User:Kamil Kielczewski/sandbox

From Wikipedia, the free encyclopedia

Main equation[edit]

The Cauchy momentum equation is a vector partial differential equation put forth by Cauchy that describes the non-relativistic momentum transport in any continuum.[1] In convective (or Lagrangian) form it is written:

where

  • is the flow velocity vector field, which depends on time and space.
  • is the time,
  • is material derivative equal to ,
  • is the density at the point considered in the continuum (for which the continuity equation holds),
  • is the stress tensor,
  • contains all of the body forces acceleration (sometimes simply gravitational acceleration) (,
  • is divergence of stress tensor[2][3][4].

Note that above we use column vectors only for clarity - however equation is written using physical components (which are neither covariant(bottom index - row vectors) nor contravariant(top index - column vectors) )[5]. However if we chose non-orthogonal curvilinear coordinate system the we should calculate and write equation in covariant(row vectors) or contravariant(column vectors) form.

After an appropriate change of variables, it can also be written in conservation form:

where j is the momentum density at a given space-time point, F is the flux associated to the momentum density, and s contains all of the body forces per unit volume.

Differential derivation[edit]

Let’s start from generalized momentum conservation principle which can be write as follows: "The change in the momentum of the system is proportional to the resultant force acting on this system." it is expressed by formula [6]:

where is momentum in time t, is force averaged in . After dividing by and passing to the limit we get (derivative):

Lets analyse each side of above equation

Right side[edit]

The X component of forces acting on walls of cubic fluid element (green for upper-down walls; red for left-right; black for front-back).
On top chart we see approximation of function f(x) (blue line) using finit difference (yellow line). In the bottom chart we see „infinitely many times enlarged surroundings of point x1” (purple square from the upper graph). On the bottom line yellow line was not has not been applied. In the figure, using two equivalent marking derivative in addition marked .

We divide forces to body force and surface force

Surface forces works on walls of cubic fluid element. The X component of this forces, for each wall, was placed in the picture with a cubic element (in the form of a product of stress and surface area e.g. ).

Adding forces (their X components) acting on each of the cube walls, we get

After ordering and after performing similar reasoning for components (they are not in the drawing - they will be parallel vectors to the Y and Z axes, respectively) we get:

We can then write it in the symbolic operational form:

Mass forces act on the inside of the control volume - we can write them using the acceleration field (e.g. gravitational acceleration):

Left side[edit]

Lets calculate momentum of the cube:

Because we assume that that tested mass (cube) is constant in time, so

Left and Right side comparison[edit]

We have

then

then

Divide both sides by , and because we get:

which finish derivation.

Integral derivation[edit]

Applying Newton's second law (ith component) to a control volume in the continuum being modeled gives:

Then, based on the Reynolds transport theorem and using material derivative notation, one can write

where Ω represents the control volume. Since this equation must hold for any control volume, it must be true that the integrand is zero, from this the Cauchy momentum equation follows. The main step (not done above) in deriving this equation is establishing that the derivative of the stress tensor is one of the forces that constitutes Fi.[1]

Conservation form[edit]

Cauchy equations can also be put in the following form:

Cauchy momentum equation (conservation form)

simply by defining:

where j is the momentum density at the point considered in the continuum (for which the continuity equation holds), F is the flux associated to the momentum density, and s contains all of the body forces per unit volume. uu is the dyad of the velocity.

Here j and s have same number of dimensions N as the flow speed and the body acceleration, while F, being a tensor, has N2.[note 1]

In the Eulerian forms it is apparent that the assumption of no deviatoric stress brings Cauchy equations to the Euler equations.

Convective acceleration[edit]

An example of convective acceleration. The flow is steady (time-independent), but the fluid decelerates as it moves down the diverging duct (assuming incompressible or subsonic compressible flow).

A significant feature of the Navier–Stokes equations is the presence of convective acceleration: the effect of time-independent acceleration of a flow with respect to space. While individual continuum particles indeed experience time dependent acceleration, the convective acceleration of the flow field is a spatial effect, one example being fluid speeding up in a nozzle.

