You are on page 1of 6

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257405260

Depth-integrated Reynolds-averaged Navier–Stokes equations for shallow


flows over rough permeable beds

Article  in  Journal of Hydraulic Research · September 2013


DOI: 10.1080/00221686.2013.814598

CITATIONS READS

3 403

1 author:

Dubravka Pokrajac
University of Aberdeen
72 PUBLICATIONS   1,409 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Dubravka Pokrajac on 09 November 2015.

The user has requested enhancement of the downloaded file.


This article was downloaded by: [Dubravka Pokrajac]
On: 07 October 2013, At: 12:57
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer
House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Hydraulic Research


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/tjhr20

Depth-integrated Reynolds-averaged Navier–Stokes


equations for shallow flows over rough permeable
beds
a
Dubravka Pokrajac
a
School of Engineering, University of Aberdeen, Aberdeen, UK
Published online: 04 Sep 2013.

To cite this article: Dubravka Pokrajac , Journal of Hydraulic Research (2013): Depth-integrated Reynolds-
averaged Navier–Stokes equations for shallow flows over rough permeable beds, Journal of Hydraulic Research, DOI:
10.1080/00221686.2013.814598

To link to this article: http://dx.doi.org/10.1080/00221686.2013.814598

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of
the Content. Any opinions and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied
upon and should be independently verified with primary sources of information. Taylor and Francis shall
not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other
liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or
arising out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Journal of Hydraulic Research, iFirst, 2013, 1–4
http://dx.doi.org/10.1080/00221686.2013.814598
© 2013 International Association for Hydro-Environment Engineering and Research

Technical note

Depth-integrated Reynolds-averaged Navier–Stokes equations for shallow flows


over rough permeable beds
DUBRAVKA POKRAJAC, Reader, School of Engineering, University of Aberdeen, Aberdeen, UK
Email: d.pokrajac@abdn.ac.uk

ABSTRACT
Depth-integrated Reynolds-averaged Navier–Stokes (DIRANS) equations for free surface flows over permeable rough beds are rigorously derived in
order to clarify the term which quantifies momentum flux through the bed surface.
Downloaded by [Dubravka Pokrajac] at 12:57 07 October 2013

Keywords: DIRANS equations; shallow water equations; surface-subsurface momentum exchange

1 Introduction grains, with the base plane. Its size has to be sufficient to provide
a statistically representative sample of bed geometry, i.e. much
Free surface flows over permeable beds occur in both the natural larger than a typical grain and much smaller than macroscopic
environment, e.g. in gravel bed streams and gravel beaches, and heterogeneities such as bed forms. Geometry of all surfaces that
in man-made structures such as coastal defence structures, porous enclose the control volume is defined by their outer unit normal
pavements, etc. Shallow free surface flows are often described vectors, n. Figure 1 also shows the Cartesian right-handed coor-
using shallow water equations (SWEs). For flows over permeable dinate system, which consists of two horizontal axes, x, y, and a
beds SWEs contain source terms, which account for surface- vertical axis z. The corresponding velocity components are u, v,
subsurface exchange of mass and momentum. There is a certain w, respectively. The thickness of the control volume in y direction
degree of confusion in the literature regarding the source term in is B. For simplicity it is assumed that the volume and momen-
the momentum equation. This note presents a rigorous derivation tum fluxes through the lateral sides of the control volume (y=0
of depth-averaged equations for free-surface flows over a rigid and y = B) are zero, i.e. the depth-averaged equations have one
permeable base. The resulting expression for the source term in principal direction, x. The extension to two-dimensional SWEs
the momentum equation is compared with the literature. is straightforward.
Figure 1 shows the control volume ∀ which will be used for Equations for ensemble-averaged flow are obtained by aver-
the derivation of depth-averaged equations. It extends between aging Navier–Stokes equations over a series of statistical real-
the two vertical sections, S1 and S2 , the free surface, Ss , and the izations, herein called runs. In reality the free surface usually
bed area, Sb . In order to define Sb we first have to specify a “base changes its position between the individual runs. However, in
plane”, i.e. an imaginary plane which defines the lower limit this note it is assumed, for simplicity, that the free surface
of the control volume across the gaps between the grains. The does not change between the individual runs. For the treat-
choice of the base plane does not influence the mathematical ment of the fluctuating free surface in depth-averaged equations
form of the depth-averaged equations, however, depending on the reader is referred to Pokrajac and Kikkert (2011). Further-
this choice the resulting depth-averaged quantities will be more more, pressure distribution along flow cross-sections is assumed
or less physically meaningful. It seems reasonable to adopt the hydrostatic. The extension to non-hydrostatic conditions can be
base plane close to the centres of the top layer of grains, but found in, e.g. Steffler and Jin (1993) and Ghamry and Steffler
other choices such as zero-plane or even grain crests are also (2005).
possible. Further discussion of the optimum position of the base The remaining text shows how the Reynolds-averaged
plane is beyond the scope of this note. The bed surface Sb coin- Navier–Stokes (RANS) equations are integrated over the con-
cides with either a grain surface or, across the gaps between the trol volume shown in Fig. 1 to produce depth-integrated RANS

Revision received 8 August 2012; accepted 10 June 2013/Currently open for discussion.

