fbpx
Wikipedia

Shallow water equations

The shallow-water equations (SWE) are a set of hyperbolic partial differential equations (or parabolic if viscous shear is considered) that describe the flow below a pressure surface in a fluid (sometimes, but not necessarily, a free surface).[1] The shallow-water equations in unidirectional form are also called Saint-Venant equations, after Adhémar Jean Claude Barré de Saint-Venant (see the related section below).

Output from a shallow-water equation model of water in a bathtub. The water experiences five splashes which generate surface gravity waves that propagate away from the splash locations and reflect off the bathtub walls.

The equations are derived[2] from depth-integrating the Navier–Stokes equations, in the case where the horizontal length scale is much greater than the vertical length scale. Under this condition, conservation of mass implies that the vertical velocity scale of the fluid is small compared to the horizontal velocity scale. It can be shown from the momentum equation that vertical pressure gradients are nearly hydrostatic, and that horizontal pressure gradients are due to the displacement of the pressure surface, implying that the horizontal velocity field is constant throughout the depth of the fluid. Vertically integrating allows the vertical velocity to be removed from the equations. The shallow-water equations are thus derived.

While a vertical velocity term is not present in the shallow-water equations, note that this velocity is not necessarily zero. This is an important distinction because, for example, the vertical velocity cannot be zero when the floor changes depth, and thus if it were zero only flat floors would be usable with the shallow-water equations. Once a solution (i.e. the horizontal velocities and free surface displacement) has been found, the vertical velocity can be recovered via the continuity equation.

Situations in fluid dynamics where the horizontal length scale is much greater than the vertical length scale are common, so the shallow-water equations are widely applicable. They are used with Coriolis forces in atmospheric and oceanic modeling, as a simplification of the primitive equations of atmospheric flow.

Shallow-water equation models have only one vertical level, so they cannot directly encompass any factor that varies with height. However, in cases where the mean state is sufficiently simple, the vertical variations can be separated from the horizontal and several sets of shallow-water equations can describe the state.

Equations edit

 
A one-dimensional diagram representing the shallow water model.

Conservative form edit

The shallow-water equations are derived from equations of conservation of mass and conservation of linear momentum (the Navier–Stokes equations), which hold even when the assumptions of shallow-water break down, such as across a hydraulic jump. In the case of a horizontal bed, with negligible Coriolis forces, frictional and viscous forces, the shallow-water equations are:

 

Here η is the total fluid column height (instantaneous fluid depth as a function of x, y and t), and the 2D vector (u,v) is the fluid's horizontal flow velocity, averaged across the vertical column. Further g is acceleration due to gravity and ρ is the fluid density. The first equation is derived from mass conservation, the second two from momentum conservation.[3]

Non-conservative form edit

Expanding the derivatives in the above using the product rule, the non-conservative form of the shallow-water equations is obtained. Since velocities are not subject to a fundamental conservation equation, the non-conservative forms do not hold across a shock or hydraulic jump. Also included are the appropriate terms for Coriolis, frictional and viscous forces, to obtain (for constant fluid density):

 

where

u is the velocity in the x direction, or zonal velocity
v is the velocity in the y direction, or meridional velocity
H is the mean height of the horizontal pressure surface
h is the height deviation of the horizontal pressure surface from its mean height, where h: η(x, y, t) = H(x, y) + h(x, y, t)
b is the topographical height from a reference D, where b: H(x, y) = D + b(x,y)
g is the acceleration due to gravity
f is the Coriolis coefficient associated with the Coriolis force. On Earth, f is equal to 2Ω sin(φ), where Ω is the angular rotation rate of the Earth (π/12 radians/hour), and φ is the latitude
k is the viscous drag coefficient
ν is the kinematic viscosity
Animation of the linearized shallow-water equations for a rectangular basin, without friction and Coriolis force. The water experiences a splash which generates surface gravity waves that propagate away from the splash location and reflect off the basin walls. The animation is created using the exact solution of Carrier and Yeh (2005) for axisymmetrical waves.[4]

It is often the case that the terms quadratic in u and v, which represent the effect of bulk advection, are small compared to the other terms. This is called geostrophic balance, and is equivalent to saying that the Rossby number is small. Assuming also that the wave height is very small compared to the mean height (hH), we have (without lateral viscous forces):

 

One-dimensional Saint-Venant equations edit

The one-dimensional (1-D) Saint-Venant equations were derived by Adhémar Jean Claude Barré de Saint-Venant, and are commonly used to model transient open-channel flow and surface runoff. They can be viewed as a contraction of the two-dimensional (2-D) shallow-water equations, which are also known as the two-dimensional Saint-Venant equations. The 1-D Saint-Venant equations contain to a certain extent the main characteristics of the channel cross-sectional shape.

The 1-D equations are used extensively in computer models such as TUFLOW, Mascaret (EDF), SIC (Irstea), HEC-RAS,[5] SWMM5, ISIS,[5] InfoWorks,[5] Flood Modeller, SOBEK 1DFlow, MIKE 11,[5] and MIKE SHE because they are significantly easier to solve than the full shallow-water equations. Common applications of the 1-D Saint-Venant equations include flood routing along rivers (including evaluation of measures to reduce the risks of flooding), dam break analysis, storm pulses in an open channel, as well as storm runoff in overland flow.

Equations edit

 
Cross section of the open channel.

The system of partial differential equations which describe the 1-D incompressible flow in an open channel of arbitrary cross section – as derived and posed by Saint-Venant in his 1871 paper (equations 19 & 20) – is:[6]

 

(1)

and

 

(2)

where x is the space coordinate along the channel axis, t denotes time, A(x,t) is the cross-sectional area of the flow at location x, u(x,t) is the flow velocity, ζ(x,t) is the free surface elevation and τ(x,t) is the wall shear stress along the wetted perimeter P(x,t) of the cross section at x. Further ρ is the (constant) fluid density and g is the gravitational acceleration.

Closure of the hyperbolic system of equations (1)–(2) is obtained from the geometry of cross sections – by providing a functional relationship between the cross-sectional area A and the surface elevation ζ at each position x. For example, for a rectangular cross section, with constant channel width B and channel bed elevation zb, the cross sectional area is: A = B (ζ − zb) = B h. The instantaneous water depth is h(x,t) = ζ(x,t) − zb(x), with zb(x) the bed level (i.e. elevation of the lowest point in the bed above datum, see the cross-section figure). For non-moving channel walls the cross-sectional area A in equation (1) can be written as:

 
with b(x,h) the effective width of the channel cross section at location x when the fluid depth is h – so b(x, h) = B(x) for rectangular channels.[7]

The wall shear stress τ is dependent on the flow velocity u, they can be related by using e.g. the Darcy–Weisbach equation, Manning formula or Chézy formula.

