mu \,\langle {\varvec{v}}_{{\mathcal {S}}}^\prime \otimes {\varvec{v}}_{{\mathcal {S}}}^\prime \rangle , \end{aligned}$$
(32.247)

we deduce (we omit the angular brackets)

$$\begin{aligned} \begin{array}{r} \displaystyle \frac{\partial }{\partial t}\left( \mu {\varvec{v}}_{{\mathcal {S}}}\right) +\text {Div}\left( \mu {\varvec{v}}_{{\mathcal {S}}}\otimes {\varvec{v}}_{{\mathcal {S}}}-{{\varvec{R}}}_{{\mathcal {S}}}\right) - \frac{\partial }{\partial \xi ^{\mathfrak a}}\left( \mu {\varvec{v}}_{{\mathcal {S}}}\right) w^{\mathfrak a}=\quad \quad \quad \\ \displaystyle -[\![\rho {\varvec{v}}\otimes ({\varvec{v}}-{\varvec{w}})+ p\varvec{I}-{\varvec{R}}]\!]\,{\varvec{n}}_b+\mu {{\varvec{g}}}, \end{array} \end{aligned}$$
(32.248)

which explains the last line in Table 32.6.

Next we want to write (32.248) using the components of vectors and tensors with respect to the local basis \(\{{{\varvec{\tau }}}_1,{{\varvec{\tau }}}_2,{\varvec{n}}_b\}\), which will give (32.182) and (32.183). To this end we use the formulae (for simplicity in this derivation we omit the lower index \(b\) in \({\mathcal U}_b\) and \({\varvec{n}}_b\) referring to the basal surface)

$$\begin{aligned} \begin{array}{c} \displaystyle \frac{\partial {{\varvec{\tau }}}_{\mathfrak a}}{\partial \xi ^{\mathfrak b}}=\varGamma ^{\mathfrak c}_{\mathfrak a\mathfrak b}{{\varvec{\tau }}}_{\mathfrak c} + b_{\mathfrak a\mathfrak b}{\varvec{n}},\quad \frac{\partial {{\varvec{n}}}}{\partial \xi ^{\mathfrak a}}=- b_{\mathfrak a\mathfrak b}{{\varvec{\tau }}}^{\mathfrak b} ,\quad \frac{\partial {\varvec{n}}}{\partial \,t}=-g^{\mathfrak a\mathfrak b} \left\{ \frac{\partial {\mathcal U}}{\partial \xi ^{\mathfrak a}}+b_{\mathfrak c\mathfrak b}w^{\mathfrak c}\right\} {{\varvec{\tau }}_{\mathfrak b}},\\ \displaystyle \frac{\partial {{\varvec{\tau }}}_{\mathfrak a}}{\partial \,t}=\frac{\partial {\varvec{w}}}{\partial \xi ^{\mathfrak a}}=\left\{ \frac{\partial w^{\mathfrak b}}{\partial \xi ^{\mathfrak a}}+w^{\mathfrak c}\varGamma ^{\mathfrak b}_{\mathfrak c\mathfrak a}-{\mathcal U}b_{\mathfrak a\mathfrak c}g^{\mathfrak c\mathfrak b}\right\} {{\varvec{\tau }}_{\mathfrak b}}+ \left\{ \frac{\partial \mathcal U}{\partial \xi ^{\mathfrak a}}+ w^{\mathfrak b}b_{\mathfrak b\mathfrak a} \right\} {\varvec{n}}.\\ \end{array} \end{aligned}$$
(32.249)

The first two relations are derived in (32.111) and (32.112). The third and fourth are proved in an analogous way. Indeed, starting with the representations

$$\begin{aligned}&\frac{\partial n^{k}}{\partial t} = \dot{a}n^{k} + \dot{b}_{{\mathfrak a}}\tau ^{{\mathfrak a}k}\; \left( = \dot{a}n^{k} + \dot{b}^{{\mathfrak b}}\tau _{{\mathfrak b}}^{k}\right) ,\nonumber \\ \\&\frac{\partial \tau _{{\mathfrak a}}^{k}}{\partial t} = \dot{d}_{{\mathfrak a}}n^{k} + \dot{e}_{{\mathfrak a}}^{{\mathfrak b}}\tau _{{\mathfrak b}}^{k},\nonumber \end{aligned}$$
(32.250)

with yet unknown \(\dot{a}\), \(\dot{b}^{{\mathfrak b}}\), \(\dot{e}_{\mathfrak a}^{\mathfrak b}\), we readily deduce

