Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

31
Chapter 6 Conservation of Momentum: Fluids and Elastic Solids The description of the motion of fluids plays a fundamental role in applied mathematics. The basic model is a system of partial dierential equations of evolution type. 6.1 Equations of Fluid Motion Fluids consists of molecules; thus, on a microscopic level, a fluid is a discrete material. To obtain a useful approximation, we will describe fluid motion on the macroscopic level by taking into account forces that act on a parcel of fluid, which we assume to be a collection of suciently many molecules of the fluid so that the continuity assumption is valid. More precisely, we will assume the fluid to be a continuous medium contained in three-dimensional space R 3 such that every parcel of the fluid, no matter how small in compar- ison with the whole body of fluid, can be viewed as a continuous material. Mathematically, a parcel of fluid (at every moment of time) is a bounded open subset A of the fluid that is assumed to be in an open set R whose closure is the region containing the fluid. To ensure correctness of the math- ematical operations that follow, we assume that each parcel A has a C 1 boundary. At every given moment of time, a particle of fluid is identified with a point in R. Let = (x, t) denote the density and u = u(x, t) the velocity of the fluid at position x 2 R and time t 2 R. The motion of the fluid is modeled by 257

Transcript of Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Page 1: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Chapter 6

Conservation of Momentum:Fluids and Elastic Solids

The description of the motion of fluids plays a fundamental role in appliedmathematics. The basic model is a system of partial di↵erential equations ofevolution type.

6.1 Equations of Fluid Motion

Fluids consists of molecules; thus, on a microscopic level, a fluid is a discretematerial. To obtain a useful approximation, we will describe fluid motion onthe macroscopic level by taking into account forces that act on a parcel offluid, which we assume to be a collection of su�ciently many molecules ofthe fluid so that the continuity assumption is valid. More precisely, we willassume the fluid to be a continuous medium contained in three-dimensionalspace R3 such that every parcel of the fluid, no matter how small in compar-ison with the whole body of fluid, can be viewed as a continuous material.

Mathematically, a parcel of fluid (at every moment of time) is a boundedopen subset A of the fluid that is assumed to be in an open set R whoseclosure is the region containing the fluid. To ensure correctness of the math-ematical operations that follow, we assume that each parcel A has a C1

boundary. At every given moment of time, a particle of fluid is identifiedwith a point in R.

Let ⇢ = ⇢(x, t) denote the density and u = u(x, t) the velocity of the fluidat position x 2 R and time t 2 R. The motion of the fluid is modeled by

257

Page 2: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

258 Fluids

di↵erential equations for ⇢ and u determined by the fluid’s internal materialproperties, its container, and the external forces acting on the fluid.

The components of u with respect to the usual coordinates of three-dimensional space are denoted by (u

1

, u2

, u3

). We assume that the functionu is twice continuously di↵erentiable and satisfies other properties that arenecessary for the correctness of the mathematical operations used in thefollowing discussion. We will make remarks about the additional propertiesof u as needed.

The changing position of the particle of fluid with initial position x0

2 Rat time t = 0 is given by the curve t 7! �(x

0

, t) in R that is the solution ofthe initial value problem (IVP)

⇠ = u(⇠, t), ⇠(0) = x0

(6.1)

(see Appendix A.3 for a theorem on existence of solutions of ordinary dif-ferential equations). When � is viewed as a function of two variables � :R3⇥R ! R3 it is called the flow of u. Note that the notation �(A, t), whereA is a subset of R3, is the set of all points in R3 obtained by solving to timet the di↵erential equation with initial condition ⇠(0) = x

0

for each x0

in A.For each such initial point, the value of the solution at time t (a point in R3)belongs to the set �(A, t).

Let x1

, x2

, and x3

denote the Cartesian coordinates in R3 and e1

, e2

, e3

the usual unit direction vectors. Using this notation, the velocity vector u isu = u

1

e1

+ u2

e2

+ u3

e3

.The gradient operator in Cartesian coordinates (also called nabla or del)

is

r := e1

@

@x1

+ e2

@

@x2

+ e3

@

@x3

or, equivalently,

r :=

0

B@

@@x1

@@x2

@@x3

1

CA .

This operator applied to a function f : R3 ! R gives its gradient in Cartesiancoordinates

rf =

0

BB@

@f@x1

@f@x2

@f@x3

1

CCA .

Page 3: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 259

Applied to a vector field u (with the notation r ·u, where · denotes the usualinner product in Euclidean space), the gradient operator gives the divergencein Cartesian coordinates

r · u =@u

1

@x1

+@u

2

@x2

+@u

3

@x3

.