Further, equation (1) is the continuity equation, expressing conservation of water volume for this incompressible homogeneous fluid. Equation (2) is the momentum equation, giving the balance between forces and momentum change rates.

The bed slope S(x), friction slope Sf(x, t) and hydraulic radius R(x, t) are defined as:

 
 
and
 

Consequently, the momentum equation (2) can be written as:[7]

 

(3)

Conservation of momentum edit

The momentum equation (3) can also be cast in the so-called conservation form, through some algebraic manipulations on the Saint-Venant equations, (1) and (3). In terms of the discharge Q = Au:[8]

 

(4)

where A, I1 and I2 are functions of the channel geometry, described in the terms of the channel width B(σ,x). Here σ is the height above the lowest point in the cross section at location x, see the cross-section figure. So σ is the height above the bed level zb(x) (of the lowest point in the cross section):

 

Above – in the momentum equation (4) in conservation form – A, I1 and I2 are evaluated at σ = h(x,t). The term g I1 describes the hydrostatic force in a certain cross section. And, for a non-prismatic channel, g I2 gives the effects of geometry variations along the channel axis x.

In applications, depending on the problem at hand, there often is a preference for using either the momentum equation in non-conservation form, (2) or (3), or the conservation form (4). For instance in case of the description of hydraulic jumps, the conservation form is preferred since the momentum flux is continuous across the jump.

Characteristics edit

 
Characteristics, domain of dependence and region of influence, associated with location P = (xP,tP) in space x and time t.

The Saint-Venant equations (1)–(2) can be analysed using the method of characteristics.[9][10][11][12] The two celerities dx/dt on the characteristic curves are:[8]

 
with
 

The Froude number Fr = |u| / c determines whether the flow is subcritical (Fr < 1) or supercritical (Fr > 1).

For a rectangular and prismatic channel of constant width B, i.e. with A = B h and c = gh, the Riemann invariants are:[9]

 
and
 
so the equations in characteristic form are:[9]
 

The Riemann invariants and method of characteristics for a prismatic channel of arbitrary cross-section are described by Didenkulova & Pelinovsky (2011).[12]

The characteristics and Riemann invariants provide important information on the behavior of the flow, as well as that they may be used in the process of obtaining (analytical or numerical) solutions.[13][14][15][16]

Hamiltonian structure for frictionless flow edit

In case there is no friction and the channel has a rectangular prismatic cross section, the Saint-Venant equations have a Hamiltonian structure.[17] The Hamiltonian H is equal to the energy of the free-surface flow:

 
with constant B the channel width and ρ the constant fluid density. Hamilton's equations then are:
 
since A/∂ζ = B).

Derived modelling edit

Dynamic wave edit

The dynamic wave is the full one-dimensional Saint-Venant equation. It is numerically challenging to solve, but is valid for all channel flow scenarios. The dynamic wave is used for modeling transient storms in modeling programs including Mascaret (EDF), SIC (Irstea), HEC-RAS,[18] InfoWorks_ICM 2016-10-25 at the Wayback Machine,[19] MIKE 11,[20] Wash 123d[21] and SWMM5.

In the order of increasing simplifications, by removing some terms of the full 1D Saint-Venant equations (aka Dynamic wave equation), we get the also classical Diffusive wave equation and Kinematic wave equation.

Diffusive wave edit

For the diffusive wave it is assumed that the inertial terms are less than the gravity, friction, and pressure terms. The diffusive wave can therefore be more accurately described as a non-inertia wave, and is written as:

 

The diffusive wave is valid when the inertial acceleration is much smaller than all other forms of acceleration, or in other words when there is primarily subcritical flow, with low Froude values. Models that use the diffusive wave assumption include MIKE SHE[22] and LISFLOOD-FP.[23] In the SIC (Irstea) software this options is also available, since the 2 inertia terms (or any of them) can be removed in option from the interface.

Kinematic wave edit

For the kinematic wave it is assumed that the flow is uniform, and that the friction slope is approximately equal to the slope of the channel. This simplifies the full Saint-Venant equation to the kinematic wave:

 

The kinematic wave is valid when the change in wave height over distance and velocity over distance and time is negligible relative to the bed slope, e.g. for shallow flows over steep slopes.[24] The kinematic wave is used in HEC-HMS.[25]

Derivation from Navier–Stokes equations edit

The 1-D Saint-Venant momentum equation can be derived from the Navier–Stokes equations that describe fluid motion. The x-component of the Navier–Stokes equations – when expressed in Cartesian coordinates in the x-direction – can be written as:

 

where u is the velocity in the x-direction, v is the velocity in the y-direction, w is the velocity in the z-direction, t is time, p is the pressure, ρ is the density of water, ν is the kinematic viscosity, and fx is the body force in the x-direction.

  1. If it is assumed that friction is taken into account as a body force, then   can be assumed as zero so:
     
  2. Assuming one-dimensional flow in the x-direction it follows that:[26]
     
  3. Assuming also that the pressure distribution is approximately hydrostatic it follows that:[26]
     
    or in differential form:
     
    And when these assumptions are applied to the x-component of the Navier–Stokes equations:
     
  4. There are 2 body forces acting on the channel fluid, namely, gravity and friction:
     
    where fx,g is the body force due to gravity and fx,f is the body force due to friction.
  5. fx,g can be calculated using basic physics and trigonometry:[27]
     
    where Fg is the force of gravity in the x-direction, θ is the angle, and M is the mass.
     
    Figure 1: Diagram of block moving down an inclined plane.
    The expression for sin θ can be simplified using trigonometry as:
     
    For small θ (reasonable for almost all streams) it can be assumed that:
     
    and given that fx represents a force per unit mass, the expression becomes:
     
  6. Assuming the energy grade line is not the same as the channel slope, and for a reach of consistent slope there is a consistent friction loss, it follows that:[28]
     
  7. All of these assumptions combined arrives at the 1-dimensional Saint-Venant equation in the x-direction:
     
     
    where (a) is the local acceleration term, (b) is the convective acceleration term, (c) is the pressure gradient term, (d) is the friction term, and (e) is the gravity term.
Terms

The local acceleration (a) can also be thought of as the "unsteady term" as this describes some change in velocity over time. The convective acceleration (b) is an acceleration caused by some change in velocity over position, for example the speeding up or slowing down of a fluid entering a constriction or an opening, respectively. Both these terms make up the inertia terms of the 1-dimensional Saint-Venant equation.