ISSN 0022-1686 print/ISSN 1814-2079 online


http://www.tandfonline.com
1
2 D. Pokrajac Journal of Hydraulic Research, iFirst (2013)

Figure 1. A control volume is located between two vertical cross-sections, S1 and S2 , free surface, Ss , and the bed surface, Sb . Flow depth in a
cross-section is found by averaging (zs − zb ) over the thickness B (middle panel). Note that the lower limit of the flow depth becomes the reference
bed level for depth-averaged equations and quantities. The bed surface Sb (right panel) follows the grain surfaces wherever they are higher than the
base plane and the area of the gaps between the grains, Sg everywhere else
Downloaded by [Dubravka Pokrajac] at 12:57 07 October 2013

(DIRANS) equations. The note ends with a brief discussion of Note that the lower limit of hi defines a reference bed level for
the momentum exchange term in DIRANS equations. the depth-avereged equations (Fig. 1).
• Free surface
2 DIRANS equations    B  zs2
Qs = ūnx dS + w̄nz dS = − ūs dz dy
2.1 Mass balance Ss Ss 0 zs1
 B  x2
The Reynolds-averaged mass balance equation applied to the + w̄s dx dy (3)
control volume shown in Fig. 1 can be expressed in integral 0 x1

form as:
 where the subscript s means “at the free-surface”. The kine-
(ūnx + w̄nz ) dS = 0 (1) matic condition at the free-surface
S
Dzs ∂zs ∂zs
where the overbar indicates ensemble-averaging, S is the sur- w̄s = = + ūs
Dt ∂t ∂x
face enclosing the control volume (Fig. 1), and nx , nz are the
x, and z components respectively of the outer unit vector n for combined with Eq. (3) gives
the surface S. The term within the brackets in (1) represents the    
B zs2 B x2
velocity in the direction normal to S, i.e. in n direction. ∂zs
Qs = − ūs dz dy + dx dy
Integration of (1) over the parts of S (S1 , S2 , Ss , Sb ) yields the 0 zs1 0 x1 ∂t
expressions for discharges through:  B  x2
∂zs
+ ūs dx dy
0 x1 ∂x
• Flow cross-sections  x2  B  x2
∂zs ∂h
 = dy dx = B dx (4)
∂t ∂t
(ūnx + w̄nz ) dS = −Bh1 U1 , x1 0 x1
S1
 • Bed surface
(ūnx + w̄nz ) dS = Bh2 U2 , (2)  
S2
Qb = (ūnx + w̄nz ) dS = (ūnx + w̄nz ) dSg
Sb Sg
where h1 , h2 , U1 , U2 are the average flow depths and cross- 
sectional velocities for S1 and S2 defined as (Fig. 1): = υg dSg (5)
Sg
 B  zsi  B
1 1
hi = dz dy = (zsi − zbi ) dy, where Sg is the area of gap between the grains (right panel in
B 0 zbi B 0
 B  zsi
Fig. 1), and υg is the velocity component normal to Sg in the
1 direction of the outer normal vector (i.e. positive towards the
Ui = ū dz dy, i = 1, 2
Bhi 0 zbi bed). Note that flow through the gaps may also have a velocity
Journal of Hydraulic Research, iFirst (2013) DIRANS equations for shallow flows over rough permeable beds 3

component within the gap area which does not contribute to shape of velocity profiles across the gaps between the grains,
the discharge. Combining Eqs. (2)–(5) yields εb , the term can be expressed as:
 x2 
∂h
B dx − Bh1 U1 + Bh2 U2 + Qb = 0 (6) ρεb ūb  υg Sg dSg = ρεb ūb Qb ,
x1 ∂t
Sg

1
To express Eq. (6) in differential form volume fluxes in Eq. (6) where ūb  = ūb dSg ,
Sg Sg
are assumed to be sufficiently smooth to allow for conversion of
the integral Eq. (6) into
i.e. the angular brackets indicate spatial averaging over the
∂h ∂hU total area of the gaps, usually called intrinsic averaging.
+ = −q (7)
∂t ∂x
• Pressure
where q = Qb /Bx is the volume flux into the permeable bed
  B  zs1
per unit horizontal projection of the bed area with a “+” sign for
pnx dS = −ρg (zs1 − z) dz dy
flow into the bed. S 0 zb1
 B  zs2
Downloaded by [Dubravka Pokrajac] at 12:57 07 October 2013

+ ρg (zs2 − z) dz dy
2.2 Momentum balance 0 zb2
 
The Reynolds-averaged equation for u-momentum balance in + ρg (zs − zb )nx dS + pF nx dS
integral form can be written as Sb Sb
   x2
h21 h2 ∂zb
  = ρgB − 2 + ρgB h dx + Fx
∂ρ ū 2 2 ∂x
d∀ + ρ ū (ūnx + w̄nz )dS x1
∀ ∂t S
 
