You are on page 1of 10

CHAPTER

General Conservation Equations Revisited: Velocity Potential Equation



Dynamics of compressible fluids, like other subjects in which the nonlinear character of the basic equations plays a decisive role, is far from the perfection envisaged by Laplace as the goal of a mathematical theory.

Richard Courant and K. O. Friedrichs, 1948

303

304

CHAPTER 8 General Conservation Equations Revisited: Velocity Potential Equation

8.1 I INTRODUCTION

In this chapter, the genera) conservation equations derived in Chap. 6 are snnpune for the special case of irrotational flow, discussed below. This simplification is dramatic; it allows the separate continuity, momentum, and energy equations the requisite dependent variables p, p, Y, T, etc., to cascade into one n ... ·""rn'ft equation with one dependent variable-a new variable defined below as the potential. In tbis chapter, the velocity potential equation will be derived; in tum, Chap. 9 it will be employed for the approximate solution of several important lems in compressible flow.

8.2 I IRROTATIONAL FLOW

The concept of rotation in a moving fluid was introduced in Sec. 6.6. The vorticity a point property of the flow, and is given by \l x Y. Vorticity is twice the velocity of a fluid element, \l x V = 200. A flow where \l x V 1= 0 throughout calJed a rotational flow. Some typical examples of rotational flows are illustrated

l

8.2 Irrotational Row

305

M~>J -

Viscous flow inside a bOM dary la y er

Inviscid flow behind a curved shock wave

Ftgure 8.1 ! Examples of rotational flows.

M~>l

Flowfleld over a sharp wedge Or cone

'Vxv=o

Two-dimensional or axisymmetric nozzle flows

Flowfield behind the shock wave on a slender, sharp nosed body is almost irrotational. For analysis, we usually assume 'V x V = 0 for this case,

Figure 8.2 J Examples of irrotational Haws.

8.1 for the region inside a boundary layer and the inviscid flow behind a curved wave (see Sec. 6.6). In contrast, a flow where \7 x V = 0 everywhere is called irrotational flow. Some typical examples of irrotational flows are shown in

8.2 for the flowfield over a sharp wedge or cone, the two-dimensional or axflow through a nozzle, and the flow over slender bodies. If the slender is moving supersonically, the attendant shock wave will be slightly curved, and

306

CHAPTER 8 General Conservation Equations Revisited: Velocity Potential Equation

hence, strictly speaking, the flow field will be slightly rotational. However, it is usually practical to ignore this, and to assume V x V ~ 0 for such cases.

Irrotational Bows are usually simpler to analyze than rotational flows; the irrotationality condition V x V = 0 adds an extra simplification to the general equations of motion. Fortunately, as exemplified in Fig. 8.2, a number of practical flowfields can be treated as irrotational. Therefore, a study of irrotatioual flow is of great practical value in fluid dynamics.

Consider an irrotational flow in more detail. In cartesian coordinates, the mathematical statement of irrotational flow is

j k
VxV= a 0 a
ax oy oz
u v w = i (aw _ av) _ j (~ w _ au) + k (av _ au) = 0

a y oz ax az ax oy

For this equality to hold at every point in the flow,

av au

ax aY

ow av

oy az

ow au ax = oz

Equations (8.1) are caned the irrotationality conditions. Now consider Euler's equation [Eq. (6.29)J without body forces.

DV

p Dt = -Vp

For steady flow, the x component of this equation is

au au au Bp

pu- +pv- +pw- =--

ax oy az ax

ap au au au

-- dx = pu- dx + pu- dx + pw- dx

ax ax ay OZ

or

But from Eq. (8.1),

au OU

=

oy ax

and

au aw

-=-

az ox

Substituting the above relations into Eq. (8.2), we have

op au av ow

-- dx = pu- dx + pv- dx +pw- dx

h ax h h

op I au2 I av2 1 aw2

-- dx = -p-dx+ -p-dx+ -p- dx

ax 2 ax 2 ox 2 ox

or

(8.1)

(8.2)

(8.3)

8.2 Irrotational Flow

307

Similarly, by considering the y and z components of Euler's equation,

8p 1 Ju2 1 Bv2 1 aw2

--dy= -p-dy+-p-dy+ -p-dy

By 2 By 2 ay 2 ay

(8.4)

op 1 au2 1 ov2 I aw2

-- dz = -p-dz+ -p- dz+ -p- d :

vz 2 az 2 8z 2 8z

(8.5)

Adding Eqs. (8.3) through (8.5), we obtain

(BP 8p op) 1 ay2 1 av2 I av2

- -dx+-dy+-dz =-p-dx+-p-dy+-p-dz

ax oy oz 2 ax 2 By 2 OZ

(8.6)

where y2 = u2 + v2 + w2.