The pressure gradient term (c) describes how pressure changes with position, and since the pressure is assumed hydrostatic, this is the change in head over position. The friction term (d) accounts for losses in energy due to friction, while the gravity term (e) is the acceleration due to bed slope.

Wave modelling by shallow-water equations edit

Shallow-water equations can be used to model Rossby and Kelvin waves in the atmosphere, rivers, lakes and oceans as well as gravity waves in a smaller domain (e.g. surface waves in a bath). In order for shallow-water equations to be valid, the wavelength of the phenomenon they are supposed to model has to be much larger than the depth of the basin where the phenomenon takes place. Somewhat smaller wavelengths can be handled by extending the shallow-water equations using the Boussinesq approximation to incorporate dispersion effects.[29] Shallow-water equations are especially suitable to model tides which have very large length scales (over hundred of kilometers). For tidal motion, even a very deep ocean may be considered as shallow as its depth will always be much smaller than the tidal wavelength.

 
Tsunami generation and propagation, as computed with the shallow-water equations (red line; without frequency dispersion)), and with a Boussinesq-type model (blue line; with frequency dispersion). Observe that the Boussinesq-type model (blue line) forms a soliton with an oscillatory tail staying behind. The shallow-water equations (red line) form a steep front, which will lead to bore formation, later on. The water depth is 100 meters.

Turbulence modelling using non-linear shallow-water equations edit

 
A snapshot from simulation of shallow-water equations in which shock waves are present

Shallow-water equations, in its non-linear form, is an obvious candidate for modelling turbulence in the atmosphere and oceans, i.e. geophysical turbulence. An advantage of this, over Quasi-geostrophic equations, is that it allows solutions like gravity waves, while also conserving energy and potential vorticity. However, there are also some disadvantages as far as geophysical applications are concerned - it has a non-quadratic expression for total energy and a tendency for waves to become shock waves.[30] Some alternate models have been proposed which prevent shock formation. One alternative is to modify the "pressure term" in the momentum equation, but it results in a complicated expression for kinetic energy.[31] Another option is to modify the non-linear terms in all equations, which gives a quadratic expression for kinetic energy, avoids shock formation, but conserves only linearized potential vorticity.[32]