$$\begin{aligned}&\dot{a} =\frac{\partial n^{k}}{\partial t}n_{k} = \frac{1}{2}\frac{\partial (n^{k}\cdot n_{k})}{\partial t} = \frac{1}{2}\frac{\partial (1)}{\partial t} = 0 \nonumber \\&\dot{b}^{{\mathfrak b}} = \frac{\partial n^{k}}{\partial t}\tau _{k}^{{\mathfrak b}} = \dot{b}^{{\mathfrak a}} \underbrace{\tau _{{\mathfrak a}}^{k}\tau ^{{\mathfrak b}}_{k}}_{\delta _{{\mathfrak a}}^{{\mathfrak b}}} = \dot{b}^{{\mathfrak a}}\delta _{{\mathfrak a}}^{{\mathfrak b}}, \\&\frac{\partial \tau _{{\mathfrak a}}^{k}}{\partial t}n^{k} = \dot{d}_{{\mathfrak a}}, \nonumber \\&\frac{\partial \tau _{{\mathfrak a}}^{k}}{\partial t}\tau _{k}^{{\mathfrak c}} = \dot{e}_{{\mathfrak a}}^{{\mathfrak b}} \tau _{{\mathfrak b}}^{k} \tau _{k}^{{\mathfrak c}} = \dot{e}_{\mathfrak a} ^{{\mathfrak b}}g_{\mathfrak b\mathfrak c} = \dot{e}_{{\mathfrak a}}^{\mathfrak c}.\nonumber \end{aligned}$$
(32.251)

Explicit evaluation of \(\dot{b}_{\mathfrak b}\) follows by recognizing that

$$\begin{aligned} \frac{\partial }{\partial t}\left( n^{k}\tau _{{\mathfrak b}k}\right) = 0 \quad \longrightarrow \quad \underbrace{\frac{\partial n^{k}}{\partial t} \tau _{{\mathfrak b}k}}_{\dot{b}_{{\mathfrak b}}} + n^{k}\frac{\partial \tau _{{\mathfrak b}}}{\partial t} = 0. \end{aligned}$$
(32.252)

So,

$$\begin{aligned} \dot{b}_{{\mathfrak b}}&= - n^{k}\frac{\partial \tau _{{\mathfrak b}k}}{\partial t} = -n^{k}\left\{ \frac{\partial }{\partial t} \left( \frac{\partial r_{k}}{\partial \xi ^{{\mathfrak b}}}\right) \right\} = - n^{k}\left\{ \frac{\partial }{\partial \xi ^{{\mathfrak b}}} \underbrace{\left( \frac{\partial r_{k}}{\partial t}\right) }_{=w^{{\mathfrak c}}\tau _{k{\mathfrak c}}+{\mathcal U}n_{k}}\right\} \nonumber \\&= - n^{k}\left\{ \frac{\partial w^{{\mathfrak c}}}{\partial \xi ^{{\mathfrak b}}}\tau _{k{\mathfrak b}} + w^{{\mathfrak c}}\underbrace{\frac{\partial \tau _{k{\mathfrak c}}}{\partial \xi ^{{\mathfrak b}}}}_{\varGamma _{{\mathfrak b\mathfrak c}}^{{\mathfrak d}}\tau _{{\mathfrak d}k}+b_{{\mathfrak b\mathfrak c}}n_{k}} + \frac{\partial {\mathcal U}}{\partial \xi ^{{\mathfrak b}}}n_{k} + {\mathcal U}\frac{\partial n_{k}}{\partial \xi ^{{\mathfrak b}}}\right\} \nonumber \\&= -\left\{ \frac{\partial {\mathcal U}}{\partial \xi ^{{\mathfrak b}}} + b_{\mathfrak b\mathfrak c}w^{{\mathfrak c}}\right\} , \end{aligned}$$
(32.253)

