Sunteți pe pagina 1din 26

Chapter 5 Inviscid Flows and Irrotational Flows

This chapter is divided between two topics that are related but must be distinguished carefully: inviscid ows and irrotational ows. Even though there is considerable overlap between the two topics, inviscid ows (Euler equation) can contain vorticity. The mechanisms by which viscous forces forces introduce vorticity in a ow will be treated in Ch. 8 and 9. And another distinction must be made right between inviscid ows, which have unique solutions (if irrotational), and very-large Reynolds number ows, which are generally turbulent.

5.1

Inviscid ow

In the absence of viscosity, Eulers equations are the expression of Newtons second law for incompressible ow. As seen above p u2 t u + u = ( + + gz) = B 2 and in index notations 1 t ui + uj j ui = gi3 i p 119 (5.2) (5.1)

120

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

5.2

Bernoullis equation

There are several forms of Bernoullis equations, relying on more or less restrictive assumptions. The student should make a clear distinction between them. In all cases, we start from the Navier-Stokes equations: t u + u = B where B= , (5.3)

p u2 + + gz, (5.4) 2 and look for ways to isolate the gradient in the r.h.s. (The Bernoulli term B is a scalar and cannot be confused with the body force per unit volume, which is a vector.)

5.2.1

The strong form

The simplest way is to assume that the ow is irrotational. A necessary and sucient condition for this to hold is that u= = 0. (5.5)

(see Helmholtz decomposition). Then, the Magnus and viscous terms vanish, and the time derivative can be brought inside the gradient, to give (t + implying that t + B = 0 (5.7) has a constant value in the entire eld. The assumptions made here seem to hold in simple wave motion, where the time-dependent potential is needed (see Ch. 11 for simple examples). If, furthermore, the ow is steady, there is also a useful relation known as Eulers normal equation. The normal pressure gradient is only balanced by the centripetal acceleration. So, 1 u2 n p = R where R is the local radius of curvature of the streamline. (5.8) p u2 + + gz) = 0 2 (5.6)

5.2. BERNOULLIS EQUATION

121

5.2.2

The weaker form

A separate version of Bernoullis equation is not as restrictive: allowing vorticity, the initial assumption is that viscous eects can be neglected. Thus, we work with Eulers equation t u + u = B (5.9)

Then, project the equation on the direction of the velocity vector itself, i.e. along the streamlines. This cancels out the Magnus term, and we have t u2 = u 2 B (5.10)

In words, the local rate of change of kinetic energy (per unit mass) comes at the expense of the convective (directional) derivative of the Bernoulli expression. If we restrict ourselves to steady ows, we see that p u2 + gz B= + 2 (5.11)

is constant along any streamline. This does not preclude B from having dierent values along dierent streamlines. This is the form of the equation generally presented in undergraduate texts; streamline independence is usually achieved through reservoirs where static conditions prevail. At this point, we see that the undergraduate version of Bernoullis equation (along streamlines) is only part of the picture. Incompressible inviscid ows remain as strong conditions, but unsteadiness can be accommodated (Fig. 5.1), and the streamline restriction disappears in irrotational ow. We saw in Ch. 3 how viscous eects can be handled as losses along streamlines. Furthermore, if there is vorticity in steady ow, not only is B constant along streamlines, but is is also constant along vortex lines since B=0 (5.12)

5.2.3

Bernoulli and energy

Equation (5.10) is derived directly from the momentum equation. But it can also be read as a scalar equation for kinetic energy, covered in Ch.3. This is not a new idea: in elementary frictionless dynamics, the energy equation

122

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

Figure 5.1: Bernoullis equation and assumptions

5.2. BERNOULLIS EQUATION

123

Figure 5.2: Bernoulli and energy balance

124

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

(kinetic and potential) is derived from F=ma by projecting Newtons second law on the trajetory. In Bernoullis equation, we recognize the kinetic and potential energies easily; the pressure term can be interpreted as the mechanical part of internal energy for an incompressible uid, missing in the case of point particles. If these mechanical contributions are subtracted from the rst law of thermodynamics, all that is left is the thermal part of internal energy: so, after mechanical terms are accounted for in agreement with Navier-Stokes, the (total) energy equation reduces to an equation for temperature only, with friction as an exchange term with the mechanical world. We also recover the hydrostatic equation (when V = 0).