See also edit

Notes edit

  1. ^ Vreugdenhil, C.B. (1986). Numerical Methods for Shallow-Water Flow. Water Science and Technology Library. Vol. 13. Springer, Dordrecht. p. 262. doi:10.1007/978-94-015-8354-1. ISBN 978-90-481-4472-3.
  2. ^ (PDF). Archived from the original (PDF) on 2012-03-16. Retrieved 2010-01-22.
  3. ^ Clint Dawson and Christopher M. Mirabito (2008). "The Shallow Water Equations" (PDF). Retrieved 2013-03-28.
  4. ^ Carrier, G. F.; Yeh, H. (2005), "Tsunami propagation from a finite source", Computer Modeling in Engineering & Sciences, 10 (2): 113–122, doi:10.3970/cmes.2005.010.113
  5. ^ a b c d S. Néelz; G Pender (2009). . Joint Environment Agency/Defra Flood and Coastal Erosion Risk Management Research and Development Programme (Science Report: SC080035): 5. Archived from the original on 8 September 2019. Retrieved 2 December 2016.
  6. ^ Saint-Venant, A.J.C. Barré de (1871), "Théorie du mouvement non permanent des eaux, avec application aux crues des rivières et a l'introduction de marées dans leurs lits", Comptes Rendus de l'Académie des Sciences, 73: 147–154 and 237–240
  7. ^ a b Chow, Ven Te (1959), Open-channel hydraulics, McGraw-Hill, OCLC 4010975, §18-1 & §18-2.
  8. ^ a b Cunge, J. A., F. M. Holly Jr. and A. Verwey (1980), Practical aspects of computational river hydraulics, Pitman Publishing, ISBN 0 273 08442 9, §§2.1 & 2.2
  9. ^ a b c Whitham, G. B. (1974) Linear and Nonlinear Waves, §§5.2 & 13.10, Wiley, ISBN 0-471-94090-9
  10. ^ Lighthill, J. (2005), Waves in fluids, Cambridge University Press, ISBN 978-0-521-01045-0, §§2.8–2.14
  11. ^ Meyer, R. E. (1960), Theory of characteristics of inviscid gas dynamics. In: Fluid Dynamics/Strömungsmechanik, Encyclopedia of Physics IX, Eds. S. Flügge & C. Truesdell, Springer, Berlin, ISBN 978-3-642-45946-7, pp. 225–282
  12. ^ a b Didenkulova, I.; Pelinovsky, E. (2011). "Rogue waves in nonlinear hyperbolic systems (shallow-water framework)". Nonlinearity. 24 (3): R1–R18. Bibcode:2011Nonli..24R...1D. doi:10.1088/0951-7715/24/3/R01. S2CID 59438883.
  13. ^ Harris, M. W.; Nicolsky, D. J.; Pelinovsky, E. N.; Rybkin, A. V. (2015-03-01). "Runup of Nonlinear Long Waves in Trapezoidal Bays: 1-D Analytical Theory and 2-D Numerical Computations". Pure and Applied Geophysics. 172 (3–4): 885–899. Bibcode:2015PApGe.172..885H. doi:10.1007/s00024-014-1016-3. ISSN 0033-4553. S2CID 55004099.
  14. ^ Harris, M. W.; Nicolsky, D. J.; Pelinovsky, E. N.; Pender, J. M.; Rybkin, A. V. (2016-05-01). "Run-up of nonlinear long waves in U-shaped bays of finite length: analytical theory and numerical computations". Journal of Ocean Engineering and Marine Energy. 2 (2): 113–127. doi:10.1007/s40722-015-0040-4. ISSN 2198-6444. S2CID 123725815.
  15. ^ Garayshin, V. V.; Harris, M. W.; Nicolsky, D. J.; Pelinovsky, E. N.; Rybkin, A. V. (2016-04-10). "An analytical and numerical study of long wave run-up in U-shaped and V-shaped bays". Applied Mathematics and Computation. 279: 187–197. doi:10.1016/j.amc.2016.01.005.
  16. ^ Anderson, Dalton; Harris, Matthew; Hartle, Harrison; Nicolsky, Dmitry; Pelinovsky, Efim; Raz, Amir; Rybkin, Alexei (2017-02-02). "Run-Up of Long Waves in Piecewise Sloping U-Shaped Bays". Pure and Applied Geophysics. 174 (8): 3185. Bibcode:2017PApGe.174.3185A. doi:10.1007/s00024-017-1476-3. ISSN 0033-4553. S2CID 132114728.
  17. ^ Lannes, D. (2013). The Water Waves Problem: Mathematical Analysis and Asymptotics. Mathematical Surveys and Monographs. American Mathematical Society. p. 174. ISBN 9780821894705. LCCN 2012046540.
  18. ^ Brunner, G. W. (1995), HEC-RAS River Analysis System. Hydraulic Reference Manual. Version 1.0 Rep., DTIC Document.
  19. ^ Searby, D.; Dean, A.; Margetts J. (1998), Christchurch harbour Hydroworks modelling., Proceedings of the WAPUG Autumn meeting, Blackpool, UK.
  20. ^ Havnø, K., M. Madsen, J. Dørge, and V. Singh (1995), MIKE 11-a generalized river modelling package, Computer models of watershed hydrology., 733–782.
  21. ^ Yeh, G.; Cheng, J.; Lin, J.; Martin, W. (1995), A numerical model simulating water flow and contaminant and sediment transport in watershed systems of 1-D stream-river network, 2-D overland regime, and 3-D subsurface media . Computer models of watershed hydrology, 733–782.
  22. ^ DHI (Danish Hydraulic Institute) (2011), MIKE SHE User Manual Volume 2: Reference Guide, edited.
  23. ^ Bates, P., T. Fewtrell, M. Trigg, and J. Neal (2008), LISFLOOD-FP user manual and technical note, code release 4.3. 6, University of Bristol.
  24. ^ Novak, P., et al., Hydraulic Modelling – An Introduction: Principles, Methods and Applications. 2010: CRC Press.
  25. ^ Scharffenberg, W. A., and M. J. Fleming (2006), Hydrologic Modeling System HEC-HMS: User's Manual, US Army Corps of Engineers, Hydrologic Engineering Center.
  26. ^ a b Vincent., Fromion (2009). Modeling and control of hydrosystems. Springer. ISBN 9781848826243. OCLC 401159458.
  27. ^ "Inclined Planes". www.physicsclassroom.com. Retrieved 2017-05-16.
  28. ^ Methods., Haestad (2007). Computer applications in hydraulic engineering : connecting theory to practice. Bentley Institute Press. ISBN 978-0971414167. OCLC 636350249.
  29. ^ Dingemans, M.W. (1997), Wave propagation over uneven bottoms, Advanced Series on Ocean Engineering 13, World Scientific, Singapore, pp. 473 & 516, ISBN 978-981-02-0427-3
  30. ^ Augier, Pierre; Mohanan, Ashwin Vishnu; Lindborg, Erik (2019-09-17). "Shallow water wave turbulence". Journal of Fluid Mechanics. 874: 1169–1196. Bibcode:2019JFM...874.1169A. doi:10.1017/jfm.2019.375. ISSN 1469-7645. S2CID 198976015.
  31. ^ Bühler, Oliver (1998-09-01). "A Shallow-Water Model that Prevents Nonlinear Steepening of Gravity Waves". Journal of the Atmospheric Sciences. 55 (17): 2884–2891. Bibcode:1998JAtS...55.2884B. doi:10.1175/1520-0469(1998)055<2884:ASWMTP>2.0.CO;2. ISSN 0022-4928.
  32. ^ Lindborg, Erik; Mohanan, Ashwin Vishnu (2017-11-01). "A two-dimensional toy model for geophysical turbulence". Physics of Fluids. 29 (11): 111114. Bibcode:2017PhFl...29k1114L. doi:10.1063/1.4985990. ISSN 1070-6631.

Further reading edit

External links edit

  • Derivation of the shallow-water equations from first principles (instead of simplifying the Navier–Stokes equations, some analytical solutions)