proving (32.249)\(_3\). Similarly,

$$\begin{aligned} \dot{d}_{{\mathfrak a}}&= \frac{\partial \tau _{{\mathfrak a}}^{k}}{\partial t}n_{k} = \frac{\partial }{\partial t}\left\{ \frac{\partial r^{k}}{\partial \xi ^{{\mathfrak a}}}\right\} n_{k} = \frac{\partial }{\partial \xi ^{{\mathfrak a}}}\left\{ \frac{\partial r_{k}}{\partial t}\right\} n_{k} = \frac{\partial }{\partial \xi ^{{\mathfrak a}}}\left( w^{{\mathfrak b}}\tau _{{\mathfrak b}}^{k} + {\mathcal U}n^{k} \right) \nonumber \\&= \left\{ \frac{\partial w^{{\mathfrak b}}}{\partial \xi ^{{\mathfrak a}}}\tau _{{\mathfrak b}}^{k} + w^{{\mathfrak b}} \underbrace{\frac{\partial \tau _{\mathfrak b}^{k}}{\partial \xi ^{{\mathfrak a}}}}_{(\varGamma _{{\mathfrak a\mathfrak b}}^{\mathfrak c}\tau _{\mathfrak c}^{k}+b_{{\mathfrak a\mathfrak b}}n^{k}), \; \text {see} \; (32.111)_1} + \frac{\partial {\mathcal U}}{\partial \xi ^{{\mathfrak a}}}n^{k} + {\mathcal U}\frac{\partial n^{k}}{\partial \xi ^{{\mathfrak a}}}\right\} n^{k} \nonumber \\&= \left( w^{{\mathfrak b}}b_{{\mathfrak a\mathfrak b}} + \frac{\partial {\mathcal U}}{\partial \xi ^{{\mathfrak a}}}\right) , \end{aligned}$$
(32.254)

as well as

$$\begin{aligned} \dot{e}_{{\mathfrak a}}^{{\mathfrak c}}&= \frac{\partial \tau _{{\mathfrak a}}^{k}}{\partial t}\tau _{k}^{{\mathfrak c}} = \frac{\partial }{\partial t} \left\{ \frac{\partial r^{k}}{\partial \xi ^{{\mathfrak a}}}\right\} \tau _{k}^{{\mathfrak c}} = \frac{\partial }{\partial \xi ^{{\mathfrak a}}}\left( w^{{\mathfrak b}}\tau _{\mathfrak b}^{k} + {\mathcal U}n^{k}\right) \tau _{k}^{{\mathfrak c}} \nonumber \\&= \left\{ \frac{\partial w^{{\mathfrak b}}}{\partial \xi ^{{\mathfrak a}}}\tau _{{\mathfrak b}}^{k} + w^{{\mathfrak b}}\underbrace{ \frac{\partial \tau _{{\mathfrak b}}^{k}}{\partial \xi ^{{\mathfrak a}}}}_{\varGamma _{{\mathfrak a\mathfrak b}}^{{\mathfrak d}}\tau _{\mathfrak d}^{k}+b_{{\mathfrak a\mathfrak b}}n^{k}} + \frac{\partial {\mathcal U}}{\partial \xi ^{{\mathfrak a}}}n^{k} + {\mathcal U}\underbrace{\frac{\partial n^{k}}{\partial \xi ^{\mathfrak a}}}_{-b_{\mathfrak a\mathfrak d}\tau ^{{\mathfrak d}k}}\right\} \tau _{k}^{\mathfrak c} \nonumber \\&= \frac{\partial w^{{\mathfrak b}}}{\partial \xi ^{{\mathfrak a}}}\tau _{\mathfrak b}^{k}\tau _{k}^{{\mathfrak c}} + \left\{ w^{{\mathfrak b}}\left( \varGamma _{{\mathfrak a\mathfrak b}}^{\mathfrak d}\tau _{\mathfrak d}^{k} + b_{{\mathfrak a\mathfrak b}}n^{k}\right) + \frac{\partial {\mathcal U}}{\partial \xi ^{\mathcal a}}n^{k} + {\mathcal U}(-b_{{\mathcal ad}}\tau ^{{\mathfrak d}k})\right\} \tau _{k}^{\mathfrak c} \nonumber \\&= \frac{\partial w^{{\mathfrak b}}}{\partial \xi ^{{\mathfrak a}}}\delta _{{\mathfrak b}}^{{\mathfrak d}} + w^{{\mathfrak b}}\varGamma _{{\mathfrak a\mathfrak b}}^{{\mathfrak d}}\underbrace{\tau _{{\mathfrak d}}^{k}\tau _{k}^{{\mathfrak c}}}_{\delta _{{\mathfrak d}}^{{\mathfrak c}}} - b_{{\mathfrak a\mathfrak d}}{\mathcal U} \underbrace{\tau ^{{\mathfrak d}k}\tau _{k}^{{\mathfrak c}}}_{g^{{\mathfrak dc}}} \nonumber \\&= \frac{\partial w^{{\mathfrak c}}}{\partial \xi ^{{\mathfrak a}}} + w^{{\mathfrak b}}\varGamma _{{\mathfrak a\mathfrak b}}^{\mathfrak c} - b_{{\mathfrak a\mathfrak d}}g^{\mathfrak dc}. \end{aligned}$$
(32.255)

Relations (32.254) and (32.255) prove (32.249)\(_4\).