The expression (u ·r)u, often written u ·ru, is the vector

0

@u1

@u1@x1

+ u2

@u1@x2

+ u3

@u1@x3

u1

@u2@x1

+ u2

@u2@x2

+ u3

@u2@x3

u1

@u3@x1

+ u2

@u3@x2

+ u3

@u3@x3

1

A .

A parcel A of fluid identified at time zero is moved to �(A, t) at time t.Reynolds’s transport theorem states that

d

dt

Z

�(A,t)

⇢(x, t) dx =

Z

�(A,t)

⇢t(x, t) + div(⇢u)(x, t) dx (6.2)

(see A.11). By conservation of mass, the rate of change of the total mass inA does not change as the parcel is transported by the flow; therefore, theleft-hand side of equation (6.2) vanishes. Since A may be taken arbitrarilysmall (for example, a ball with arbitrarily small radius), it follows that

⇢t + div(⇢u) = 0 (6.3)

or, equivalently,⇢t +r · (⇢u) = 0. (6.4)

Equation (6.4) (or (6.3)) is called the equation of continuity; it statesthat the mass of a fluid is conserved by the fluid motion. Equation (6.2) is ageneral statement of the rate of change of total mass that holds as long as uis an arbitrary (smooth) vector field with flow �.

A di↵erential equation for the velocity field u is obtained from the equa-tion for the conservation of momentum using Newton’s second law of motion.Note that a fluid has mass. It might also be charged. Thus, a fluid is sub-jected to body forces, which by definition are forces that act per unit of massor per unit of charge. The most important body force is the gravitationalforce, which acts on every fluid simply because fluids have mass. Unlike themotion of particles or rigid bodies, fluids (by definition) have internal stress

Page 4: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

260 Fluids

(force per area) that is caused by the action of the fluid on itself. Stress ismodeled by a function � that assigns a vector in R3 to each pair consistingof a point (x, t) in space-time and a unit-length (space) vector ⌘ in R3 atthis point. The value of the stress function at this pair is called the stressvector at the point x in the direction of the outer normal ⌘ at time t on imag-inary surfaces in space that contain this point and have this outer normalat time t. The stress vector has units of force per area. Using conservationof momentum and angular momentum, it is possible to prove that the stressfunction at each point in space-time is a symmetric linear transformation ofspace (see [17]). Let us simply incorporate these facts about the stress asassumptions.

A linear and symmetric transformation on vectors may be representedby a (diagonalizable) matrix in the usual Cartesian coordinates. Thus, foreach point (x, t) in space-time, �(x, t) may be viewed as a symmetric matrix,which is thus defined by six numbers (the elements on and above the maindiagonal of the matrix).

Total stress over the current position of parcel A at time t is given by

TS :=

Z

@�(A,t)

�(x, t)⌘(x, t) dS,

where ⌘ is the outer unit normal on the boundary of �(A, t) and dS is theelement of surface area. Using the body force b per unit of mass, the totalbody force on �(A, t) is

TB :=

Z

�(A,t)

⇢(x, t)b(x, t) dV ,

where dV is the element of volume.The total momentum of �(A, t) is

Z

�(A,t)

⇢(x, t)u(x, t) dV .

By Newton’s second law of motion (the time rate of change of momen-tum on a body is equal to the sum of the forces acting on the body), themathematical expression for the conservation of momentum is

d

dt

Z

�(A,t)

⇢(x, t)u(x, t) dV =

Z

@�(A,t)

�(x, t)⌘(x) dS +

Z

�(A,t)

⇢(x, t)b(x, t) dV .

(6.5)

Page 5: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 261

The region �(A, t) in space is moving with time. Using the equation ofcontinuity (6.3), the transport theorem (A.11), and some algebra, it followsthat the left-hand side of the momentum balance (6.5) is given by

d

dt

Z

�(A,t)

⇢(x, t)u(x, t) dV =

Z

�(A,t)

⇢(x, t)(ut + (u ·r)u)(x, t) dV , (6.6)

where ut denotes the partial derivative of u with respect to t.The expression ut + (u ·r)u that appears in the right-hand side of equa-

tion (6.6) is the material derivative of the velocity field u, which is often

denoted byDu

Dt(x, t); its definition takes into account the motion of the fluid

particles with time:

d

dtu(�(t, x

0

), t) = ut(�(t, x0

), t) +Du(�(t, x0

), t)�(t, x0

)

= ut(�(t, x0

), t) +Du(�(t, x0

), t)u(�(t, x0

), t)

= (ut + (u ·r)u)(�(t, x0

), t)

=Du

Dt(�(t, x

0

), t).