shallow, water, equations, shallow, water, equations, hyperbolic, partial, differential, equations, parabolic, viscous, shear, considered, that, describe, flow, below, pressure, surface, fluid, sometimes, necessarily, free, surface, shallow, water, equations, . The shallow water equations SWE are a set of hyperbolic partial differential equations or parabolic if viscous shear is considered that describe the flow below a pressure surface in a fluid sometimes but not necessarily a free surface 1 The shallow water equations in unidirectional form are also called Saint Venant equations after Adhemar Jean Claude Barre de Saint Venant see the related section below Output from a shallow water equation model of water in a bathtub The water experiences five splashes which generate surface gravity waves that propagate away from the splash locations and reflect off the bathtub walls The equations are derived 2 from depth integrating the Navier Stokes equations in the case where the horizontal length scale is much greater than the vertical length scale Under this condition conservation of mass implies that the vertical velocity scale of the fluid is small compared to the horizontal velocity scale It can be shown from the momentum equation that vertical pressure gradients are nearly hydrostatic and that horizontal pressure gradients are due to the displacement of the pressure surface implying that the horizontal velocity field is constant throughout the depth of the fluid Vertically integrating allows the vertical velocity to be removed from the equations The shallow water equations are thus derived While a vertical velocity term is not present in the shallow water equations note that this velocity is not necessarily zero This is an important distinction because for example the vertical velocity cannot be zero when the floor changes depth and thus if it were zero only flat floors would be usable with the shallow water equations Once a solution i e the horizontal velocities and free surface displacement has been found the vertical velocity can be recovered via the continuity equation Situations in fluid dynamics where the horizontal length scale is much greater than the vertical length scale are common so the shallow water equations are widely applicable They are used with Coriolis forces in atmospheric and oceanic modeling as a simplification of the primitive equations of atmospheric flow Shallow water equation models have only one vertical level so they cannot directly encompass any factor that varies with height However in cases where the mean state is sufficiently simple the vertical variations can be separated from the horizontal and several sets of shallow water equations can describe the state Contents 1 Equations 1 1 Conservative form 1 2 Non conservative form 2 One dimensional Saint Venant equations 2 1 Equations 2 2 Conservation of momentum 2 3 Characteristics 2 4 Hamiltonian structure for frictionless flow 2 5 Derived modelling 2 5 1 Dynamic wave 2 5 2 Diffusive wave 2 5 3 Kinematic wave 2 6 Derivation from Navier Stokes equations 3 Wave modelling by shallow water equations 4 Turbulence modelling using non linear shallow water equations 5 See also 6 Notes 7 Further reading 8 External linksEquations edit nbsp A one dimensional diagram representing the shallow water model Conservative form edit The shallow water equations are derived from equations of conservation of mass and conservation of linear momentum the Navier Stokes equations which hold even when the assumptions of shallow water break down such as across a hydraulic jump In the case of a horizontal bed with negligible Coriolis forces frictional and viscous forces the shallow water equations are r h t r h u x r h v y 0 r h u t x r h u 2 1 2 r g h 2 r h u v y 0 r h v t y r h v 2 1 2 r g h 2 r h u v x 0 displaystyle begin aligned frac partial rho eta partial t amp frac partial rho eta u partial x frac partial rho eta v partial y 0 3pt frac partial rho eta u partial t amp frac partial partial x left rho eta u 2 frac 1 2 rho g eta 2 right frac partial rho eta uv partial y 0 3pt frac partial rho eta v partial t amp frac partial partial y left rho eta v 2 frac 1 2 rho g eta 2 right frac partial rho eta uv partial x 0 end aligned nbsp Here h is the total fluid column height instantaneous fluid depth as a function of x y and t and the 2D vector u v is the fluid s horizontal flow velocity averaged across the vertical column Further g is acceleration due to gravity and r is the fluid density The first equation is derived from mass conservation the second two from momentum conservation 3 Non conservative form edit Expanding the derivatives in the above using the product rule the non conservative form of the shallow water equations is obtained Since velocities are not subject to a fundamental conservation equation the non conservative forms do not hold across a shock or hydraulic jump Also included are the appropriate terms for Coriolis frictional and viscous forces to obtain for constant fluid density h t x H h u y H h v 0 u t u u x v u y f v g h x k u n 2 u x 2 2 u y 2 v t u v x v v y f u g h y k v n 2 v x 2 2 v y 2 displaystyle begin aligned frac partial h partial t amp frac partial partial x Bigl H h u Bigr frac partial partial y Bigl H h v Bigr 0 3pt frac partial u partial t amp u frac partial u partial x v frac partial u partial y fv g frac partial h partial x ku nu left frac partial 2 u partial x 2 frac partial 2 u partial y 2 right 3pt frac partial v partial t amp u frac partial v partial x v frac partial v partial y fu g frac partial h partial y kv nu left frac partial 2 v partial x 2 frac partial 2 v partial y 2 right end aligned nbsp where u is the velocity in the x direction or zonal velocity v is the velocity in the y direction or meridional velocity H is the mean height of the horizontal pressure surface h is the height deviation of the horizontal pressure surface from its mean height where h h x y t H x y h x y t b is the topographical height from a reference D where b H x y D b x y g is the acceleration due to gravity f is the Coriolis coefficient associated with the Coriolis force On Earth f is equal to 2W sin f where W is the angular rotation rate of the Earth p 12 radians hour and f is the latitude k is the viscous drag coefficient n is the kinematic viscosity source source source source Animation of the linearized shallow water equations for a rectangular basin without friction and Coriolis force The water experiences a splash which generates surface gravity waves that propagate away from the splash location and reflect off the basin walls The animation is created using the exact solution of Carrier and Yeh 2005 for axisymmetrical waves 4 It is often the case that the terms quadratic in u and v which represent the effect of bulk advection are small compared to the other terms This is called geostrophic balance and is equivalent to saying that the Rossby number is small Assuming also that the wave height is very small compared to the mean height h H we have without lateral viscous forces h t H u x v y 0 u t f v g h x k u v t f u g h y k v displaystyle begin aligned frac partial h partial t amp H left frac partial u partial x frac partial v partial y right 0 3pt frac partial u partial t amp fv g frac partial h partial x ku 3pt frac partial v partial t amp fu g frac partial h partial y kv