Anda di halaman 1dari 9

Euler equations (fluid dynamics)

From Wikipedia, the free encyclopedia

In fluid dynamics, the Euler equations are a set of equations governing inviscid flow. They are named after
Leonhard Euler. The equations represent conservation of mass (continuity), momentum, and energy, corresponding
to the Navier–Stokes equations with zero viscosity and heat conduction terms. Historically, only the continuity and
momentum equations have been derived by Euler. However, fluid dynamics literature often refers to the full set –
including the energy equation – together as "the Euler equations".[1]

Like the Navier-Stokes equations, the Euler equations are usually written in one of two forms: the "conservation
form" and the "non-conservation form". The conservation form emphasizes the physical interpretation of the
equations as conservation laws through a control volume fixed in space. The non-conservation form emphasizes
changes to the state of a control volume as it moves with the fluid.

The Euler equations can be applied to compressible as well as to incompressible flow – using either an appropriate
equation of state or assuming that the divergence of the flow velocity field is zero, respectively.

Contents
1 History
2 Conservation and component form
3 Conservation and vector form
4 Non-conservation form with flux Jacobians
4.1 Flux Jacobians for an ideal gas
4.2 Linearized form
4.3 Uncoupled wave equations for the linearized one-dimensional case
5 Shock waves
6 The equations in one spatial dimension
7 Steady flow in streamline coordinates
7.1 Streamline curvature theorem
8 See also
9 Notes
10 Further reading
11 External links

History
The Euler equations first appeared in published form in Euler's article "Principes généraux du mouvement des
fluides," published in Mémoires de l'Academie des Sciences de Berlin in 1757 (in this article Euler actually
published only the general form of the continuity equation and the momentum equation;[2] the energy conservation
equation will be obtained a century later). They were among the first partial differential equations to be written
down. At the time Euler published his work, the system of equations consisted of the momentum and continuity
equations, thus it was underdetermined except in the case of an incompressible fluid. An additional equation, which
was later to be called the adiabatic condition, was supplied by Pierre-Simon Laplace in 1816.

During the second half of the 19th century, it was found that the equation related to the conservation of energy must
at all times be kept, while the adiabatic condition is a consequence of the fundamental laws in the case of smooth
solutions. With the discovery of the special theory of relativity, the concepts of energy density, momentum density,
and stress were unified into the concept of the stress-energy tensor, and energy and momentum were likewise
unified into a single concept, the energy-momentum vector.[3]

Conservation and component form


In differential form, the equations are:

where

ρ is the fluid mass density,


u is the fluid velocity vector, with components u, v, and w,
E = ρ e + ½ ρ ( u2 + v 2 + w2 ) is the total energy per unit volume, with e being the internal energy per unit
mass for the fluid,
p is the pressure,
denotes the tensor product, and
0 being the zero vector.

These equations may be expressed in subscript notation. The second equation includes the divergence of a dyadic
product, and may be clearer in subscript notation:

where the i and j subscripts label the three Cartesian components: ( x 1 , x 2 , x 3 ) = ( x , y , z ) and ( u1 , u2 , u3 )
= ( u , v , w ). These equations may be more succinctly expressed using Einstein notation, in which matched indices
imply a sum over those indices and and :

The above equations are expressed in conservation form, as this format emphasizes their physical origins (and is
often the most convenient form for computational fluid dynamics simulations). By subtracting the velocity times the
mass conservation term, the second equation (momentum conservation), can also be expressed as:

or, in vector notation:

but this form for the momentum conservation equation obscures the direct connection between the Euler equations
and Newton's second law of motion. Similarly, by subtracting the velocity times the above momentum conservation
term, the third equation (energy conservation), can also be expressed as:

or

Conservation and vector form


In vector and conservation form, the Euler equations become:

where
This form makes it clear that fx, fy and fz are fluxes.

The equations above thus represent conservation of mass, three components of momentum, and energy. There are
thus five equations and six unknowns. Closing the system requires an equation of state; the most commonly used is
the ideal gas law (i.e. p = ρ (γ−1) e, where ρ is the density, γ is the adiabatic index, and e the internal energy).