We also need for a scalar valued \(f\), for vector fields \({\varvec{u}}, {\varvec{v}}\) and for a second order tensor field \({\varvec{T}}\), defined on the surface \({\mathcal S}\) the following rules of differentiation

$$\begin{aligned} \begin{array}{c} \displaystyle \mathrm {Div}\,(f{\varvec{v}})=f \mathrm {Div}\,{\varvec{v}}+{\mathrm {Grad}}\,f\cdot {\varvec{v}}, \quad \mathrm {Div}\,(f{\varvec{T}})=f \mathrm {Div}\, {{\varvec{T}}} +{{\varvec{T}}}{\mathrm {Grad}}\,f, \\ \displaystyle \mathrm {Div}\,({{\varvec{u}}}\otimes {{\varvec{v}}})=v^{\mathfrak a}\frac{\partial {\varvec{u}}}{\partial \xi ^{\mathfrak a}}+ (\mathrm {Div}\,{{\varvec{v}}})\,{{\varvec{u}}}, \quad \mathrm {Div}\,{\varvec{n}}=-2K,\quad \mathrm {Div}\, ({\varvec{n}}\otimes {\varvec{n}})=-2K{\varvec{n}},\\ \displaystyle \mathrm {Div}\,({\varvec{n}}\otimes {{\varvec{\tau }}}_{\mathfrak a})=-b_{\mathfrak a\mathfrak b}{{\varvec{\tau }}}^{\mathfrak b}+\varGamma ^{\mathfrak b}_{\mathfrak a\mathfrak b}{\varvec{n}},\quad \mathrm {Div}\,( {{\varvec{\tau }}}_{\mathfrak a}\otimes {\varvec{n}})=-2K {{\varvec{\tau }}}_{\mathfrak a}, \end{array} \end{aligned}$$
(32.256)

where

$$\begin{aligned} {\mathrm {Grad}}\,f\equiv \frac{\partial \,f}{\partial \xi ^{\mathfrak a}}\,{{\varvec{\tau }}}^{\mathfrak a}, \quad \mathrm {Div}\, {\varvec{v}}\equiv \frac{\partial {\varvec{v}}}{\partial \xi ^{\mathfrak a}}\cdot {{\varvec{\tau }}}^{\mathfrak a},\quad \mathrm {Div}\,{{\varvec{T}}}\equiv \frac{\partial {{\varvec{T}}}}{\partial \xi ^{\mathfrak a}}\,{{\varvec{\tau }}}^{\mathfrak a}. \end{aligned}$$

Thus, using the decomposition

$$\begin{aligned} {\varvec{v}}_{\mathcal {S}}={\varvec{v}}_{{\mathcal {S}}\parallel }+{\mathcal U}{\varvec{n}}=v^{\mathfrak a}{{\varvec{\tau }}}_{\mathfrak a}+{\mathcal U}{\varvec{n}}, \end{aligned}$$

we obtain

$$\begin{aligned} \begin{array}{l} \displaystyle \frac{\partial }{\partial \,t}(\mu {\varvec{v}}_{\mathcal {S}})\!=\!\frac{\partial \mu v^{\mathfrak a}}{\partial \,t}\,{{\varvec{\tau }}}_{\mathfrak a}+ \mu v^{\mathfrak b}\left\{ \frac{\partial w^{\mathfrak a}}{\partial \xi ^{\mathfrak b}}+w^{\mathfrak c}\varGamma ^{\mathfrak a}_{\mathfrak c\mathfrak b} - {\mathcal U}b_{\mathfrak b\mathfrak c} g^{\mathfrak c\mathfrak a}\right\} {{\varvec{\tau }}}_{\mathfrak a} -\\ \displaystyle \quad \quad \quad \quad \;\; \mu {\mathcal U} g^{\mathfrak a\mathfrak b}\left\{ \frac{\partial \mathcal U}{\partial \xi ^{\mathfrak b}} +b_{\mathfrak b\mathfrak c}w^{\mathfrak c} \right\} {{\varvec{\tau }}}_{\mathfrak a}+ \left\{ \frac{\partial \mu \,\mathcal U}{\partial t}+\mu v^{\mathfrak a}\frac{\partial \mathcal U}{\partial \xi ^{\mathfrak a}} +\mu b_{\mathfrak b\mathfrak a}v^{\mathfrak a} w^{\mathfrak b}\right\} {\varvec{n}}. \end{array} \end{aligned}$$
(32.257)