end aligned nbsp One dimensional Saint Venant equations editThe one dimensional 1 D Saint Venant equations were derived by Adhemar Jean Claude Barre de Saint Venant and are commonly used to model transient open channel flow and surface runoff They can be viewed as a contraction of the two dimensional 2 D shallow water equations which are also known as the two dimensional Saint Venant equations The 1 D Saint Venant equations contain to a certain extent the main characteristics of the channel cross sectional shape The 1 D equations are used extensively in computer models such as TUFLOW Mascaret EDF SIC Irstea HEC RAS 5 SWMM5 ISIS 5 InfoWorks 5 Flood Modeller SOBEK 1DFlow MIKE 11 5 and MIKE SHE because they are significantly easier to solve than the full shallow water equations Common applications of the 1 D Saint Venant equations include flood routing along rivers including evaluation of measures to reduce the risks of flooding dam break analysis storm pulses in an open channel as well as storm runoff in overland flow Equations edit nbsp Cross section of the open channel The system of partial differential equations which describe the 1 D incompressible flow in an open channel of arbitrary cross section as derived and posed by Saint Venant in his 1871 paper equations 19 amp 20 is 6 A t A u x 0 displaystyle frac partial A partial t frac partial left Au right partial x 0 nbsp 1 and u t u u x g z x P A t r displaystyle frac partial u partial t u frac partial u partial x g frac partial zeta partial x frac P A frac tau rho nbsp 2 where x is the space coordinate along the channel axis t denotes time A x t is the cross sectional area of the flow at location x u x t is the flow velocity z x t is the free surface elevation and t x t is the wall shear stress along the wetted perimeter P x t of the cross section at x Further r is the constant fluid density and g is the gravitational acceleration Closure of the hyperbolic system of equations 1 2 is obtained from the geometry of cross sections by providing a functional relationship between the cross sectional area A and the surface elevation z at each position x For example for a rectangular cross section with constant channel width B and channel bed elevation zb the cross sectional area is A B z zb B h The instantaneous water depth is h x t z x t zb x with zb x the bed level i e elevation of the lowest point in the bed above datum see the cross section figure For non moving channel walls the cross sectional area A in equation 1 can be written as A x t 0 h x t b x h d h displaystyle A x t int 0 h x t b x h dh nbsp with b x h the effective width of the channel cross section at location x when the fluid depth is h so b x h B x for rectangular channels 7 The wall shear stress t is dependent on the flow velocity u they can be related by using e g the Darcy Weisbach equation Manning formula or Chezy formula Further equation 1 is the continuity equation expressing conservation of water volume for this incompressible homogeneous fluid Equation 2 is the momentum equation giving the balance between forces and momentum change rates The bed slope S x friction slope Sf x t and hydraulic radius R x t are defined as S d z b d x displaystyle S frac mathrm d z mathrm b mathrm d x nbsp S f t r g R displaystyle S mathrm f frac tau rho gR nbsp and R A P displaystyle R frac A P nbsp Consequently the momentum equation 2 can be written as 7 u t u u x g h x g S f S 0 displaystyle frac partial u partial t u frac partial u partial x g frac partial h partial x g left S mathrm f S right 0 nbsp 3 Conservation of momentum edit The momentum equation 3 can also be cast in the so called conservation form through some algebraic manipulations on the Saint Venant equations 1 and 3 In terms of the discharge Q Au 8 Q t x Q 2 A g I 1 g A S f S g I 2 0 displaystyle frac partial Q partial t frac partial partial x left frac Q 2 A g I 1 right g A left S f S right g I 2 0 nbsp 4 where A I1 and I2 are functions of the channel geometry described in the terms of the channel width B s x Here s is the height above the lowest point in the cross section at location x see the cross section figure So s is the height above the bed level zb x of the lowest point in the cross section A s x 0 s B s x d s I 1 s x 0 s s s B s x d s and I 2 s x 0 s s s B s x x d s displaystyle begin aligned A sigma x amp int 0 sigma B sigma x mathrm d sigma I 1 sigma x amp int 0 sigma sigma sigma B sigma prime x mathrm d sigma qquad text and I 2 sigma x amp int 0 sigma sigma sigma frac partial B sigma x partial x mathrm d sigma end aligned nbsp Above in the momentum equation 4 in conservation form A I1 and I2 are evaluated at s h x t The term g I1 describes the hydrostatic force in a certain cross section And for a non prismatic channel g I2 gives the effects of geometry variations along the channel axis x In applications depending on the problem at hand there often is a preference for using either the momentum equation in non conservation form 2 or 3 or the conservation form 4 For instance in case of the description of hydraulic jumps the conservation form is preferred since the momentum flux is continuous across the jump Characteristics edit nbsp Characteristics domain of dependence and region of influence associated with location P xP tP in space x and time t The Saint Venant equations 1 2 can be analysed using the method of characteristics 9 10 11 12 The two celerities dx dt on the characteristic curves are 8 d x d t u c displaystyle frac mathrm d x mathrm d t u pm c nbsp with c g A B displaystyle c sqrt frac gA B nbsp The Froude number Fr u c determines whether the flow is subcritical Fr lt 1 or supercritical Fr gt 1 For a rectangular and prismatic channel of constant width B i e with A B h and c gh the Riemann invariants are 9 r u 2 g h displaystyle r u 2 sqrt gh nbsp and r u 2 g h displaystyle r u 2 sqrt gh nbsp so the equations in characteristic form are 9 d d t u 2 g h g S S f along d x d t u g h and d d t u 2 g h g S S f along d x d t u g h displaystyle begin aligned amp frac mathrm d mathrm d t left u 2 sqrt gh right g left S S f right amp amp text along quad frac mathrm d x mathrm d t u sqrt gh quad text and amp frac mathrm d mathrm d t left u 2 sqrt gh right g left S S f right amp amp text along quad frac mathrm d x mathrm d t u sqrt gh end aligned nbsp The Riemann invariants and method of characteristics for a prismatic channel of arbitrary cross section are described by Didenkulova amp Pelinovsky 2011 12 The characteristics and Riemann invariants provide important information on the behavior of the flow as well as that they may be used in the process of obtaining analytical or numerical solutions 13 14 15 16 Hamiltonian structure for frictionless flow edit In case there is no friction and the channel has a rectangular prismatic cross section the Saint Venant equations have a Hamiltonian structure 17 The Hamiltonian H is equal to the energy of the free surface flow H r 1 2 A u 2 1 2 g B z 2 d x displaystyle H rho int left frac 1 2 Au 2 frac 1 2 gB zeta 2 right mathrm d x nbsp with constant B the channel width and r the constant fluid density Hamilton s equations then are r B z t x H u r B z t A u x r A t A u x 0 r B u t x H z r B u t u u x g z x 0 displaystyle begin aligned amp rho B frac partial zeta partial t frac partial partial x left frac partial H partial u right