D'Alembert's paradox


In fluid dynamics, d'Alembert's paradox is a contradiction reached in 1752 by French mathematician Jean le Rond d'Alembert. D'Alembert proved that – for incompressible and inviscid potential flowthe drag force is zero on a body moving with constant velocity relative to the fluid. Zero drag is in direct contradiction to the observation of substantial drag on bodies moving relative to fluids, such as air and water; especially at high velocities corresponding with high Reynolds numbers. It is a particular example of the reversibility paradox.
D’Alembert, working on a 1749 Prize Problem of the Berlin Academy on flow drag, concluded: "It seems to me that the theory, developed in all possible rigor, gives, at least in several cases, a strictly vanishing resistance, a singular paradox which I leave to future Geometers to elucidate". A physical paradox indicates flaws in the theory.
Fluid mechanics was thus discredited by engineers from the start, which resulted in an unfortunate split – between the field of hydraulics, observing phenomena which could not be explained, and theoretical fluid mechanics explaining phenomena which could not be observed – in the words of the Chemistry Nobel Laureate Sir Cyril Hinshelwood.
According to scientific consensus, the occurrence of the paradox is due to the neglected effects of viscosity. In conjunction with scientific experiments, there were huge advances in the theory of viscous fluid friction during the 19th century. With respect to the paradox, this culminated in the discovery and description of thin boundary layers by Ludwig Prandtl in 1904. Even at very high Reynolds numbers, the thin boundary layers remain as a result of viscous forces. These viscous forces cause friction drag on streamlined objects, and for bluff bodies the additional result is flow separation and a low-pressure wake behind the object, leading to form drag.
The general view in the fluid mechanics community is that, from a practical point of view, the paradox is solved along the lines suggested by Prandtl. A formal mathematical proof is lacking, and difficult to provide, as in so many other fluid-flow problems involving the Navier–Stokes equations.

Viscous friction: Saint-Venant, Navier and Stokes

First steps towards solving the paradox were made by Saint-Venant, who modelled viscous fluid friction. Saint-Venant states in 1847:
Soon after, in 1851, Stokes calculated the drag on a sphere in Stokes flow, known as Stokes' law. Stokes flow is the low Reynolds-number limit of the Navier–Stokes equations describing the motion of a viscous liquid.
However, when the flow problem is put into a non-dimensional form, the viscous Navier–Stokes equations converge for increasing Reynolds numbers towards the inviscid Euler equations, suggesting that the flow should converge towards the inviscid solutions of potential flow theory – having the zero drag of the d'Alembert paradox. Of this, there is no evidence found in experimental measurements of drag and flow visualisations. This again raised questions concerning the applicability of fluid mechanics in the second half of the 19th century.

Inviscid separated flow: Kirchhoff and Rayleigh

In the second half of the 19th century, focus shifted again towards using inviscid flow theory for the description of fluid drag—assuming that viscosity becomes less important at high Reynolds numbers. The model proposed by Kirchhoff
and Rayleigh
was based on the free-streamline theory of Helmholtz and consists of a steady wake behind the body. Assumptions applied to the wake region include: flow velocities equal to the body velocity, and a constant pressure. This wake region is separated from the potential flow outside the body and wake by vortex sheets with discontinuous jumps in the tangential velocity across the interface.
In order to have a non-zero drag on the body, the wake region must extend to infinity. This condition is indeed fulfilled for the Kirchhoff flow perpendicular to a plate. The theory correctly states the drag force to be proportional to the square of the velocity.
In first instance, the theory could only be applied to flows separating at sharp edges. Later, in 1907, it was extended by Levi-Civita to flows separating from a smooth curved boundary.
It was readily known that such steady flows are not stable, since the vortex sheets develop so-called Kelvin–Helmholtz instabilities. But this steady-flow model was studied further in the hope it still could give a reasonable estimate of drag. Rayleigh asks "... whether the calculations of resistance are materially affected by this circumstance as the pressures experienced must be nearly independent of what happens at some distance in the rear of the obstacle, where the instability would first begin to manifest itself."
However, fundamental objections arose against this approach: Kelvin observed that if a plate is moving with constant velocity through the fluid the velocity in the wake is equal to that of the plate. The infinite extent of the wake—widening with the distance from the plate, as obtained from the theory—results in an infinite kinetic energy in the wake, which must be rejected on physical grounds.
Moreover, the observed pressure differences between front and back of the plate, and resulting drag forces, are much larger than predicted: for a flat plate perpendicular to the flow the predicted drag coefficient is
CD=0.88, while in experiments CD''=2.0 is found. This is mainly due to suction at the wake side of the plate, induced by the unsteady flow in the real wake.
So, this theory is found to be unsatisfactory as an explanation of drag on a body moving through a fluid. Although it can be applied to so-called cavity flows where, instead of a wake filled with fluid, a vacuum cavity is assumed to exist behind the body.