The total stress in the direction of an arbitrary constant vector field v inR3 is given by Z

@�(A,t)

(�(x, t)⌘(x)) · v dS.

Using the symmetry of the stress tensor, we have

(�⌘) · v = (�v) · ⌘.

This fact together with the divergence theorem implies

Z

@�(A,t)

(�(x, t)⌘(x)) · v dS =

Z

�(A,t)

div(�(x, t)v) dV .

By an easy calculation in components with � represented as a matrix and va constant vector, we have the identity

r(�v) = (r · �)v. (6.7)

Page 6: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

262 Fluids

By our notation, div(�v) = r(�v). Thus, we may view r ·� (the divergenceof �), as a linear operation on vectors. In Cartesian coordinates, this operatoris given by

r · � =

0

@@�11@x1

+ @�12@x2

+ @�13@x3

@�21@x1

+ @�22@x2

+ @�23@x3

@�31@x1

+ @�32@x2

+ @�33@x3

1

A .

Using these facts, the total stress is

TS =

Z

@�(A,t)

�(x, t)⌘(x) dS =

Z

�(A,t)

r · �(x, t) dV . (6.8)

By substitution of formulas (6.6) and (6.8) into the conservation of mo-mentum equation (6.5), we have the equivalent integral expression

Z

�(A,t)

�⇢Du

Dt�r · � � ⇢b

�dV = 0 (6.9)

for conservation of momentum. This equation holds for every parcel of fluid.Under the assumption that the integrand is continuous (which might bea strong assumption in some circumstances), the di↵erential equation forconservation of momentum is

⇢Du

Dt= r · � + ⇢b; (6.10)

or, equivalently, Cauchy’s equation

⇢(ut + (u ·r)u) = r · � + ⇢b. (6.11)

There is a third conservation law: the conservation of energy. It is re-quired, for example, in case temperature changes are important (see, forexample, [43]). But, for simplicity, we will treat only situations where thislaw can reasonably be ignored.

Equation (6.11) holds for arbitrary elastic media, not just fluids. Thismodel would be exact if internal stress could be described exactly by a sym-metric tensor that was derived without additional assumptions from elec-tromagnetism. While a fundamental representation of internal stress mightbe possible in principle, it would likely be so complex as to be useless formaking predictions. Instead, an approximation—based on physical princi-ples and physical intuition—called a constitutive law for the type of fluid

Page 7: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 263

(or elastic material) being modeled is used to define the stress tensor as afunction of the other state variables. The choice of constitutive stress lawis the most important ingredient in a viable model. Since constitutive lawsare approximations, predictions derived from corresponding models must bevalidated by physical experiments.

Perhaps the simplest model for internal stress arises from the constitutiveassumption that the stress is the same in all directions (no shear stress). Inthis case, there is a scalar function p on space-time, called the pressure, suchthat the stress tensor is given by

�(x, t) = �p(x, t)I, (6.12)

where I denotes the identity transformation on space. The minus sign istaken because the stress on the boundary of a fluid parcel is �⌘, where ⌘ isthe outer unit normal on the boundary of the parcel and a positive pressure(from outside the parcel) is assumed to act in the direction of the innernormal. Indeed, pressure is correctly defined to be the normal component offorce per area. A fluid that satisfies constitutive law (6.12) is called an idealfluid. Using formula (6.7), the divergence of the stress for an ideal fluid is

r · � = �rp,

and the corresponding equations of motion

⇢(ut + (u ·r)u) =�rp+ ⇢b,

⇢t +r · (⇢u) = 0. (6.13)

System (6.13) is four equations, counting the three components of thevector equation for conservation of momentum, for five unknowns: the threecomponents of the velocity u, the density ⇢ and the pressure p. One moreequation is needed to close the system. The missing ingredient is conser-vation of energy. A physically realistic approach to conservation of energyrequires a digression into thermodynamics. While the kinetic energy 1

2

⇢kuk2is a quantity from classical mechanics, models of internal energy require adeeper analysis that is not completely understood. Thus, the most generalform of the equations of fluid dynamics remain somewhat controversial. For-tunately, for some practical applications, internal energy considerations canbe bypassed or modeled with relatively simple constitutive laws. Anotherapproach to closing the system is to make a simplifying assumption on the

Page 8: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

264 Fluids

nature of the fluid. Two standard (closely related) possibilities are the as-sumptions that the fluid is incompressible (that is, the flow preserves volume)or that the density of the fluid is constant.