rho left B frac partial zeta partial t frac partial Au partial x right rho left frac partial A partial t frac partial Au partial x right 0 amp rho B frac partial u partial t frac partial partial x left frac partial H partial zeta right rho B left frac partial u partial t u frac partial u partial x g frac partial zeta partial x right 0 end aligned nbsp since A z B Derived modelling edit Dynamic wave edit The dynamic wave is the full one dimensional Saint Venant equation It is numerically challenging to solve but is valid for all channel flow scenarios The dynamic wave is used for modeling transient storms in modeling programs including Mascaret EDF SIC Irstea HEC RAS 18 InfoWorks ICM Archived 2016 10 25 at the Wayback Machine 19 MIKE 11 20 Wash 123d 21 and SWMM5 In the order of increasing simplifications by removing some terms of the full 1D Saint Venant equations aka Dynamic wave equation we get the also classical Diffusive wave equation and Kinematic wave equation Diffusive wave edit For the diffusive wave it is assumed that the inertial terms are less than the gravity friction and pressure terms The diffusive wave can therefore be more accurately described as a non inertia wave and is written as g h x g S f S 0 displaystyle g frac partial h partial x g S f S 0 nbsp The diffusive wave is valid when the inertial acceleration is much smaller than all other forms of acceleration or in other words when there is primarily subcritical flow with low Froude values Models that use the diffusive wave assumption include MIKE SHE 22 and LISFLOOD FP 23 In the SIC Irstea software this options is also available since the 2 inertia terms or any of them can be removed in option from the interface Kinematic wave edit For the kinematic wave it is assumed that the flow is uniform and that the friction slope is approximately equal to the slope of the channel This simplifies the full Saint Venant equation to the kinematic wave S f S 0 displaystyle S f S 0 nbsp The kinematic wave is valid when the change in wave height over distance and velocity over distance and time is negligible relative to the bed slope e g for shallow flows over steep slopes 24 The kinematic wave is used in HEC HMS 25 Derivation from Navier Stokes equations edit This section possibly contains original research Please improve it by verifying the claims made and adding inline citations Statements consisting only of original research should be removed April 2018 Learn how and when to remove this message The 1 D Saint Venant momentum equation can be derived from the Navier Stokes equations that describe fluid motion The x component of the Navier Stokes equations when expressed in Cartesian coordinates in the x direction can be written as u t u u x v u y w u z p x 1 r n 2 u x 2 2 u y 2 2 u z 2 f x displaystyle frac partial u partial t u frac partial u partial x v frac partial u partial y w frac partial u partial z frac partial p partial x frac 1 rho nu left frac partial 2 u partial x 2 frac partial 2 u partial y 2 frac partial 2 u partial z 2 right f x nbsp where u is the velocity in the x direction v is the velocity in the y direction w is the velocity in the z direction t is time p is the pressure r is the density of water n is the kinematic viscosity and fx is the body force in the x direction If it is assumed that friction is taken into account as a body force then n displaystyle nu nbsp can be assumed as zero so n 2 u x 2 2 u y 2 2 u z 2 0 displaystyle nu left frac partial 2 u partial x 2 frac partial 2 u partial y 2 frac partial 2 u partial z 2 right 0 nbsp Assuming one dimensional flow in the x direction it follows that 26 v u y w u z 0 displaystyle v frac partial u partial y w frac partial u partial z 0 nbsp Assuming also that the pressure distribution is approximately hydrostatic it follows that 26 p r g h displaystyle p rho gh nbsp or in differential form p r g h displaystyle partial p rho g partial h nbsp And when these assumptions are applied to the x component of the Navier Stokes equations p x 1 r 1 r r g h x g h x displaystyle frac partial p partial x frac 1 rho frac 1 rho frac rho g left partial h right partial x g frac partial h partial x nbsp There are 2 body forces acting on the channel fluid namely gravity and friction f x f x g f x f displaystyle f x f x g f x f nbsp where fx g is the body force due to gravity and fx f is the body force due to friction fx g can be calculated using basic physics and trigonometry 27 F g sin 8 g M displaystyle F g sin theta gM nbsp where Fg is the force of gravity in the x direction 8 is the angle and M is the mass nbsp Figure 1 Diagram of block moving down an inclined plane The expression for sin 8 can be simplified using trigonometry as sin 8 opp hyp displaystyle sin theta frac text opp text hyp nbsp For small 8 reasonable for almost all streams it can be assumed that sin 8 tan 8 opp adj S displaystyle sin theta tan theta frac text opp text adj S nbsp and given that fx represents a force per unit mass the expression becomes f x g g S displaystyle f x g gS nbsp Assuming the energy grade line is not the same as the channel slope and for a reach of consistent slope there is a consistent friction loss it follows that 28 f x f S f g displaystyle f x f S f g nbsp All of these assumptions combined arrives at the 1 dimensional Saint Venant equation in the x direction u t u u x g h x g S f S 0 displaystyle frac partial u partial t u frac partial u partial x g frac partial h partial x g S f S 0 nbsp a b c d e displaystyle a quad b quad c qquad d quad e nbsp where a is the local acceleration term b is the convective acceleration term c is the pressure gradient term d is the friction term and e is the gravity term Terms The local acceleration a can also be thought of as the unsteady term as this describes some change in velocity over time The convective acceleration b is an acceleration caused by some change in velocity over position for example the speeding up or slowing down of a fluid entering a constriction or an opening respectively Both these terms make up the inertia terms of the 1 dimensional Saint Venant equation The pressure gradient term c describes how pressure changes with position and since the pressure is assumed hydrostatic this is the change in head over position The friction term d accounts for losses in energy due to friction while the gravity term e is the acceleration due to bed slope Wave modelling by shallow water equations editShallow water equations can be used to model Rossby and Kelvin waves in the atmosphere rivers lakes and oceans as well as gravity waves in a smaller domain e g surface waves in a bath In order for shallow water equations to be valid the wavelength of the phenomenon they are supposed to model has to be much larger than the depth of the basin where the phenomenon takes place Somewhat smaller wavelengths can be handled by extending the shallow water equations using the Boussinesq approximation to incorporate dispersion effects 29 Shallow water equations are especially suitable to model tides which have very large length scales over hundred of kilometers For tidal motion even a very deep ocean may be considered as shallow as its depth will always be much smaller than the tidal wavelength nbsp Tsunami generation and propagation as computed with the shallow water equations red line without frequency dispersion and with a