Equation (8.6) is in the form of perfect differentials, and can be written as

or

-dp = ~p d(y2) I dp=-pVdV I

(8.7)

Equation (8.7) is a special form of Euler's equation which holds for any direction throughout an irrotational inviscid flow with no body forces. If the flow were rotational, Eq. (8.7) would hold only along a streamline. However, for an irrotational How, the changes in pressure dp and velocity dV in Eq. (8.7) can be taken in aoy direction, not necessarily just along a streamline.

Euler's equation embodies one of the most fundamental physical characteristics of fluid flow-a physical characteristic that is easily seen in the form given by Eq. (8.7). Namely, in an inviscid flow if the pressure decreases along a given direction [dp is negative in Eq. (8.7)], the velocity must increase in the same direction [in Eq. (8.7), dY must be positive]; similarly, if the pressure increases along a given direction [dp is positive in Eq. (8.7)], the velocity must decrease in the same direction [in Eq. (8.7), dV must be negative). In the popular literature this is sometimes called the "Bernoulli principle" because in the early eighteenth century Daniel Bernoulli observed this physical effect Although he worked hard to properly quantify it, he was unsuccessful. His friend and colleague, Leonard Euler, was the first to obtain the proper quantitative relation, namely Eq. (8.7). This equation dates from 1753. (See Reference 134 for more historical details on Bernoulli and Euler, and their contribution to fluid dynamics.)

The Bernoulli principle is very easy to understand physically. Consider a fluid element moving with velocity Y in the s direction as sketched in Fig. 8.3. If the pres~ sure decreases in the s direction as shown in Fig. 8.3a (this is defined as e favorable pressure gradient), the pressure on the left face will be higher than that on the right face, exerting a net force on the fluid element acting toward the right, and hence

308

CHAPTER 8 General Conservation Equations Revisited: Velocity Potential Equation

Pressure decreases in the s direction, thus accelerating the fluid element towards the right.

Pressure increases in the s direction, thus decelerating the

fluid element

v -

v
-
p .....
Net force
ds
I
(C/) p +dp

(dp is negau vel

p + ell'

p

....

Nel force

(dp is positive)

--- cis -----+1.1

~------------~s

~------------+ s

(b)

Figure 8.3 I Illustration of pressure gradient effect on the velocity of a fluid element. (a) Decreasing pressure in the flow direction increases the velocity. (b) Increasing pressure in the flow direction decreases the velocity,

accelerating it in the s direction. Clearly, in a region of decreasing pressure, the fluid element will increase its velocity. Conversely, if the pressure increases in the s direction as shown in Fig. 8.3b (this is defined as an adverse pressure gradient), the pressure on the light face will be higher than that on the left face, exerting a net force on the fluid element acting toward the left. and hence decelerating it in the s direction. Clearly, in a region of increasing pressure. the fluid element will decrease its velocity.

8.31 THE VELOCITY POTENTIAL EQUATION

Consider a vector A. If V x A = 0 everywhere, then A can always be expressed as V ~, where ~ is a scalar function. This stems directly from the vector identity, curl (grad) == O. Hence,

V x V~ = 0

where ~ is any scalar function. For irrotational flow, V x V = O. Hence, we can define a scalar function, ~ = ~(x, y, z), such that

where <I> is called the velocity potential. In cartesian coordinates, since

v = ui + vj + wk

and

a<p a~ a<p

V<P = -i+ -j +-k

ax By· oz

(8.8)

8.3 The Velocity Potential Equation

then, by comparison,

a<p atll atll

u=- v=- w=-

ax ay az

Hence, if the velocity potential is known, the velocity can be obtained directly from Eq. (8.8) or (8.9).

As derived next, the velocity potential can be obtained from a single partial differential equation which physically describes an irrotational flow. In addition, we will assume steady, isentropic flow. For simplicity, we will adopt subscript notation for derivatives of <P as follows: atlljax == 1fJ .. , atlljay == «1>)", aipjaz == tIlz' etc. Thus, the continuity equation, Eq. (6.5), for steady flow becomes

(8.9)

'V. (pV) = 0

o(pu) i1(pv) 8(pw)

--+--+--=0

ax oy oz

o a. a.

-,--- plfJ..I' + -;- plfJy + - p<l>t = 0

ax oy oz

. op op. ap

p(<P .... + «l>yy + til,;:) + IfJx- + cJ>y- + <Pz- = 0

ax ay GZ

(8.10)

Since we are striving for an equation completely in terms of <JI, we eliminate p from Eq. (8. J 0) by using Euler's equation in the form ofEq. (8.7), which for an irrotarional flow applies in any direction:

(8.11)

Prom the speed of sound, a2 = (op/8p)s. Recalling that the flow is isentropic, any change in pressure dp in the flow is followed by a corresponding isentropic change in density, dp. Hence,

dp (OP) = (/2

dp = 8p ,

dp

dp=-

a2

(8. L2)

Combining Eqs. (8.11) and (8.12):

p . (til; + <P~ + <P~)

dp = -- d .

a2 2

(8.13)

309

(S.lS)

310

CHAPTER 8 General Conservation Equations Revisited: Velocity Potential Equation

Considering changes in the x direction, Eq. (S .13) directly yields

or

(S.14)

Similarly,

(8.15)

(8.16)

Substituting Eqs. (8.14) through (8.16) into Eq. (8.10), canceling the p that appears in each term, and factoring out the second derivatives of q." we have

(8.17)

Equation (8. J 7) is called the velocity potential equation.