The incompressibility assumption is equivalent to assuming the velocityfield is divergence free: div u = 0. In this case, using the identity

r · (⇢u) = r⇢ · u+ ⇢ru,

the (five scalar) equation of motion

⇢(ut + (u ·r)u) =�rp+ ⇢b,

⇢t +r⇢ · u= 0, (6.14)

r · u= 0. (6.15)

are called Euler’s equations for an ideal fluid.A more restrictive assumption is constant density; it leads to the (four)

equations

⇢(ut + (u ·r)u) =�rp+ ⇢b,

r · u= 0 (6.16)

for the four unknowns for the components of u and the pressure.Because our fluid is confined to a region R of space, boundary conditions

must be imposed. Physical experiments show that the correct boundarycondition is u ⌘ 0 on the solid boundaries of R. This is called the no-slipboundary condition. To demonstrate this fact yourself, consider cleaning ametal plate by using a hose to spray it with water; for example, try cleaninga dirty automobile. As the pressure of the water increases, the size of theparticles of dirt that can be removed decreases. But, it is very di�cult toremove all the dirt by spraying alone. This can be checked by polishingwith a clean cloth. In fact, the velocity of the spray decreases rapidly in theboundary layer near the plate. Dirt particles with su�ciently small diameterare not subjected to flow velocities that are high enough to dislodge them.

Euler’s equations are not well-posed with the no-slip boundary condition;that is, under this boundary condition, the equations of motion do not haveunique solutions depending continuously on the initial position of the fluid.The mathematically correct boundary condition for Euler’s equations is thatthe fluid does not penetrate the boundary; or, equivalently, the fluid velocityis everywhere tangent to the boundary. The no-slip condition is allowed,

Page 9: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 265

but not required. While this fact implies that Euler’s equations cannot bethe correct model for physical fluids, Euler’s equations give experimentallyverifiable predictions as long as measurements are taken far from the fluid’sboundary. Indeed, this observation is fundamental in many applications (forexample, flow over an airplane wing) where there is a thin layer of fluidnear the boundary that must obey a more realistic stress constitutive law inorder to satisfy the no-slip boundary condition; but, away from this layer,the motion of the fluid is well-approximated by Euler’s equations.

To obtain a more realistic model, the viscosity of the fluid must be takeninto account. The modeling process requires several assumptions, which aredi↵erent for di↵erent types of fluids. For fluids similar to water and air, calledNewtonian fluids, the main assumptions are the isotropy of the fluid (thatis, the fluid is the same in all directions) and a linear relation between stressand velocity. For some fluids, for example blood, accurate models requirenonlinear stress-velocity relations.

The modeling process to obtain the stress-velocity relation for Newtonianfluids is described in books on fluid mechanics (see, for example, [17, 43]). Adiscussion and derivation of this stress-strain relation in the context of theequation of motion for elastic media is presented in Section 6.13. The sameunderlying ideas lead to stress tensors for fluids. The stresses are all given bypressure (due to molecular motion). This part of the stress is modeled by theEulerian fluid where � = �pI. Stresses in moving fluids are determined byconstitutive laws, usually formulated as linear relations between stress andstrain; that is, Hooke’s law. Strains are relative changes in length. Theyare measured by the (infinitesimal) changes in positions of material pointsof the fluid, which is modeled by detecting the stretching and compressionof vectors (representing directions) at points in the fluid. Strain, the relativechange in length, is dimensionless.

Let p be a point in the fluid and v a unit vector at p. The motion of p inthe moving fluid starting at some fixed time T with velocity field u is givenby t 7! �(p, t), where

@�

@t(p, t) = u(�(p, t), t) (6.17)

and �(p, T ) = 0. The temporal variable t measures time starting at time T .The displacement of p after time t is given by �(p, t) := �(p, t)� p. Let ↵ bea curve at p parameterized by s with tangent vector v; that is, s 7! ↵(s) isa curve such that ↵(0) = p and ↵0(0) = v. The fluid flow moves each point,hence the curve tangent to v. The infinitesimal distortion of the curve is the

Page 10: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

266 Fluids

directional derivative

d

ds�(↵(s), t)

��s=0

= D�(p, t)v,

where D denotes the derivative of the transformation p 7! �(p, t) for fixed t.The length of ↵ starting at p is

`(s) :=

Z s

0

|↵0(⌧)| d⌧,

and the length along the distorted curve is

L(s) :=

Z s

0

|D�(↵(⌧), t)↵0(⌧)| d⌧.