Boussinesq type model blue line with frequency dispersion Observe that the Boussinesq type model blue line forms a soliton with an oscillatory tail staying behind The shallow water equations red line form a steep front which will lead to bore formation later on The water depth is 100 meters Turbulence modelling using non linear shallow water equations edit nbsp A snapshot from simulation of shallow water equations in which shock waves are present Shallow water equations in its non linear form is an obvious candidate for modelling turbulence in the atmosphere and oceans i e geophysical turbulence An advantage of this over Quasi geostrophic equations is that it allows solutions like gravity waves while also conserving energy and potential vorticity However there are also some disadvantages as far as geophysical applications are concerned it has a non quadratic expression for total energy and a tendency for waves to become shock waves 30 Some alternate models have been proposed which prevent shock formation One alternative is to modify the pressure term in the momentum equation but it results in a complicated expression for kinetic energy 31 Another option is to modify the non linear terms in all equations which gives a quadratic expression for kinetic energy avoids shock formation but conserves only linearized potential vorticity 32 See also editWaves and shallow waterNotes edit Vreugdenhil C B 1986 Numerical Methods for Shallow Water Flow Water Science and Technology Library Vol 13 Springer Dordrecht p 262 doi 10 1007 978 94 015 8354 1 ISBN 978 90 481 4472 3 The Shallow Water Equations PDF Archived from the original PDF on 2012 03 16 Retrieved 2010 01 22 Clint Dawson and Christopher M Mirabito 2008 The Shallow Water Equations PDF Retrieved 2013 03 28 Carrier G F Yeh H 2005 Tsunami propagation from a finite source Computer Modeling in Engineering amp Sciences 10 2 113 122 doi 10 3970 cmes 2005 010 113 a b c d S Neelz G Pender 2009 Desktop review of 2D hydraulic modelling packages Joint Environment Agency Defra Flood and Coastal Erosion Risk Management Research and Development Programme Science Report SC080035 5 Archived from the original on 8 September 2019 Retrieved 2 December 2016 Saint Venant A J C Barre de 1871 Theorie du mouvement non permanent des eaux avec application aux crues des rivieres et a l introduction de marees dans leurs lits Comptes Rendus de l Academie des Sciences 73 147 154 and 237 240 a b Chow Ven Te 1959 Open channel hydraulics McGraw Hill OCLC 4010975 18 1 amp 18 2 a b Cunge J A F M Holly Jr and A Verwey 1980 Practical aspects of computational river hydraulics Pitman Publishing ISBN 0 273 08442 9 2 1 amp 2 2 a b c Whitham G B 1974 Linear and Nonlinear Waves 5 2 amp 13 10 Wiley ISBN 0 471 94090 9 Lighthill J 2005 Waves in fluids Cambridge University Press ISBN 978 0 521 01045 0 2 8 2 14 Meyer R E 1960 Theory of characteristics of inviscid gas dynamics In Fluid Dynamics Stromungsmechanik Encyclopedia of Physics IX Eds S Flugge amp C Truesdell Springer Berlin ISBN 978 3 642 45946 7 pp 225 282 a b Didenkulova I Pelinovsky E 2011 Rogue waves in nonlinear hyperbolic systems shallow water framework Nonlinearity 24 3 R1 R18 Bibcode 2011Nonli 24R 1D doi 10 1088 0951 7715 24 3 R01 S2CID 59438883 Harris M W Nicolsky D J Pelinovsky E N Rybkin A V 2015 03 01 Runup of Nonlinear Long Waves in Trapezoidal Bays 1 D Analytical Theory and 2 D Numerical Computations Pure and Applied Geophysics 172 3 4 885 899 Bibcode 2015PApGe 172 885H doi 10 1007 s00024 014 1016 3 ISSN 0033 4553 S2CID 55004099 Harris M W Nicolsky D J Pelinovsky E N Pender J M Rybkin A V 2016 05 01 Run up of nonlinear long waves in U shaped bays of finite length analytical theory and numerical computations Journal of Ocean Engineering and Marine Energy 2 2 113 127 doi 10 1007 s40722 015 0040 4 ISSN 2198 6444 S2CID 123725815 Garayshin V V Harris M W Nicolsky D J Pelinovsky E N Rybkin A V 2016 04 10 An analytical and numerical study of long wave run up in U shaped and V shaped bays Applied Mathematics and Computation 279 187 197 doi 10 1016 j amc 2016 01 005 Anderson Dalton Harris Matthew Hartle Harrison Nicolsky Dmitry Pelinovsky Efim Raz Amir Rybkin Alexei 2017 02 02 Run Up of Long Waves in Piecewise Sloping U Shaped Bays Pure and Applied Geophysics 174 8 3185 Bibcode 2017PApGe 174 3185A doi 10 1007 s00024 017 1476 3 ISSN 0033 4553 S2CID 132114728 Lannes D 2013 The Water Waves Problem Mathematical Analysis and Asymptotics Mathematical Surveys and Monographs American Mathematical Society p 174 ISBN 9780821894705 LCCN 2012046540 Brunner G W 1995 HEC RAS River Analysis System Hydraulic Reference Manual Version 1 0 Rep DTIC Document Searby D Dean A Margetts J 1998 Christchurch harbour Hydroworks modelling Proceedings of the WAPUG Autumn meeting Blackpool UK Havno K M Madsen J Dorge and V Singh 1995 MIKE 11 a generalized river modelling package Computer models of watershed hydrology 733 782 Yeh G Cheng J Lin J Martin W 1995 A numerical model simulating water flow and contaminant and sediment transport in watershed systems of 1 D stream river network 2 D overland regime and 3 D subsurface media Computer models of watershed hydrology 733 782 DHI Danish Hydraulic Institute 2011 MIKE SHE User Manual Volume 2 Reference Guide edited Bates P T Fewtrell M Trigg and J Neal 2008 LISFLOOD FP user manual and technical note code release 4 3 6 University of Bristol Novak P et al Hydraulic Modelling An Introduction Principles Methods and Applications 2010 CRC Press Scharffenberg W A and M J Fleming 2006 Hydrologic Modeling System HEC HMS User s Manual US Army Corps of Engineers Hydrologic Engineering Center a b Vincent Fromion 2009 Modeling and control of hydrosystems Springer ISBN 9781848826243 OCLC 401159458 Inclined Planes www physicsclassroom com Retrieved 2017 05 16 Methods Haestad 2007 Computer applications in hydraulic engineering connecting theory to practice Bentley Institute Press ISBN 978 0971414167 OCLC 636350249 Dingemans M W 1997 Wave propagation over uneven bottoms Advanced Series on Ocean Engineering 13 World Scientific Singapore pp 473 amp 516 ISBN 978 981 02 0427 3 Augier Pierre Mohanan Ashwin Vishnu Lindborg Erik 2019 09 17 Shallow water wave turbulence Journal of Fluid Mechanics 874 1169 1196 Bibcode 2019JFM 874 1169A doi 10 1017 jfm 2019 375 ISSN 1469 7645 S2CID 198976015 Buhler Oliver 1998 09 01 A Shallow Water Model that Prevents Nonlinear Steepening of Gravity Waves Journal of the Atmospheric Sciences 55 17 2884 2891 Bibcode 1998JAtS 55 2884B doi 10 1175 1520 0469 1998 055 lt 2884 ASWMTP gt 2 0 CO 2 ISSN 0022 4928 Lindborg Erik Mohanan Ashwin Vishnu 2017 11 01 A two dimensional toy model for geophysical turbulence Physics of Fluids 29 11 111114 Bibcode 2017PhFl 29k1114L doi 10 1063 1 4985990 ISSN 1070 6631 Further reading editBattjes J A Labeur R J 2017 Unsteady flow in open channels Cambridge University Press doi 10 1017 9781316576878 ISBN 978 1 107 15029 4 Vreugdenhil C B 1994 Numerical Methods for Shallow Water Flow Kluwer Academic Publishers ISBN 978 0792331643External links editDerivation of the shallow water equations from first principles instead of simplifying the Navier Stokes equations some analytical solutions Retrieved from https en wikipedia org w index php title Shallow water equations amp oldid 1221691404, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.