5.2.4

Simplied Croccos equation

Rather than asking about the conditions for B to be constant Bernoullis question), we can also wonder what makes B change. We already touched on these topics in Ch. 3, with the viscous losses, and above with the shear layer. Consider steady inviscid (incompressible) isothermal ow. Then u= B (5.13)

Bernoullis equation states that B is constant along streamlines and along vortex lines. If vorticity is aligned with the streamlines (helicity is maximum), B is constant again. Croccos theorem states how B changes from streamline to streamline according to the equation above: the Lamb vector gives the magnitude and direction of B. In 2D ows, where vorticity is perpendicular to the plane of the ow (and therefore to the streamlines), the gradient of B is normal to the streamlines and to the vortex lines. See BatchelorSection 3.5 p.156-160 (where H is enthalpy) for the compressible ow version. See also Currie Section 3.3 p.57. Compare the derivation with viscous losses along streamlines in Ch. 3. This helps us understand what happens to B at the base of a manometer tube (Fig. 5.4) mounted on the wall normal to some ow. While Bernoullis equation, or its variant with losses, may be applied in the ow (along streamlines), and is certainly applicable in the hydrostatic conditions of the manometer, it is denitely not applicable across the shear layer that separates the ow from the manometer (Fig. 5.4). According to Croccos equation, the value of B changes across the shear layer. A similar situation is encountered in boundary layers (Ch 8).

5.2. BERNOULLIS EQUATION

125

Figure 5.3: Croccos theorem for 2-D inviscid ows.

Figure 5.4: The shear layer at the base of a manometer resets the value of the Bernoulli term.

126

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

5.2.5

Applications of Bernoullis equation

Students should ll in this information from suitable undegraduate texts: 1/2 page of text and a gure for each topic, maybe a simple mind-map with relevant topics, would do. Every student at this level should know about these phenomena. Venturi Pitot tube Free surface Siphons Nozzles Cavitation Aerodynamic lift

5.2.6

Lagranges theorem

Before Navier and Stokes reintroduced Newtons viscosity in the momentum balance, the theory of inviscid ow progressed rapidly, and an early theorem by Lagrange is worth mentioning, The theorem states that inviscid irrotational ow remains irrotational forever. The proof is easy. Starting from Eulers equation, cancel the Magnus term, and take the curl to get the evolution of vorticity. t u = t = B=0 (5.14)

This is the Eulerian equivalent of Kelvins theorem (Lagrangian, below); it applies only to irrotational ows, whereas Kelvins theorem has no such restriction.

5.2.7

Creation of vorticity

Under the Boussinesq approximation, it is possible to have density variations in the ow while preserving the simple form the continuity equation. Then p u2 + gz) t u + u = ( + 2 (5.15)

5.2. BERNOULLIS EQUATION

127

Figure 5.5: Creation of vorticity from density gradients and the vorticity equation takes the form t + ( u) = 1 p (5.16)

From Shapiros movie, if density gradient is not parallel to pressure gradient, creation of vorticity in the ow is possible. This remark underscores the importance of the conditions attached to Lagranges and Kelvins theorem (next). We will see other mechanisms of vorticity creation in relation to boundary layers, ow separation and rotating ows.

5.2.8

Kelvins theorem

In Ch. 2 we introduced the circulation and its relation to vorticity = u dr = dA (5.17)

An exact equation was derived by Kelvin under the assumption of inviscid ow of uniform density, under the assumption that the contour of integration is a material line moving with the uid. Let us write the Euler equation in terms of the material derivative 1 dt ui = gi3 i p (5.18)

128

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

Figure 5.6: About Kelvins theorem and multiply by dxi taken along a material line. Then, simple manipulations yield dxi dt ui = dt (ui dxi ) ui dt dxi = dt (ui dxi ) d( u2 ) 2 1 1 gdxi i3 dxi i p = gdx3 dp

(5.19)

where the dierentials of u2 , x3 and p are along the contour. Under the standard assumption that all functions are smooth enough, the integral of these dierentials along the contour must vanish, since we come back to the starting value at the end of integration. Hence dt = 0 (5.20)

We can read the surface integral of vorticity as equivalent to an average vorticity normal to the surface multiplied by the area of this surface. If the product of the two is constant for an arbitrary material line moving with the ow, it means that a reduction in area is associated with an increase in average vorticity, and conversely. This mechanism is very important in a