Note the odd form for the energy equation; see Rankine–Hugoniot equation. The extra terms involving p may be
interpreted as the mechanical work done on a fluid element by its neighbor fluid elements. These terms sum to zero
in an incompressible fluid.

The well-known Bernoulli's equation can be derived by integrating Euler's equation along a streamline, under the
assumption of constant density and a sufficiently stiff equation of state.

Non-conservation form with flux Jacobians


Expanding the fluxes can be an important part of constructing numerical solvers, for example by exploiting
(approximate) solutions to the Riemann problem. From the original equations as given above in vector and
conservation form, the equations are written in a non-conservation form as:

where Ax, Ay and Az are called the flux Jacobians, which are matrices equal to:

Here, the flux Jacobians Ax, Ay and Az are still functions of the state vector m, so this form of the Euler equations is
nonlinear, just like the original equations. This non-conservation form is equivalent to the original Euler equations in
conservation form, at least in regions where the state vector m varies smoothly.

Flux Jacobians for an ideal gas

The ideal gas law is used as the equation of state, to derive the full Jacobians in matrix form, as given below:[4]

Flux Jacobians in matrix form for an ideal gas


The x-direction flux Jacobian:
The y-direction flux Jacobian:

The z-direction flux Jacobian:

Where .

The total enthalpy H is given by:

and the speed of sound a is given as:

Linearized form

The linearized Euler equations are obtained by linearization of the Euler equations in non-conservation form with flux
Jacobians, around a state m = m0, and are given by:
where Ax,0, Ay,0 and Az,0 are the values of respectively Ax, Ay and Az at some reference state m = m0.

Uncoupled wave equations for the linearized one-dimensional case

The Euler equations can be transformed into uncoupled wave equations if they are expressed in characteristic
variables instead of conserved variables. As an example, the one-dimensional (1-D) Euler equations in linear flux-
Jacobian form is considered:

The matrix Ax,0 is diagonalizable, which means it can be decomposed into:

Here r1, r2, r3 are the right eigenvectors of the matrix Ax,0 corresponding with the eigenvalues λ1 , λ2 and λ3.

Defining the characteristic variables as:

Since Ax,0 is constant, multiplying the original 1-D equation in flux-Jacobian form with P−1 yields:

The equations have been essentially decoupled and turned into three wave equations, with the eigenvalues being the
wave speeds. The variables wi are called Riemann invariants or, for general hyperbolic systems, they are called
characteristic variables.

Shock waves
The Euler equations are nonlinear hyperbolic equations and their general solutions are waves. Much like the familiar
oceanic waves, waves described by the Euler Equations 'break' and so-called shock waves are formed; this is a
nonlinear effect and represents the solution becoming multi-valued. Physically this represents a breakdown of the
assumptions that led to the formulation of the differential equations, and to extract further information from the
equations we must go back to the more fundamental integral form. Then, weak solutions are formulated by working
in 'jumps' (discontinuities) into the flow quantities – density, velocity, pressure, entropy – using the Rankine–
in 'jumps' (discontinuities) into the flow quantities – density, velocity, pressure, entropy – using the Rankine–
Hugoniot shock conditions. Physical quantities are rarely discontinuous; in real flows, these discontinuities are
smoothed out by viscosity. (See Navier–Stokes equations)

Shock propagation is studied – among many other fields – in aerodynamics and rocket propulsion, where
sufficiently fast flows occur.

The equations in one spatial dimension


For certain problems, especially when used to analyze compressible flow in a duct or in case the flow is cylindrically
or spherically symmetric, the one-dimensional Euler equations are a useful first approximation. Generally, the Euler
equations are solved by Riemann's method of characteristics. This involves finding curves in plane of independent
variables (i.e., x and t) along which partial differential equations (PDE's) degenerate into ordinary differential
equations (ODE's). Numerical solutions of the Euler equations rely heavily on the method of characteristics.

Steady flow in streamline coordinates


In the case of steady flow, it is convenient to choose the Frenet–Serret frame along a streamline as the coordinate
system for describing the momentum part of the Euler equations:[5]

where v, p and ρ denote the velocity, the pressure and the density, respectively.

