HOME

TheInfoList



OR:

In
fluid dynamics In physics and engineering, fluid dynamics is a subdiscipline of fluid mechanics that describes the flow of fluids— liquids and gases. It has several subdisciplines, including ''aerodynamics'' (the study of air and other gases in motion) an ...
, Prandtl–Batchelor theorem states that ''if in a two-dimensional laminar flow at high Reynolds number closed streamlines occur, then the
vorticity In continuum mechanics, vorticity is a pseudovector field that describes the local spinning motion of a continuum near some point (the tendency of something to rotate), as would be seen by an observer located at that point and traveling along wit ...
in the closed streamline region must be a constant''. A similar statement holds true for axisymmetric flows. The theorem is named after
Ludwig Prandtl Ludwig Prandtl (4 February 1875 – 15 August 1953) was a German fluid dynamicist, physicist and aerospace scientist. He was a pioneer in the development of rigorous systematic mathematical analyses which he used for underlying the science of ...
and
George Batchelor George Keith Batchelor FRS (8 March 1920 – 30 March 2000) was an Australian applied mathematician and fluid dynamicist. He was for many years a Professor of Applied Mathematics in the University of Cambridge, and was founding head of the De ...
.
Prandtl Ludwig Prandtl (4 February 1875 – 15 August 1953) was a German fluid dynamicist, physicist and aerospace scientist. He was a pioneer in the development of rigorous systematic mathematical analyses which he used for underlying the science of ...
in his celebrated 1904 paper stated this theorem in arguments,
George Batchelor George Keith Batchelor FRS (8 March 1920 – 30 March 2000) was an Australian applied mathematician and fluid dynamicist. He was for many years a Professor of Applied Mathematics in the University of Cambridge, and was founding head of the De ...
unaware of this work proved the theorem in 1956. The problem was also studied in the same year by
Richard Feynman Richard Phillips Feynman (; May 11, 1918 – February 15, 1988) was an American theoretical physicist, known for his work in the path integral formulation of quantum mechanics, the theory of quantum electrodynamics, the physics of the superflu ...
and
Paco Lagerstrom Paco Axel Lagerstrom (February 24, 1914 – February 16, 1989) was an applied mathematician and aeronautical engineer. He was trained formally in mathematics, but worked for much of his career in aeronautical applications. He was known for wo ...
and by W.W. Wood in 1957.


Mathematical proof

At high
Reynolds numbers In fluid mechanics, the Reynolds number () is a dimensionless quantity that helps predict fluid flow patterns in different situations by measuring the ratio between inertial and viscous forces. At low Reynolds numbers, flows tend to be domin ...
,
Euler equations 200px, Leonhard Euler (1707–1783) In mathematics and physics, many topics are named in honor of Swiss mathematician Leonhard Euler (1707–1783), who made many important discoveries and innovations. Many of these items named after Euler include ...
reduce to solving a problem for
stream function The stream function is defined for incompressible flow, incompressible (divergence-free) fluid flow, flows in two dimensions – as well as in three dimensions with axisymmetry. The flow velocity components can be expressed as the derivatives of t ...
, :\nabla^2\psi = - \omega(\psi), \quad \psi=\psi_o \text \partial D. As it stands, the problem is ill-posed since the vorticity distribution \omega(\psi) can have infinite number of possibilities, all of which satisfies the equation and the boundary condition. This is not true if the streamlines are not closed, in which case, every streamline can be traced back to infinity, where \omega(\psi) is assumed to be prescribed. The difficulty arises only when closed streamlines occur inside the flow at high Reynolds number, where \omega(\psi) is not uniquely defined. The theorem asserts that \omega(\psi) is uniquely defined in such cases by examining the limiting process Re\rightarrow \infty properly. In two-dimensional flows, the only non-zero component lies in the z direction. The steady, non-dimensional vorticity equation in this case reduces to :\mathbf \cdot \nabla\mathbf = \frac\nabla^2\omega. Integrate the equation over a surface S lying entirely in the region where we have closed streamlines, bounded by a closed contour C :\int_S\mathbf \cdot \nabla\mathbf\, d\mathbf S = \frac\int_S\nabla^2\omega\, d\mathbf S. The integrand in the left-hand side term can be written as \nabla \cdot (\omega\mathbf u) since \nabla\cdot\mathbf u=0. By divergence theorem, one obtains :\oint_C \omega\mathbf\cdot \mathbf n dl = \frac\oint_C\nabla\omega\cdot \mathbf n dl. where \mathbf n is the outward unit vector normal to the contour line element dl. The left-hand side integrand can be made zero if the contour C is taken to be one of the closed streamlines since then the velocity vector projected normal to the contour will be zero, that is to say \mathbf u\cdot \mathbf n=0. Thus one obtains :\frac\oint_C \nabla\omega \cdot \mathbf\ dl = 0 This expression is true for finite but large Reynolds number since we did not neglect the viscous term before. Unlike the two-dimensional inviscid flows, where \omega=\omega(\psi) since \mathbf u\cdot \nabla \omega =0 with no restrictions on the functional form of \omega, in the viscous flows, \omega\neq \omega(\psi). But for large but finite \mathrm, we can write \omega=\omega(\psi) + \rm, and this small corrections become smaller and smaller as we increase the Reynolds number. Thus, in the limit \mathrm\rightarrow \infty, in the first approximation (neglecting the small corrections), we have : \frac\oint_C \nabla\omega \cdot \mathbf\ dl = \frac\oint_C \frac\nabla\psi \cdot \mathbf\ dl = 0. Since d\omega/d\psi is constant for a given streamline, we can take that term outside the integral, :\frac\frac\oint_C \nabla\psi \cdot \mathbf\ dl = 0. One may notice that the integral is negative of the circulation since :\Gamma = -\oint_C\mathbf u\cdot d\mathbf =-\int_S \omega d\mathbf S = \int_S \nabla^2\psi d\mathbf = \oint_C \nabla \psi \cdot \mathbf dl where we used the Stokes theorem for circulation and \omega=-\nabla^2\psi. Thus, we have :\frac\frac = 0. The circulation around those closed streamlines is not zero (unless the velocity at each point of the streamline is zero with a possible discontinuous vorticity jump across the streamline) . The only way the above equation can be satisfied is only if :\frac = 0, i.e., vorticity is not changing across these closed streamlines, thus proving the theorem. Of course, the theorem is not valid inside the boundary layer regime. This theorem cannot be derived from the Euler equations.Lagerstrom, P. A. (1975). Solutions of the Navier–Stokes equation at large Reynolds number. SIAM Journal on Applied mathematics, 28(1), 202-214.


References

{{DEFAULTSORT:Prandtl-Batchelor theorem Fluid dynamics