The di↵erence of the squares of these lengths is easily computed using theTaylor approximation to be

L2(s)� `2(s) = (|D�(p, t)v|2 � |v|2)s2 +O(s3)

= (hD�(p, t)v,D�(p, t)vi � hv, vi)s2 +O(s3)

= h(D�(p, t)TD�(p, t)� I)v, vis2 +O(s3).

To take into account the size of the displacement, the linear transformationD�(p, t)TD�(p, t)� I is recast in the form

D�(p, t)TD�(p, t)� I = (Du(p) + I)T (Du(p) + I)� I

=Du(p)TDu(p) +Du(p)T +Du(p).

The Lagrange-Green (strain) tensor E is defined by

E(p, t)(v, w) =1

2h(D�(p, t)TD�(p, t)� I)v, wi

=1

2h(D�(p, t)TD�(p, t) +D�(p, t)T +D�(p, t))v, wi,

where the factor 1

2

is inserted to agree with the usual definition of this tensor(which is used in Section 6.13) and w is a vector at p. A displacement (ordeformation) is called small if the product D�(p, t)TD�(p, t) may be safelyneglected. From a mathematical perspective, the same result is achieved by

Page 11: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 267

simply declaring the linear strain tensor " to be the linear approximation ofthe Lagrange-Green tensor; that is, the linear strain tensor is given by

"(p, t)(v, w) =1

2h(D�(p, t)T +D�(p, t))v, wi. (6.18)

The spatial derivative of the fluid motion is determined by di↵erentiationof the di↵erential equation (6.17) in the direction v:

@

@tD�(p, t)v = Du(�(p, t), t)D�(p, t)v,

with D�(p, T ) = I. By an algebraic manipulation, this di↵erential equationcan be recast into the form

@

@t((D�(p, t)� I)v = Du(�(p, t), t)(D�(p, t)� I)v +Du(�(p, t), t)v.

At t = T , the time rate-of-change of the di↵erential of displacement at p inthe direction v is

@

@t(D�(p, t)v

��t=T

= Du(p, T )v.

The choice of T is arbitrary. Thus, the time rate-of-change of the linearstrain tensor is given by

"(p, t)(v, w) =1

2h(Du(p, t)T +Du(p, t))v, wi, (6.19)

where u is the fluid velocity. It has units of inverse time. Stress has unitsof force per area. Hooke’s law (stress is proportional to strain) suggests thedynamic stress must be some constant µ, which has units of mass per lengthper time, times " so that

� = �pI + 2µ".

The factor 2 is put in to remove the factor 1/2 in the Lagrange-Green tensor.Our construction of the stress tensor is almost correct! The only problem

is the interpretation of Hooke’s law. The strain is a matrix; simple propor-tionality is too strong. Hooke’s law is more general: it states that stress islinearly related to strain. More generally, there is some matrix valued lineartransformation of matrices K so that

� = �pI +K ".

Page 12: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

268 Fluids

There is no reason to take K = 2µI. On physical grounds, K applied to" is isotropic and symmetric. Isotropic means that K is invariant underrotations. Using these hypotheses, it is possible to prove that K has to be aspecial type of linear transformation. In fact, there are only two viscositiesµ and � and

� = �pI +K " = �pI + 2µ"+ � div uI.

For reasons that arise from thermodynamics, the viscosities are related by� = �2/3µ. The final form of the constitutive relation is the stress tensor

� = �pI +K " = �pI + 2µ"� 2

3µ div uI. (6.20)

Using �ij = 0 if i 6= j and �ij = 1 if i = j, the components of " are givenby

�ij = �p�ij + µ�@ui

@xj

+@uj

@xi

�� 2

3µ(r · u)�ij.

The function p is the pressure, µ is called the viscosity, and 1/2(@ui/@xj +@uj/@xi) is the (i, j)-component of the linear strain tensor. With this choiceof �, the boundary value problem

⇢(ut + (u ·r)u) =�rp+ µ�u+µ

3r(r · u) + ⇢b,

⇢t +r · (⇢u) = 0, (6.21)

u= 0 on @R,

consisting of Cauchy’s equation and the continuity equation—which are to-gether called the Navier-Stokes equations for a Newtonian fluid—is well-posed with the no-slip boundary condition. Here �u denotes the Laplacianapplied component-wise to the vector u; that is,

�u =

0

B@

@2u1

@x21+ @2u1

@x22+ @2u1

@x23

@2u2

@x21+ @2u2

@x22+ @2u2

@x23

@2u3

@x21+ @2u3

@x22+ @2u3

@x23

1

CA .