Regardless of what kind of continuum is being dealt with, convective acceleration is a nonlinear effect. Convective acceleration is present in most flows (exceptions include one-dimensional incompressible flow), but its dynamic effect is disregarded in creeping flow (also called Stokes flow). Convective acceleration is represented by the nonlinear quantity u · ∇u, which may be interpreted either as (u · ∇)u or as u · (∇u), with u the tensor derivative of the velocity vector u. Both interpretations give the same result.[7]

Advection operator vs tensor derivative[edit]

The convection term can be written as (u · ∇)u, where u · ∇ is the advection operator. This representation can be contrasted to the one in terms of the tensor derivative.[7] The tensor derivative u is the component-by-component derivative of the velocity vector, defined by [∇u]mi = ∂m vi, so that

Lamb form[edit]

The vector calculus identity of the cross product of a curl holds:

where the Feynman subscript notation a is used, which means the subscripted gradient operates only on the factor a.

Lamb in his famous classical book Hydrodynamics (1895),[8] still in print, used this identity to change the convective term of the flow velocity in rotational form, i.e. without a tensor derivative:[9][full citation needed][10]

where the vector is called the Lamb vector. The Cauchy momentum equation becomes:

Using the identity:

the Cauchy equation becomes:

In fact, in case of an external conservative field, by defining its potential φ:

In case of a steady flow the time derivative of the flow velocity disappears, so the momentum equation becomes:

And by projecting the momentum equation on the flow direction, i.e. along a streamline, the cross product disappears due to a vector calculus identity of the triple scalar product:

If the stress tensor is isotropic, then only the pressure enters, and the Euler momentum equation in the steady incompressible case becomes:

In the steady incompressible case the mass equation is simply:

that is, the mass conservation for a steady incompressible flow states that the density along a streamline is constant. This leads to a considerable simplification of the Euler momentum equation:

The convenience of defining the total head for an inviscid liquid flow is now apparent:

in fact, the above equation can be simply written as:

That is, the momentum balance for a steady inviscid and incompressible flow in an external conservative field states that the total head along a streamline is constant.

Irrotational flows[edit]

The Lamb form is also useful in irrotational flow, where the curl of the velocity (called vorticity) ω = ∇ × u is equal to zero. In that case, the convection term reduces to

Stresses[edit]

The effect of stress in the continuum flow is represented by the p and ∇ · τ terms; these are gradients of surface forces, analogous to stresses in a solid. Here p is the pressure gradient and arises from the isotropic part of the Cauchy stress tensor. This part is given by the normal stresses that occur in almost all situations. The anisotropic part of the stress tensor gives rise to ∇ · τ, which usually describes viscous forces; for incompressible flow, this is only a shear effect. Thus, τ is the deviatoric stress tensor, and the stress tensor is equal to:[11][full citation needed]

where I is the identity matrix in the space considered and τ the shear tensor.

All non-relativistic momentum conservation equations, such as the Navier–Stokes equation, can be derived by beginning with the Cauchy momentum equation and specifying the stress tensor through a constitutive relation. By expressing the shear tensor in terms of viscosity and fluid velocity, and assuming constant density and viscosity, the Cauchy momentum equation will lead to the Navier–Stokes equations. By assuming inviscid flow, the Navier–Stokes equations can further simplify to the Euler equations.

The divergence of the stress tensor can be written as

The effect of the pressure gradient on the flow is to accelerate the flow in the direction from high pressure to low pressure.

As written in the Cauchy momentum equation, the stress terms p and τ are yet unknown, so this equation alone cannot be used to solve problems. Besides the equations of motion—Newton's second law—a force model is needed relating the stresses to the flow motion.[12] For this reason, assumptions based on natural observations are often applied to specify the stresses in terms of the other flow variables, such as velocity and density.

External forces[edit]

The vector field f represents body forces per unit mass. Typically, these consist of only gravity acceleration, but may include others, such as electromagnetic forces. In non-inertial coordinate frames, other "inertial accelerations" associated with rotating coordinates may arise.