5.2. BERNOULLIS EQUATION

129

Figure 5.7: Vortex stretching number of applications covered later in this course. Perhaps more importantly, if the material contour initiates its travel in a region where there is no vorticity, the associated particles will remain in irrotational motion as long as viscous eects are negligible. (Note the interesting exception in Ch. 10 where the addition of a weak rotational force (Coriolis) and stretching amplies the creation of weak vorticity). Vortex stretching A dierent formulation of the same idea can be obtained from the vorticity equation. For inviscid ow, the material derivative of vorticity is d t i = j j ui The square vorticity (enstrophy) then obeys the equation dt 2 = j j ui i = j sij i 2 (5.22) (5.21)

because the symmetry of i j cancels the nonsymmetric part of the velocity derivative (see linear algebra). In particular, we can select the local axes to coincide with the principal axes of the local rate of strain. Then for each direction 2 (i) 2 dt = (s(i) (i) ) (no sum) (5.23) 2

130

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

Therefore, following a material point (Lagrangian description), the squared vorticity in each direction (no sum) varies exponentially in time
2 (i) es(i) t

(5.24)

For the component of vorticity long the primary extensional axis, we have an exponential increase; along the compression axis, exponential decrease of square vorticity. Because the latter approaches zero while the former increases very rapidly, vorticity tends to align itself with the extensional axis and increases in magnitude. The idea here is the same as in Kelvins theorem, the presentation quite dierent, the result equivalent.

5.2.9

Helmholtzs vortex theorems

Recall (Ch. 2) Helmholtzs theorems 1. vortex lines cannot end in the uid, they can only end at a boundary. 2. in inviscid incompressible ow, uid elements that lie on a vortex line remain on vortex line (this means that vorticity is a marker just as dye). 3. circulation is the same for all sections of a vortex tube. Theorems 1 and 3 were proved and explained in Ch. 2; theorem 2 invokes inviscid dynamics and can nally be addressed. Consider the vorticity eld, with vorticity vectors distributed through space; consider the associated vortex lines; and nally select the vortex lines supported by some smooth line so as to make a vortex sheet. Then, draw a material contour within this vortex sheet. Since the vorticity is in the sheet, the normal to the sheet is normal to vorticity, and therefore = n dA = 0 (5.25)

for any such contour. By Kelvins theorem, it will remain zero. This implies that, as the arbitrary material contour deforms in the ow, it remains within the vortex sheet. Now, take some support line intersecting the rst one, and the correspoding vortex sheet, and arbitrary contours within the sheet, and reach the same conclusion: the material points remain on the vortex sheet.

5.3. IRROTATIONAL FLOW

131

Figure 5.8: Helmholtz second theorem Finally, since a vortex line is the intersection of two vortex sheets, its material points remain within each of the sheets, therefore remain on the same vortex line. See Panton Section 13.8 p.338, Acheson Section 5.3 p.162-3 for more information. This theorem is very important: as long as eects of viscosity (diusion) and density changes change be neglected, vortex lines act as ow markers, and conversely dye injected along a vortex line will continue to show the vortex line as it moves downstream. This is a good point to view Shapiros movie again (Fig. 5.9) is repeated here from Ch. 2).

5.3

Irrotational ow

Irrotational means that there is no vorticity. Kelvins theorem provides a good understanding for the conditions under which this could happen. Now, let us explore the consequences.

132

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

A.H. Shapiros movie: Vorticity


Kinematics Denition, vorticity meter Potential vortex Vorticity and circulation: Stokes theorem Kelvins theorem: inviscid, Lagrangian, vortex stretching Start-up vortex Dynamics
1 Bernoullis equation: conditions for B = p + 2 V 2 + g z to be constant

Croccos equation: how B changes Potential ow, singular vortices Model ( = 0) Motion of vortex pair, induced velocity Helmholtzs theorems (material lines, vortex tubes, vortex lines end only at boundary) How to introduce vorticity Singularities unsteadiness rotational forces pressure gradient along boundary (viscous eect) See Ch. 9 p and interface

Figure 5.9: Kinematics and dynamics of vorticity and circulation, A. Shapiros movie.