In some realistic situations (for example, the flow of water at constanttemperature) it is reasonable to make two further assumptions: the viscosityµ and density ⇢ are constant. Using the equation of continuity, the con-stant density assumption implies that the vector field u is divergence free.

Page 13: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 269

This is equivalent to the assumption that the fluid is incompressible (see Ap-pendix A.11). Under these assumptions, the Navier-Stokes equations (6.21)simplify to

⇢(ut + (u ·r)u) =�rp+ µ�u+ ⇢b,

r · u= 0, (6.22)

u= 0on @R.

Exercise 6.1. (a) Prove that if u is a su�ciently smooth vector field, then

r ·�u = �(r · u).

Conclude that if u divergence free, then r ·�u = 0. (b) Use part (a) to concludethat for incompressible flow with body force b satisfying r ·b = 0 (such a body forceis called conservative), the pressure is given by

�p = �⇢r · ((u ·r)u).

Exercise 6.2. Consider incompressible flow with no body force. Let � be a har-monic function (�� = 0) on three-dimensional space and f an arbitrary real func-tion of one real variable. (a) Show that the velocity field u(x, t) = f(t)r� solvesthe Navier-Stokes equations, ignoring the boundary condition. Find the exact ex-pression for the pressure. (b) Is it possible to find nontrivial solutions as in part(a) on a bounded domain such that the velocity field vanishes at the boundary?DIscuss.

Exercise 6.3. Show that it is not necessary to assume the density is constant(everywhere) to obtain equations (6.22) from equations (6.21). More precisely,prove that it su�ces to have D⇢/Dt = 0 (that is, the density is constant alongfluid particle paths).

Exercise 6.4. Define

�(x) =

8<

:

1, x 0;1� 3x2 + 2x3, 0 < x < 1;

0, x � 1.

(a) Show that � is continuously di↵erentiable and write out the formula for �0. (b)Recall the conservation of mass formula (the continuity equation) ⇢t+div(⇢u) = 0.Suppose that u is the constant vector field (2, 0, 0) and the density is distributed in

Page 14: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

270 Fluids

space so that the density is the same on each plane that is orthogonal to the x-axis.This means ⇢(x, y, z, t) does not depend on y or z; it can be viewed as a functionof x and t only. Using these simplifications, consider the initial value problem

⇢t + div(⇢u) = 0, ⇢(x, 0) = �(x).

Determine the value of ⇢(3, 5/4). Hint: Look for a traveling wave solution.

Exercise 6.5. The total kinetic energy of a fluid with density ⇢ and velocity uover the parcel A moving with the fluid via the flow � is

KE :=

Z

�(A,t)

1

2⇢(x, t)u(x, t) · u(x, t) dV.

Use the transport theorem and a computation in coordinates to show that

dKE

dt=

Z

�(A,t)⇢u · (ut + (ur)u) dV.

6.2 Scaling: The Reynolds Number and FroudeNumber

For simplicity, the Navier-Stokes equations (6.22) for incompressible flow isconsidered in this section. By introducing a characteristic length L and acharacteristic velocity V for the flow in question together with the corre-sponding natural time-scale ⌧ = L/V , the state variables are rendered di-mensionless via the assignment X = x/L, s = t/⌧ , U(X, s) = u(x, t)/V ,and P (X, s) = p(x, t)/(V 2⇢). Using the kinematic viscosity ⌫ := µ/⇢,the Reynolds (dimensionless) number Re := LV/⌫, and the Froude num-ber Fr := U/

pL|b|, which depends on the magnitude of the body force, the

system of equations for the velocity and pressure can be rescaled to the di-mensionless form of the Navier–Stokes equations for an incompressible fluid

Us + (U ·r)U =1

Re�U �rP +

1

Fr2b

|b| ,

r · U = 0,

U = 0 in @R⇤ (6.23)

where R⇤ is the image of the region R under the change of coordinates. Theexistence of this scaling is important: If two flows have the same Reynolds

Page 15: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 271

and Froude numbers, then the flows have the same dynamics. For example,flow around a scaled model of an airplane in a wind tunnel might be testedat the same Reynolds and Froude numbers expected for the airplane undercertain flight conditions. Perhaps the same Reynolds number can be obtainedfor testing by increasing the velocity in the wind tunnel to compensate forthe smaller length scale of the scaled model. In principle, the behavior of theflow around the scaled model is the same as for the full-sized aircraft.

6.3 The Zero Viscosity Limit

The scaled Euler’s equation in a region R⇤ is given by

Us + (U ·r)U =�rP +B,

r · U = 0,

U · ⌘ = 0 on @R⇤, (6.24)

where ⌘ is the outward unit normal vector field on @R⇤, can be viewed as anidealization of the scaled Navier–Stokes equations (6.23) for a fluid with zeroviscosity. The no-slip boundary condition for the Navier–Stokes equations isreplaced by the condition that there is no fluid passing through the boundary.