Thin boundary layers: Prandtl

The German physicist Ludwig Prandtl suggested in 1904 that the effects of a thin viscous boundary layer possibly could be the source of substantial drag. Prandtl put forward the idea that, at high velocities and high Reynolds numbers, a no-slip boundary condition causes a strong variation of the flow speeds over a thin layer near the wall of the body. This leads to the generation of vorticity and viscous dissipation of kinetic energy in the boundary layer. The energy dissipation, which is lacking in the inviscid theories, results for bluff bodies in separation of the flow. The low pressure in the wake region causes form drag, and this can be larger than the friction drag due to the viscous shear stress at the wall.
Evidence that Prandtl's scenario occurs for bluff bodies in flows of high Reynolds numbers can be seen in impulsively started flows around a cylinder. Initially the flow resembles potential flow, after which the flow separates near the rear stagnation point. Thereafter, the separation points move upstream, resulting in a low-pressure region of separated flow.
Prandtl made the hypothesis that the viscous effects are important in thin layers – called boundary layers – adjacent to solid boundaries, and that viscosity has no role of importance outside. The boundary-layer thickness becomes smaller when the viscosity reduces. The full problem of viscous flow, described by the non-linear Navier–Stokes equations, is in general not mathematically solvable. However, using his hypothesis Prandtl was able to derive an approximate model for the flow inside the boundary layer, called boundary-layer theory; while the flow outside the boundary layer could be treated using inviscid flow theory. Boundary-layer theory is amenable to the method of matched asymptotic expansions for deriving approximate solutions. In the simplest case of a flat plate parallel to the incoming flow, boundary-layer theory results in drag, whereas all inviscid flow theories will predict zero drag. Importantly for aeronautics, Prandtl's theory can be applied directly to streamlined bodies like airfoils where, in addition to surface-friction drag, there is also form drag. Form drag is due to the effect of the boundary layer and thin wake on the pressure distribution around the airfoil.

Open questions

To verify, as Prandtl suggested, that a vanishingly small cause has a large effect – substantial drag —
may be very difficult.
The mathematician Garrett Birkhoff in the opening chapter of his book Hydrodynamics from 1950, addresses a number of paradoxes of fluid mechanics and expresses a clear doubt in their official resolutions:
In particular, on d'Alembert's paradox, he considers another possible route to the creation of drag: instability of the potential flow solutions to the Euler equations. Birkhoff states:
In his 1951 review of Birkhoff's book, the mathematician James J. Stoker sharply criticizes the first chapter of the book:
In the second and revised edition of Birkhoff's Hydrodynamics in 1960, the above two statements no longer appear.
The importance and usefulness of the achievements, made on the subject of the d'Alembert paradox, are reviewed by Stewartson thirty years later. His long 1981 survey article starts with:
For many paradoxes in physics, their resolution often lies in transcending the available theory. In the case of d'Alembert's paradox, the essential mechanism for its resolution was provided by Prandtl through the discovery and modelling of thin viscous boundary layers – which are non-vanishing at high Reynolds numbers.
A new resolution, connecting to second quote of Birkhoff above, was published by , which is entirely different from Prandtl's resolution based on his boundary layer theory. The new resolution is based on the discovery supported by mathematical analysis and computation that potential flow with zero drag is an unphysical unstable formal mathematical solution of Euler's equations, which as physical flow from a basic instability at separation develops a turbulent wake creating drag. The new resolution questions Prandtl's legacy based on the concept of boundary layer and opens new possibilities in computational fluid mechanics explored in . The new resolution has led to a

Proof of zero drag in steady potential flow

Potential flow

The three main assumptions in the derivation of d'Alembert's paradox is that the steady flow is incompressible, inviscid and irrotational.
An inviscid fluid is described by the Euler equations, which together with the other two conditions read
where u denotes the flow velocity of the fluid, p the pressure, ρ the density, and is the gradient operator.
We have the second term in the Euler equation as:
where the first equality is a vector calculus identity and the second equality uses that the flow is irrotational. Furthermore, for every irrotational flow, there exists a velocity potential φ such that u = φ. Substituting this all in the equation for momentum conservation yields
Thus, the quantity between brackets must be constant. Assuming that the fluid is at rest at infinity and that the pressure is defined to be zero there, this constant is zero, and thus
which is the Bernoulli equation for unsteady potential flow.

Zero drag

Now, suppose that a body moves with constant velocity v through the fluid, which is at rest infinitely far away. Then the velocity field of the fluid has to follow the body, so it is of the form u = u, where x is the spatial coordinate vector, and thus:
Since u = φ, this can be integrated with respect to x:
The force
F that the fluid exerts on the body is given by the surface integral
where
A denotes the body surface and n the normal vector on the body surface. But it follows from that
thus
with the contribution of
R to the integral being equal to zero.
At this point, it becomes more convenient to work in the vector components. The
kth component of this equation reads
Let
V be the volume occupied by the fluid. The divergence theorem says that
The right-hand side is an integral over an infinite volume, so this needs some justification, which can be provided by appealing to potential theory to show that the velocity
u must fall off as r−3 – corresponding to a dipole potential field in case of a three-dimensional body of finite extent – where r is the distance to the centre of the body. The integrand in the volume integral can be rewritten as follows:
where first equality and then the incompressibility of the flow are used. Substituting this back into the volume integral and another application of the divergence theorem again. This yields
Substituting this in, we find that
The fluid cannot penetrate the body and thus
n · u = n · v
on the body surface. So and
Finally, the drag is the force in the direction in which the body moves, so
Hence the drag vanishes. This is d'Alembert's paradox.

Historical

*