5.3. IRROTATIONAL FLOW

133

Figure 5.10: About We start from = 0, which in turn implies (and conversely) u= Then, incompressibility leads to
2

=0

(5.26)

(5.27)

=0

(5.28)

at every instant. Laplaces equation admits unique solutions if either the eld value (: Dirichlet condition) or its normal derivative (n = un : Neuman condition) is given at every point on the boundary. In the presence of impermeable boundaries, the normal velocity must be equal to that of the boundary (which can be in motion), and the entire eld responds instantaneously to the current conditons. The no-slip condition would require the presence of viscous eects, themselves associated with vorticity. The no-slip condition is incompatible with the Laplace equation for potential ow. As similar picture emerges if we satisfy mass balance rst. Then the vector potential A=u (5.29)

134

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

Figure 5.11: Potential lines and streamfunctions. must be such that The Laplace equation now applies to the vector potential instead of the scalar velocity potential. The picture is consistent: irrotational ow is characterized by unique solutions that reect the boundary conditions throughout the eld at each instant. u= =0=
2

(5.30)

5.3.1

2D potential lines and streamlines

In 2-D ows, it is easy to show that lines of constant and lines of constant are perpendicular to each other. Indeed, we can write d = x dx + y dy = u dx + v dy = 0 so that For the streamlines d = x dx + y dy = v dx + u dy = 0 (5.33) u dy | = . dx v (5.31) (5.32)

5.3. IRROTATIONAL FLOW so

135

dy v | = . dx u dy dy | | = 1, dx dx

(5.34)

Hence

(5.35)

which proves the result. In 2D, it is also convenient to use polar coordinates when the symmetry of the ow makes this representation simpler. For the velocity potential (r, ), we need the components of the gradient: 1 ur = r and u = r Similarly, with A = (0, 0, ), the expression of the curl gives 1 ur = and u = r r (5.37) (5.36)

5.3.2

Superposition of 2D potentials

This material may have been covered at the undergraduate level. This subsection is a list of elementary ow types and a brief discussion of how to use them. The key it to realize that the Laplace equation (for and ) is linear, so that elementary solutions can be added to each other. Uniform ow We take the ow to be uniform with velocity U in the x direction: changes of coordinates can accommodate dierent directions easily. u = U and v = 0 (5.38)

Then, the potential and streamfunctions are easily obtained by integration. (Note that the arbitrary constant of integration can be ignored without loss of generality: addition of the constant does not modify the velocity eld.) = U x and = U y (5.39)

136
3

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS


Potential flow at a stagnation point

2.5

1.5

0.5

0 3

0 x

Figure 5.12: Potential lines (dashed lines) and streamlines (solid lines) of a stagnation ow Stagnation ow For this case = x2 y 2 and = Cxy (5.40) for which the streamlines are branches of hyperbola. For other ows, polar coordinates capture the symmetries more eectively; but of course the choice of coordinates is yours, for most eective solution of the problem at hand. Point source or sink Sign reversal gives a point sink, obviously. Let q be rate of ow coming out of the source, where we place the origin of coordinates. Then q q q y q ln(r) = ln( x2 + y 2 ) and = = atan( 2 ) 2 2 2 2 x + y2 (5.41) Then, the velocity components are = ur = q 1 and u = 0 2 r (5.42)

5.3. IRROTATIONAL FLOW


Potential flow around a source 3

137

3 3

0 x

Figure 5.13: Potential lines and streamlines of a point source Point vortex Dual of the source: swap and above and get y ln(r) = ln( x2 + y 2 ) ) and = = = atan( 2 2 2 2 2 2 x +y (5.43) Then, the velocity components are (5.44) ur = 0 and u = 2r

Source-sink doublet Place a source of strength q at x = a, a sink of equal strength at x = a, and superpose the ows (add the potentials); then take the limit of a 0 at xed value of = aq. The result is a doublet aligned with the x-axis. sin cos and = (5.45) = r r Then, the velocity components are cos sin ur = 2 and u = 2 (5.46) r r It can be checked that streamlines and potential lines are circles.

138

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

Potential flow around a source 3

3 3

0 x

Figure 5.14: Potential lines and streamlines of a point vortex

Potential flow around a doublet 2

1.5

0.5

0.5

1.5

2 2

1.5

0.5

0 x

0.5

1.5

Figure 5.15: Potential lines and streamlines of a doublet

5.3. IRROTATIONAL FLOW


Potential flow around a Rankine halfbody 3

139

3 3

0 x

Figure 5.16: Potential lines and streamlines of a Rankine half-body Rankine half-body Superpostion of a uniform ow and a source = U r cos + q q ln(r) and = U r sin 2 2 (5.47)

Then, the velocity components are ur = and u = (5.48)

Cylinder Superposition of a doublet and a uniform ow cos sin and = U r sin r r Then, the velocity components are = U r cos + ur = U cos cos sin and u = U sin 2 2 r r (5.49)

(5.50)

140

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

Potential flow around a cylinder 3

3 3

0 x

Figure 5.17: Potential lines and streamlines of the ow around a cylinder

Potential flow around a cylinder 3

3 3

0 x

Figure 5.18: Potential lines and streamlines for a cylinder with circulation

5.3. IRROTATIONAL FLOW Cylinder with circulation

141

Superposition of a doublet, a vortex and a uniform ow This to be lled in.