A naive expectation is that the limit of a family of solutions of the Navier–Stokes equations, as the Reynolds number increases without bound, is asolution of Euler’s equations. After all, the term �U/Re would seem toapproach zero as Re ! 1. Note, however, the possibility that the secondderivatives of the velocity field are unbounded in the limit and the di↵erentboundary conditions for the Navier-Stokes and Euler equations. For theseand other reasons, the limiting behavior of the Navier-Stokes equations forlarge values of the Reynolds number is not yet completely understood.

The necessity of di↵erent boundary conditions in passing from the Navier-Stokes to the Euler equations is the starting point for one of the most im-portant aspects of fluid dynamics, which was introduced by L. Prandtl in1904 (see [63, 66]), called boundary layer theory. The fundamental idea isthat for flows of interest in aerodynamics, for instance, the viscous e↵ectsare only important in a thin layer near the boundary where the no-slip con-dition must hold; away from the boundary, the flow is well-approximatedby Euler’s equations. Boundary layer theory will be discussed more fully inSection 6.12.3.

Page 16: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

272 Fluids

The possibility of large second derivatives of the velocity field is importantin another area of fluid dynamics: the study of turbulence, a subject that isbeyond the scope of this book.

6.4 The Low Reynolds’s Number Limit

In the zero viscosity limit, the pressure is scaled by P = p/(V 2⇢). Physically,the pressure is comparable to the momentum per unit volume. For flowswith low Reynolds’s numbers, the pressure is comparable to the forces dueto viscosity. For this reason, the correct scaling in this regime is X = x/L,s = t/⌧ , U(X, s) = u(x, t)/V , and P (X, s) = p(x, t)L/(µV ), which leads tothe equation

L

VUs + U ·rU = � 1

RerP +

1

Re�U.

After multiplying by the Reynolds number and passing to the limit as Re !0, we obtain the well-posed boundary value problem

rP =�U,

r · U = 0,

U = 0 on @R⇤. (6.25)

This system, usually called Stokes’s equations, is a useful approximation inmany di↵erent situations where the velocity is small, the viscosity is large,or the body size is small.

6.5 Flow in a Pipe

As an example of the solution of a fluid flow problem, let us consider perhapsthe most basic example of the subject: a special case of flow in a round pipe.1

Consider cylindrical coordinates r, ✓, and z where the z-axis is the axisof symmetry of a round pipe with radius a. More precisely, we consider thecoordinate transformation

x1

= r cos ✓, x2

= r sin ✓, x3

= z.

1The general nature of flow in a round pipe is not completely understood (see, forexample, B. Eckhardt (2008), Turbulence transition in pipe flow: some open questions.Nonlinearity. 21, T1–T11.)

Page 17: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 273

The basis vector fields for cylindrical coordinates are defined in terms of theusual basis of Euclidean space by

er := (cos ✓, sin ✓, 0), e✓ := (� sin ✓, cos ✓, 0), ez := (0, 0, 1).

We denote the coordinates of a vector field F on the cylinder {(r, ✓, z) : r a} with respect to the basis vector fields er, e✓, and ez by Fr, F✓, and Fz

respectively. The use of subscripts to denote coordinates in this section mustnot be confused with partial derivatives of F .

It is natural to expect that there are some flow regimes for which thevelocity field has its only nonzero component in the axial direction of thepipe; that is, the velocity field has the form

u(r, ✓, z, t) = (0, 0, uz(r, ✓, z, t)), (6.26)

where the components of this vector field are taken with respect to the basisvector fields er, e✓, ez and uz denotes the z component of the field (not thepartial derivative with respect to z).

We will express the Euler and the Navier–Stokes equations in cylindricalcoordinates. For a function f and a vector field F = Frer + F✓e✓ + Fzez onEuclidean space, the basic operators are given in cylindrical coordinates by

rf =@f

@rer +

1

r

@f

@✓e✓ +

@f

@zez,

r · F =1

r

@

@r(rFr) +

1

r

@F✓

@✓+@Fz

@z, (6.27)

