Open-channel flow

Last updated

In fluid mechanics and hydraulics, open-channel flow is a type of liquid flow within a conduit with a free surface, known as a channel. [1] [2] The other type of flow within a conduit is pipe flow. These two types of flow are similar in many ways but differ in one important respect: open-channel flow has a free surface, whereas pipe flow does not, resulting in flow dominated by gravity but not hydraulic pressure.

Contents

Central Arizona Project channel. Arizona cap canal.jpg
Central Arizona Project channel.

Classifications of flow

Open-channel flow can be classified and described in various ways based on the change in flow depth with respect to time and space. [3] The fundamental types of flow dealt with in open-channel hydraulics are:

States of flow

The behavior of open-channel flow is governed by the effects of viscosity and gravity relative to the inertial forces of the flow. Surface tension has a minor contribution, but does not play a significant enough role in most circumstances to be a governing factor. Due to the presence of a free surface, gravity is generally the most significant driver of open-channel flow; therefore, the ratio of inertial to gravity forces is the most important dimensionless parameter. [4] The parameter is known as the Froude number, and is defined as:

where is the mean velocity, is the characteristic length scale for a channel's depth, and is the gravitational acceleration. Depending on the effect of viscosity relative to inertia, as represented by the Reynolds number, the flow can be either laminar, turbulent, or transitional. However, it is generally acceptable to assume that the Reynolds number is sufficiently large so that viscous forces may be neglected. [4]

Formulation

It is possible to formulate equations describing three conservation laws for quantities that are useful in open-channel flow: mass, momentum, and energy. The governing equations result from considering the dynamics of the flow velocity vector field with components . In Cartesian coordinates, these components correspond to the flow velocity in the x, y, and z axes respectively.

To simplify the final form of the equations, it is acceptable to make several assumptions:

  1. The flow is incompressible (this is not a good assumption for rapidly-varied flow)
  2. The Reynolds number is sufficiently large such that viscous diffusion can be neglected
  3. The flow is one-dimensional across the x-axis

Continuity equation

The general continuity equation, describing the conservation of mass, takes the form:

where is the fluid density and is the divergence operator. Under the assumption of incompressible flow, with a constant control volume , this equation has the simple expression . However, it is possible that the cross-sectional area can change with both time and space in the channel. If we start from the integral form of the continuity equation:

it is possible to decompose the volume integral into a cross-section and length, which leads to the form:

Under the assumption of incompressible, 1D flow, this equation becomes:

By noting that and defining the volumetric flow rate , the equation is reduced to:

Finally, this leads to the continuity equation for incompressible, 1D open-channel flow:

Momentum equation

The momentum equation for open-channel flow may be found by starting from the incompressible Navier-Stokes equations  :

where is the pressure, is the kinematic viscosity, is the Laplace operator, and is the gravitational potential. By invoking the high Reynolds number and 1D flow assumptions, we have the equations:

The second equation implies a hydrostatic pressure , where the channel depth is the difference between the free surface elevation and the channel bottom . Substitution into the first equation gives:

where the channel bed slope . To account for shear stress along the channel banks, we may define the force term to be:

where is the shear stress and is the hydraulic radius. Defining the friction slope , a way of quantifying friction losses, leads to the final form of the momentum equation:

Energy equation

To derive an energy equation, note that the advective acceleration term may be decomposed as:

where is the vorticity of the flow and is the Euclidean norm. This leads to a form of the momentum equation, ignoring the external forces term, given by:

Taking the dot product of with this equation leads to:

This equation was arrived at using the scalar triple product . Define to be the energy density:

Noting that is time-independent, we arrive at the equation:

Assuming that the energy density is time-independent and the flow is one-dimensional leads to the simplification:

with being a constant; this is equivalent to Bernoulli's principle. Of particular interest in open-channel flow is the specific energy , which is used to compute the hydraulic head that is defined as:

with being the specific weight. However, realistic systems require the addition of a head loss term to account for energy dissipation due to friction and turbulence that was ignored by discounting the external forces term in the momentum equation.

See also

Related Research Articles

Continuum mechanics is a branch of mechanics that deals with the deformation of and transmission of forces through materials modeled as a continuous medium rather than as discrete particles. The French mathematician Augustin-Louis Cauchy was the first to formulate such models in the 19th century.

<span class="mw-page-title-main">Mach number</span> Ratio of speed of an object moving through fluid and local speed of sound

The Mach number, often only Mach, is a dimensionless quantity in fluid dynamics representing the ratio of flow velocity past a boundary to the local speed of sound. It is named after the Austrian physicist and philosopher Ernst Mach.

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

The vorticity equation of fluid dynamics describes the evolution of the vorticity ω of a particle of a fluid as it moves with its flow; that is, the local rotation of the fluid. The governing equation is:

In mathematics, the Laplace operator or Laplacian is a differential operator given by the divergence of the gradient of a scalar function on Euclidean space. It is usually denoted by the symbols , (where is the nabla operator), or . In a Cartesian coordinate system, the Laplacian is given by the sum of second partial derivatives of the function with respect to each independent variable. In other coordinate systems, such as cylindrical and spherical coordinates, the Laplacian also has a useful form. Informally, the Laplacian Δf (p) of a function f at a point p measures by how much the average value of f over small spheres or balls centered at p deviates from f (p).