Equation (8.17) is not strictly in terms of ({> only; the variable speed of sound a still appears. We need to express a in terms of .p. From the energy equation, Eq. (6.45),

ho = const

Hence, for a calorically perfect gas, this equation can be expressed as V2

CpT + 2 =cpTo

yRT V2 yRTo

--+-=--

y-1 2 y-l

a2 V2 a2

_'_+_= __ 0_

y-l 2 y-l

? ? Y - 12 2 Y - I? 2 2

a: = a~ - -- V = a - -- (u- + v + w )

o 2 0 2

8.3 The Velocity Potential Equation

311

Since Qo is a known constant of the flow, Eq. (8.18) gives the speed of sound Q as a function of ell.

In summary, Eq, (8.17) coupled with Eq. (8.18) represents a single equation for the unknown variable W. Equation (8.18) represents a combination of the continuity, momentum, and energy equations. This leads to a genera) procedure for the solution of irrotational, isentropic flowfields:

1. Solve for W from Eqs. (8.17) and (8.18) for the specified boundary conditions

of the given problem.

2. Calculate u, v, and w from Eq. (S.9). Hence, V = ../u2 +v2 + w2•

3. Calculate Q from Eq. (8.18).

4. Calculate M = Via.

5. Calculate T, p, and p from Eqs. (3.28), (3.30), and (3.31) respectively.

Hence, we see that once «fJ = W(x, y, z) is obtained, the whole fiowfield is known. This demonstrates the importance of ell.

Note that Eq. (8.17) combined with (8.1S) is a nonlinear partial differential equation. It applies to any irrotational, isentropic flow: subsonic, transonic. supersonic, or hypersonic. It also applies to incompressible flow, where a ~ 00, hence yielding the familiar Laplace's equation,

~.u + <l>yy + CIlzz = 0

Moreover, the combined Eqs. (8.17) and (8.18) is an exact equation within the framework of isentropic, irrotational flow. No mathematical assumptions (such as small perturbations) have been applied at this stage of our presentation. There is no general closed-form solution to the velocity potential equation, and hence its solution is usually approached in one of these ways:

1. Exact numerical solutions. This approach makes it difficult to formulate general trends and rules-the results are raw numbers which have to be analyzed, just like experimental data obtained in the laboratory. However, the techniques of modem computational fluid dynamics are rendering numerical solutions as everyday occurrences in compressible flow, allowing solutions to complicated applications where there would ordinarily be no solution at all. We will study aspects of computational fluid dynamics in Cbaps. 11, 12, and 17, emphasizing methods of characteristic and finite-difference solutions.

2. Transformation of variables in order to make the velocity potential equation linear, but still exact. Examples of this approach are scarce. One such method is the bodograph solution for subsonic flow, as described by Shapiro (see Ref. 16). Due to its limited usefulness, this technique will not be considered here.

3. Linearized solutions. Here, we find linear equations that are approximations to the exact nonlinear equations, but which lend themselves to closed-form analytic solution. A large number of real engineering problems lend themselves

312

C HAP T E R 8 General Conservation Equations Revisited: Velocity Potential Equation

to reasonable approximations which linearize the velocity potential equation. Aerodynamic theory historically abounds in linearized theories. This wiU be the subject of Chap. 9.

8.4 I HISTORICAL NOTE: ORIGIN OF THE CONCEPTS OF FLUID ROTATION AND VELOCITY POTENTIAL

The French mathematician Augustin Cauchy, famous for his contributions to partial differential equations and complex variables, was also active in the theory of fluid flow. In a paper presented to the Paris Academy of Sciences in 1815. he introduced the average rotation at a point in the flow. The extension of this idea to the concept of instantaneous rotation of a fluid element was made by the Englishman George Stokes at Cambridge in 1847. (See Fig. 8.4.) In a paper dealing with the viscous flow of fluids. Stokes was the first person to visualize the motion of a fluid element as the resolution of three components: pure translation, pure rotation, and pure strain. The concept of rotation of a fluid element was then applied to inviscid flows about J 5 years later by Hemann von Helmholtz.

Figure 8.4 I Sir George Stokes (1819-1903).

You might also like