Let {es , en, eb } be a Frenet–Serret orthonormal basis which consists of a tangential unit vector, a normal unit
vector, and a binormal unit vector to the streamline, respectively. Since a streamline is a curve that is tangent to the
velocity vector of the flow, the left-handed side of the above equation, the substantial derivative of velocity, can be
described as follows:

where R is the radius of curvature of the streamline.

Therefore, the momentum part of the Euler equations for a steady flow is found to have a simple form:
For barotropic flow ( ρ=ρ(p) ), Bernoulli's equation is derived from the first equation:

The second equation expresses that, in the case the streamline is curved, there should exist a pressure gradient
normal to the streamline because the centripetal acceleration of the fluid parcel is only generated by the normal
pressure gradient.

The third equation expresses that pressure is constant along the binormal axis.

Streamline curvature theorem

Let r be the distance from the center of curvature of the


streamline, then the second equation is written as follows:

where

This equation states:


The "Streamline curvature theorem" states that the
In a steady flow of an inviscid fluid without
pressure at the upper surface of an airfoil is lower
external forces, the center of curvature of the
than the pressure far away and that the pressure at
streamline lies in the direction of decreasing
the lower surface is higher than the pressure far
radial pressure.
away; hence the pressure difference between the
Although this relationship between the pressure field and upper and lower surfaces of an airfoil generates a
flow curvature is very useful, it doesn't have a name in the lift force.
English-language scientific literature.[6] Japanese fluid-
dynamicists call the relationship the "Streamline curvature theorem". [7]

This "theorem" explains clearly why there are such low pressures in the centre of vortices,[6] which consist of
concentric circles of streamlines. This also is a way to intuitively explain why airfoils generate lift forces.[6]

See also
Madelung equations
Notes
1. ^ Anderson, John D. (1995), Computational Fluid Dynamics, The Basics With Applications. ISBN 0-07-113210-4
2. ^ E226 -- Principes generaux du mouvement des fluides (http://www.math.dartmouth.edu/~euler/pages/E226.html)
3. ^ Christodoulou, Demetrios (October 2007). "The Euler Equations of Compressible Fluid Flow"
(http://www.ams.org/bull/2007-44-04/S0273-0979-07-01181-0/S0273-0979-07-01181-0.pdf). Bulletin of the
American Mathematical Society 44 (4): 581–602. doi:10.1090/S0273-0979-07-01181-0
(http://dx.doi.org/10.1090%2FS0273-0979-07-01181-0). Retrieved June 13, 2009.
4. ^ See Toro (1999)
5. ^ James A. Fay (June 1994). Introduction to Fluid Mechanics. MIT Press. ISBN 0-262-06165-1. see "4.5 Euler's
Equation in Streamline Coordinates" pp.150-pp.152 (http://books.google.com/books?id=XGVpue4954wC&pg=150)
6. ^ a b c Babinsky, Holger (November 2003), "How do wings work?" (http://www.iop.org/EJ/article/0031-
9120/38/6/001/pe3_6_001.pdf), Physics Education
7. ^ 今井 功 (IMAI, Isao) (November 1973). 『流体力学(前編)』(Fluid Dynamics 1) (in Japanese). 裳華房
(Shoukabou). ISBN 4-7853-2314-0.

Further reading
Batchelor, G. K. (1967). An Introduction to Fluid Dynamics. Cambridge University Press. ISBN 0-521-
66396-2.
Thompson, Philip A. (1972). Compressible Fluid Flow. New York: McGraw-Hill. ISBN 0-07-064405-5.
Toro, E.F. (1999). Riemann Solvers and Numerical Methods for Fluid Dynamics. Springer-Verlag.
ISBN 3-540-65966-8.

External links
An introductory explanation of the Euler equations at a conceptual level
(http://wadeawalker.wordpress.com/2013/02/07/fluid-dynamics-super-awesome-or-super-awesomer/)

Retrieved from "http://en.wikipedia.org/w/index.php?title=Euler_equations_(fluid_dynamics)&oldid=564025522"


Categories: Concepts in physics Equations of fluid dynamics

This page was last modified on 12 July 2013 at 21:39.


Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply.
By using this site, you agree to the Terms of Use and Privacy Policy.
Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.

Anda mungkin juga menyukai