In fluid dynamics, Stokes' law is an empirical law for the frictional force – also called drag force – exerted on spherical objects with very small Reynolds numbers in a viscous fluid. It was derived by George Gabriel Stokes in 1851 by solving the Stokes flow limit for small Reynolds numbers of the Navier–Stokes equations.

<span class="mw-page-title-main">Euler equations (fluid dynamics)</span> Set of quasilinear hyperbolic equations governing adiabatic and inviscid flow

In fluid dynamics, the Euler equations are a set of quasilinear partial differential equations governing adiabatic and inviscid flow. They are named after Leonhard Euler. In particular, they correspond to the Navier–Stokes equations with zero viscosity and zero thermal conductivity.

A continuity equation or transport equation is an equation that describes the transport of some quantity. It is particularly simple and powerful when applied to a conserved quantity, but it can be generalized to apply to any extensive quantity. Since mass, energy, momentum, electric charge and other natural quantities are conserved under their respective appropriate conditions, a variety of physical phenomena may be described using continuity equations.

In fluid mechanics or more generally continuum mechanics, incompressible flow refers to a flow in which the material density is constant within a fluid parcel—an infinitesimal volume that moves with the flow velocity. An equivalent statement that implies incompressibility is that the divergence of the flow velocity is zero.

In fluid dynamics, turbulence kinetic energy (TKE) is the mean kinetic energy per unit mass associated with eddies in turbulent flow. Physically, the turbulence kinetic energy is characterised by measured root-mean-square (RMS) velocity fluctuations. In the Reynolds-averaged Navier Stokes equations, the turbulence kinetic energy can be calculated based on the closure method, i.e. a turbulence model.

<span class="mw-page-title-main">Shallow water equations</span> Set of partial differential equations that describe the flow below a pressure surface in a fluid

The shallow-water equations (SWE) are a set of hyperbolic partial differential equations that describe the flow below a pressure surface in a fluid. The shallow-water equations in unidirectional form are also called Saint-Venant equations, after Adhémar Jean Claude Barré de Saint-Venant.

Conservation form or Eulerian form refers to an arrangement of an equation or system of equations, usually representing a hyperbolic system, that emphasizes that a property represented is conserved, i.e. a type of continuity equation. The term is usually used in the context of continuum mechanics.

The intent of this article is to highlight the important points of the derivation of the Navier–Stokes equations as well as its application and formulation for different families of fluids.

Pressure-correction method is a class of methods used in computational fluid dynamics for numerically solving the Navier-Stokes equations normally for incompressible flows.

The Cauchy momentum equation is a vector partial differential equation put forth by Cauchy that describes the non-relativistic momentum transport in any continuum.

In fluid dynamics, Luke's variational principle is a Lagrangian variational description of the motion of surface waves on a fluid with a free surface, under the action of gravity. This principle is named after J.C. Luke, who published it in 1967. This variational principle is for incompressible and inviscid potential flows, and is used to derive approximate wave models like the mild-slope equation, or using the averaged Lagrangian approach for wave propagation in inhomogeneous media.

In fluid dynamics, the Basset–Boussinesq–Oseen equation describes the motion of – and forces on – a small particle in unsteady flow at low Reynolds numbers. The equation is named after Joseph Valentin Boussinesq, Alfred Barnard Basset and Carl Wilhelm Oseen.

In fluid dynamics, the Craik–Leibovich (CL) vortex force describes a forcing of the mean flow through wave–current interaction, specifically between the Stokes drift velocity and the mean-flow vorticity. The CL vortex force is used to explain the generation of Langmuir circulations by an instability mechanism. The CL vortex-force mechanism was derived and studied by Sidney Leibovich and Alex D. D. Craik in the 1970s and 80s, in their studies of Langmuir circulations.

The porous medium equation, also called the nonlinear heat equation, is a nonlinear partial differential equation taking the form:

In fluid dynamics, the general equation of heat transfer is a nonlinear partial differential equation describing specific entropy production in a Newtonian fluid subject to thermal conduction and viscous forces:

References

  1. Chow, Ven Te (2008). Open-Channel Hydraulics (PDF). Caldwell, NJ: The Blackburn Press. ISBN   978-1932846188.
  2. Battjes, Jurjen A.; Labeur, Robert Jan (2017). Unsteady Flow in Open Channels. Cambridge, UK: Cambridge University Press. ISBN   9781316576878.
  3. Jobson, Harvey E.; Froehlich, David C. (1988). Basic Hydraulic Principles of Open-Channel Flow (PDF). Reston, VA: U.S. Geological Survey.
  4. 1 2 Sturm, Terry W. (2001). Open Channel Hydraulics (PDF). New York, NY: McGraw-Hill. p. 2. ISBN   9780073397870.

Further reading