= − p̄nx dS + (τxx nx + τxz nz ) dS (8) where g is acceleration due to gravity. Pressure on the bed
S S surface was split into a hydrostatic pressure, equal ρg (zs −
zb ), and the pressure caused by the flow separation from the
where ρ is density, p is pressure, and τxx , τxz are the fluid stress individual roughness elements, pF , which generates the drag
(=viscous stress + turbulent stress) components. The terms in (8) force on the grains, Fx .
are developed as follows.
• Fluid stress
• Local change of momentum within ∀   
(τxx nx + τxz nz ) dS = − τxx dS + τxx dS
  x2  B  zs  S S1 S2
∂ρ ū ∂ρ u   
d∀ = dz dy dx
∀ ∂t x1 0 zb ∂t + τxz nz dS = − τxx dS + τxx dS − Vx
 x2  B   zs  Sb S1 S2
∂ ∂zs
=ρ ūdz − ūs dy dx
x1 0 ∂t zb ∂t
 x2  x2  B where Vx is the viscous drag, i.e. the x-component of the force
∂hU ∂zs
= ρB dx − ρ ūs dy dx exerted by the fluid on the grains by viscous stress (skin fric-
x1 ∂t x1 0 ∂t tion). Note that, as stated earlier, τxx , τxz include both viscous
stress and turbulent stress.
• Momentum flux All above terms are grouped together to produce:


ρ ū(ūnx + w̄nz ) dS = −ρε1 h1 BU12 + ρε2 h2 BU22 x2
∂hU
S ρB dx − ρε1 h1 BU12 + ρε2 h2 BU22
   x1 ∂t
x2 B
∂zs  2   x2 
+ρ ūs dy dx + ρ ūb υg dSg h1 h22 ∂zb
x1 0 ∂t Sg = −ρgB − − ρgB h dx − τxx dS
2 2 x1 ∂x S1

where εi , i = 1, 2 are momentum correction factors that + τxx dS − Tx − ρεb ūb Qb (9)
account for non-uniformity of velocity profiles in the flow S2

cross-sections. The last term on the r.h.s. is the flux of


u-momentum through the bed surface. By introducing a where Tx = Vx + Fx is the x-component of the total force
momentum correction factor that accounts for the effect of the exerted on the grains. For sufficiently smooth momentum
4 D. Pokrajac Journal of Hydraulic Research, iFirst (2013)

fluxes the equation above is converted into through the bed surface multiplied with the average streamwise
velocity at the bed level, or, more precisely, at the level of the
∂hU ∂εhU 2 ∂h ∂zb base plane used for deriving depth-averaged equations. It may be
+ = −gh − gh
∂t ∂x ∂x ∂x necessary to include a momentum correction factor to account
 
∂ ∂U τb for non-uniformity of velocity across the gaps between the grains
+ hνe − − εb ūb q (10)
∂x ∂x ρ at the bed level.

where νe is the effective viscosity commonly used for


References
parametrizing fluid stress, and dividing Tx with the horizontal
bed area has produced the bed shear stress, τb .
Clarke, S., Dodd, N., Damgaard, J. (2004). Modelling flow in
and above a porous beach. J. Waterw. Port Coastal Ocean
Equation (10) shows that the momentum extracted from the
Eng. 130(5), 223–233.
surface flow per unit mass is equal to the average x-component
Ghamry, H.K., Steffler, P.M. (2005). Two-dimensional depth-
of velocity across the base plane, i.e. across the gaps between
averaged modeling of flow in curved open channels.
the grains, and also that there is a coefficient εb to account for
J. Hydraulic Res. 43(1), 44–55.
non-uniformity of this velocity. Some literature, instead of the
Kobayashi, N., Wurjanto, A. (1990). Numerical model for
velocity at the base plane, uses the average profile velocity, U
Downloaded by [Dubravka Pokrajac] at 12:57 07 October 2013

waves on rough permeable slopes. J. Coastal Res. SI(7),


(e.g. Clarke et al. 2004, Li et al. 2002, Kobayashi and Wurjanto
149–166.
1990), which is considerably larger than the velocity at the base
Li, L., Barry, D.A., Pattiaratchi, C.B., Masselink, G. (2002).
plane. Van Gent (1994) uses the x-component of the flux q, which
BeachWin: Modelling groundwater effects on swash sediment
is correct only if the average vertical velocity across the gaps
transport and beach profile changes. Environ. Model. Software
between the grains is zero, i.e. w̄b  = 0. To the author’s best
17(3), 313–320.
knowledge so far the literature has not acknowledged the non-
Pokrajac, D., Kikkert, G. (2011). RADINS equations for aerated
uniformity coefficient εb .
shallow water flows over rough beds. J. Hydraulic Res. 49(5),
630–638.
3 Conclusions Steffler, P.M., Jin, Y.C. (1993). Depth-averaged and moment
equations for moderately shallow free surface flow.
A rigorous derivation of the DIRANS equations for unsteady J. Hydraulic Res. 31(1), 5–18.
turbulent two-dimensional flows over a rigid permeable bed has Van Gent, M.R.A. 1994. The modelling of wave action
been presented. The momentum loss from the surface flow due on and in coastal structures. Coastal Eng. 22(3–4),
to infiltration into a permeable bed is equal to the volume flux 311–339.

View publication stats

You might also like