Often, these forces may be represented as the gradient of some scalar quantity χ, with f = ∇χ in which case they are called conservative forces. Gravity in the z direction, for example, is the gradient of ρgz. Because pressure from such gravitation arises only as a gradient, we may include it in the pressure term as a body force h = pχ. The pressure and force terms on the right-hand side of the Navier–Stokes equation become

Nondimensionalisation[edit]

In order to make the equations dimensionless, a characteristic length r0 and a characteristic velocity u0 need to be defined. These should be chosen such that the dimensionless variables are all of order one. The following dimensionless variables are thus obtained:

Substitution of these inverted relations in the Euler momentum equations yields:

and by dividing for the first coefficient:

Now defining the Froude number:

the Euler number:

and the coefficient of skin-friction or the one usually referred as 'drag' co-efficient in the field of aerodynamics:

by passing respectively to the conservative variables, i.e. the momentum density and the force density:

the equations are finally expressed (now omitting the indexes):

Cauchy momentum equation (nondimensional conservative form)

Cauchy equations in the Froude limit Fr → ∞ (corresponding to negligible external field) are named free Cauchy equations:

Free Cauchy momentum equation (nondimensional conservative form)

and can be eventually conservation equations. The limit of high Froude numbers (low external field) is thus notable for such equations and is studied with perturbation theory.

Finally in convective form the equations are:

Cauchy momentum equation (nondimensional convective form)

3D explicit convective forms[edit]

Cartesian 3D coordinates[edit]

In general for asymmetric stress tensor, equations have following form [2] [3] [4] [13]

Cylindrical 3D coordinates[edit]

Below we write main equation in pressure-tau form assuming that stress tensor is symmetric ():

See also[edit]

Notes[edit]

  1. ^ In 3D for example, with respect to some coordinate system, the vector j has 3 components, while the tensors σ and F have 9 (3x3), so the explicit forms written as matrices would be:
    Note, however, that if symmetrical, F will only contain 6 degrees of freedom. And F's symmetry is equivalent to σ's symmetry (which will be present for the most common Cauchy stress tensors), since dyads of vectors with themselves are always symmetrical.

References[edit]

  1. ^ a b Acheson, D. J. (1990). Elementary Fluid Dynamics. Oxford University Press. p. 205. ISBN 0-19-859679-0.
  2. ^ a b "Behavior of a Vorticity-Influenced Asymmetric Stress Tensor in Fluid Flow" (PDF). AIR FORCE WRIGHT AERONAUTICAL LABORATORIES. 1986. p. 13 (below main equation author describe ) </math>). {{cite news}}: Cite uses deprecated parameter |authors= (help)
  3. ^ a b Viscous Fluid Flow (PDF). CRC Press. 2000. p. 66,68,143,182 authors use . ISBN 0-8493-1606-5. {{cite book}}: Cite uses deprecated parameter |authors= (help)
  4. ^ a b Introduction to Chemical Engineering Fluid Mechanics. Cambridge University Press. 2016. pp. 133–136. ISBN 978-1-107-12377-9. {{cite book}}: Cite uses deprecated parameter |authors= (help)
  5. ^ "A Primer on Tensor Calculus" (PDF). 2011. p. 11 (pdf 15). {{cite web}}: Cite uses deprecated parameter |authors= (help)
  6. ^ Anderson, Jr., John D. (1995). Computational Fluid Dynamics (PDF). New York: McGraw-Hill. pp. 61–64. ISBN 0-07-001685-2.
  7. ^ a b Emanuel, G. (2001). Analytical fluid dynamics (second ed.). CRC Press. p. 6–7. ISBN 0-8493-9114-8.
  8. ^ Valorani, Nasuti. "Metodi di analisi delle turbomacchine" (PDF) (in Italian). p. 11–12.
  9. ^ See Batchelor (1967), §3.5, p. 160.
  10. ^ Weisstein, Eric W. "Convective Derivative". MathWorld.
  11. ^ Batchelor (1967) p. 142.
  12. ^ Feynman, Richard P.; Leighton, Robert B.; Sands, Matthew (1963), The Feynman Lectures on Physics, Reading, Massachusetts: Addison-Wesley, Vol. 1, §9–4 and §12–1, ISBN 0-201-02116-1
  13. ^ Adam Powell (12 April 2010). "The Navier-Stokes Equations" (PDF). p. 2 (author use ).

Category:Concepts in physics Category:Continuum mechanics Category:Partial differential equations