Then,

$$\begin{aligned}&\quad \mathrm {Div}\,(\mu {\varvec{v}}_{{\mathcal {S}}}\otimes {\varvec{v}}_{{\mathcal {S}}})\nonumber \\&= \mathrm {Div}\,(\mu {\varvec{v}}_{{\mathcal {S}}\parallel }\otimes {\varvec{v}}_{{\mathcal {S}}\parallel })+\mathrm {Div}\,(\mu {\mathcal U}{\varvec{v}}_{{\mathcal {S}}\parallel }\otimes {\varvec{n}})+\mathrm {Div}\,(\mu {\mathcal U}{\varvec{n}}\otimes {\varvec{v}}_{{\mathcal {S}}\parallel })+\mathrm {Div}\,(\mu {\mathcal U}^2{\varvec{n}}\otimes {\varvec{n}})\nonumber \\&=\mathrm {Div}\,(\mu {\varvec{v}}_{{\mathcal {S}}\parallel }\otimes {\varvec{v}}_{{\mathcal {S}}\parallel })-\mu {\mathcal U}\,v^{\mathfrak b} \,b_{\mathfrak b\mathfrak c}\,g^{\mathfrak c\mathfrak a}\,{{\varvec{\tau }}}_{\mathfrak a}+\mathrm {Div}\,(\mu {\mathcal U}{\varvec{v}}_{{\mathcal {S}}\parallel })\,{\varvec{n}}-2\mu K{\mathcal U}{\varvec{v}}_{\mathcal {S}}, \end{aligned}$$
(32.258)

and with the notations (32.181) for the components of \({{\varvec{R}}}_{\mathcal {S}}\),

$$\begin{aligned} \displaystyle \mathrm {Div}{{\varvec{R}}}_{\mathcal {S}}&= \mathrm {Div}\,(S^{\mathfrak a\mathfrak b}{{\varvec{\tau }}}_{\mathfrak a}\otimes {{\varvec{\tau }}}_{\mathfrak b}) - \left\{ S^{\mathfrak c} b_{\mathfrak c\mathfrak b}g^{\mathfrak b\mathfrak a} + 2KS^{\mathfrak a}\right\} {{\varvec{\tau }}}_{\mathfrak a}\nonumber \\&+\left\{ \mathrm {Div}\,(S^{\mathfrak a}{{\varvec{\tau }}}_{\mathfrak a})-2SK\right\} {\varvec{n}}. \end{aligned}$$
(32.259)

Finally, we have

$$\begin{aligned} \displaystyle \frac{\partial }{\partial \xi ^{\mathfrak b}}(\mu {\varvec{v}}_{\mathcal {S}})\,w^{\mathfrak b}&= w^{\mathfrak b}\left\{ \frac{\partial \mu v^{\mathfrak a}}{\partial \xi ^{\mathfrak b}}+ \mu v^{\mathfrak c}\,\varGamma ^{\mathfrak a}_{\mathfrak c\mathfrak b} -\mu {\mathcal U}\,b_{\mathfrak b\mathfrak c} g^{\mathfrak c\mathfrak a} \right\} {{\varvec{\tau }}}_{\mathfrak a}\nonumber \\&+ \displaystyle w^{\mathfrak b}\left\{ \mu v^{\mathfrak c}\,b_{\mathfrak c\mathfrak b}+\frac{\partial \mu \mathcal U}{\partial \xi ^{\mathfrak b}}\right\} {\varvec{n}}. \end{aligned}$$
(32.260)

Now, substituting (32.257)–(32.260) into (32.248) and separating the tangential and normal parts of the emerging relation yields (32.182) and (32.183).

Rights and permissions

Reprints and permissions

Copyright information

© 2014 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Hutter, K., Wang, Y., Chubarenko, I.P. (2014). Sediment Transport in Alluvial Systems. In: Physics of Lakes. Advances in Geophysical and Environmental Mechanics and Mathematics. Springer, Cham. https://doi.org/10.1007/978-3-319-00473-0_32

Download citation

Publish with us

Policies and ethics