5.3.3

F=ma!?

The previous subsections show that, for suitable boundary conditions, one does obtain unique solutions for the velocity eld by satisfying mass balance and the kinematic constraint of irrotational ow. How about F = ma? Can we have a solution without momentum balance? Answer in class, after you think it through!

5.3.4

DAlemberts paradox

It is easy to see that the streamline = 0 corresponds to a circle of radius R = /U . Applying Bernoullis equation between innity and the cylinder surface, we have (u2 + u2 ) U 2 r + p = + p |R (5.51) 2 2 Taking p = 0 as reference pressure (why is this OK?) we get the surface pressure (Fig. 5.19) as p |R = U 2 (cos 2 1) 2 (5.52)

This is maximum at the front and back stagnation points = 0 and , and minimum at = /2. Then it is easy to see that, for a surface element of area Rd, the components of force are dFx = Rd cos pR (5.53) and dFy = Rd sin pR By integration, the drag D and lift L are D= and L= dFy = 0 (5.56) dFx = 0 (5.55) (5.54)

142

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

Figure 5.19: Pressure force on the cylinder Vanishing lift is clearly the result of symmetry, broken by adding circulation. But the vanishing drag has much deeper roots. In fact, the absence of drag was proved by DAlembert for bodies of any shape whatsoever: the mismatch between this elegant result (and the vindication of potential external ows as valid models within the accuracy achievable then) and the common experience of drag was called DAlemberts paradox. It started a profound schism in the uids community, lasting for much of the nineteenth century, between the aerodynamicists (who, with Joukovky and others, took potential ows into higher mathematics) and hydraulics practitioners (who relied more on empirical data). The paradox was nally solved by Prandtl, with the discovery of the boundary layer and its profound eect on the ow (skin friction, boundary layer separation, wake formation...)

5.4

Viscous potential ows

We will see in Ch. 9 the role played by viscosity to introduce vorticity from solid boundaries into many ows. This ubiquitous complication blinded the uid mechanics community to the interest of viscous potential ows. Combining potential ow with incompressibility gives
2

u=0

(5.57)

5.5. ADVANCED TOPICS AND IDEAS FOR FURTHER READING 143 so that, although the viscous stress does not vanish, the viscous force does. For additional information, see the recent work of D.D. Joseph.

5.5

Advanced topics and ideas for further reading

Batchelors entire Ch.6 is devoted to inviscid ows with vorticity: good reading! In his Ch.5, Batchelor also covers the eect on viscosity on Kelvins theorem. In relation of Helmholtzs interpretation of vortex lines with material lines in inviscid ow, the eect of viscosity on vortex line topology (break-up and reconnections) has also been studied extensively. Complex coordinates, conformal mapping for 2D ows: see Acheson for a good overview, Panton or Currie for more details. Curries Ch. 5 is devoted to 3D potential ows. Kelvins theorem is a particular of the Poincar-Cartan invariant in Hamile tonian dynamics. There are many points of overlap between irrotational ows and Hamiltonian dynamics Finally, Onsager pointed out the possibility of having energy dissipation in inviscid ows: this requires that singularities (innite derivatives) develop as a result of vrotex stretching. This cannot happen in 2D, but the issue is still unresolved in 3D.

Problems
1. Categorize the ideas in this chapter as being for inviscid ow, irrotational ow, or both. Explain relations between these ideas. 2. Discuss (on the basis of equations) the following proposition: any ow that is irrotational is also inviscid (the converse is clearly not true). 3. Derive the equation of the free surface associated with a potential vortex (next: add a sink at the origin); discuss in what respects this is or is not a good model for the bathtub vortex. 4. Work out the pressure distribution eq. (5.52) for the potential ow around a cyclinder, and integrate it to get drag and lift; repeat with the cylinder with circulation (add a vortex).

144

CHAPTER 5. INVISCID FLOWS AND IRROTATIONAL FLOWS

S-ar putea să vă placă și