�f =1

r

@

@r

�r@f

@r

�+

1

r2@2f

@✓2+@2f

@z2

(see Exercise 6.7). To obtain the incompressible Navier–Stokes equations incylindrical coordinates, consider the unknown velocity field u = urer+u✓e✓+uzez. Write this vector field in the usual Cartesian components by usingthe definitions of the direction fields given above, insert the result into theNavier–Stokes equations, and then compute the space derivatives using theoperators given in display (6.27). After multiplication of the vector consistingof the first two of the resulting component equations (that is, the equationsin the directions ex and ey) by the matrix

✓cos ✓ sin ✓� sin ✓ cos ✓

◆,

Page 18: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

274 Fluids

we obtain the equivalent system

@ur

@t+ (u ·r)ur �

1

ru2

✓ =1

Re

��ur �

1

r2(ur + 2

@u✓

@✓)�� @p

@r,

@u✓

@t+ (u ·r)u✓ +

1

ruru✓ =

1

Re

��u✓ �

1

r2(u✓ � 2

@ur

@✓)�� 1

r

@p

@✓,

@uz

@t+ (u ·r)uz =

1

Re�uz �

@p

@z,

r · u= 0, (6.28)

where the operators r and � are represented in cylindrical coordinates asin formulas (6.27).

The Euler equations in cylindrical coordinates for the fluid motion in thepipe are obtained from system (6.28) by deleting the terms that are dividedby the Reynolds number. If the velocity field u has the form given in equa-tion (6.26), then u automatically satisfies the appropriate boundary conditionfor the incompressible Euler equation; that is, the Neumann boundary con-dition @u

@n= 0, where n is a unit normal on the cylinder that models the wall

of the pipe. Thus, the Euler equations for the (scaled) velocity and pressurefields u and p reduce to the system

@p

@r= 0,

@p

@✓= 0,

@uz

@t+ uz

@uz

@z= �@p

@z,

@uz

@z= 0.

The first two equations imply that p is a function of z and t only; thesecond two equations imply that

@uz

@t= �@p

@z. (6.29)

After di↵erentiation of equation (6.29) with respect to z, it follows that@2p/@z2 = 0. Therefore, p = ↵+�z for some functions ↵ and � that dependonly on t. Using equation (6.29), uz = v(r) �

R t

0

�(s) ds for an arbitrarychoice of initial velocity v that may depend on r, but not ✓ or z. The generalsolution, for the class of velocities that have zero first and second componentsis

u(x, y, z, t) = v(r)�Z t

0

�(s) ds), p(x, y, z, t) = ↵(t) + �(t)z.

All such solutions, the initial velocity depends only on r. The no pen-etration boundary condition is always satisfied by the assumption that the

Page 19: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 275

first two components of u vanish. The no slip boundary condition can alsobe satisfied by taking v(a) = 0. Of course, so far our pipe has infinite length.

A physically more realistic requires some pressure di↵erential to push theflow. Suppose for example the pressure is constant with value p

0

at z = 0and p

1

at z = 1, and p0

> p1

. The model should then predict a nonzero flowvelocity with positive uz component. In this case, the pressure must be

p = p0

+ (p1

� p0

)z

and the third component of velocity is

uz = v(r) + (p0

� p1

)t.

The flow moves in the expected direction, but the velocity increases withoutbound as t grows to infinity. Thus, the model predicts an unrealistic flow foran arbitrary choice of initial velocity v.

Note that for arbitrary ↵, the pair of functions uz = 0 and p = ↵(t)is a solution of the Euler model with appropriate boundary conidtion. Atfirst sight it might seem that this simple example shows that solutions ofEuler’s equations with zero initial velocity and the no penetration boundarycondition does not have unique solutions. But, this is not the case: the nonuniqueness of pressure is allowed. The reason is simple: the gradient of thepressure appears in the equations of motion. Thus, the pressure is neverunique in a fluid flow problem; it is required to be unique up to the additionof a constant function of time.

There are several cases for Euler flow with the no penetration boundarycondition. For � = 0, the pressure does not depend on the position in thepipe and the fluid velocity field is constant with respect to z (along the flowdirection in the pipe). This is called plug flow. Because of its mathematicalsimplicity, plug flow is often used as a model. For example, plug flow is oftenused to model flow in tubular reactors studied in chemical engineering.

What about Navier–Stokes flow?By considering the same pipe, the same coordinate system, and the same

hypothesis about the direction of the velocity field, the Navier–Stokes equa-tions reduce to

@p

@r= 0,

@p

@✓= 0,

@uz

@t+ uz

@uz

@z=

1

Re�uz �

@p

@z,

@uz

@z= 0,

with the no-slip boundary condition at the wall of the pipe given by

uz(a, ✓, z, t) ⌘ 0.

Page 20: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

276 Fluids

This apparently simple system of fluid equations is di�cult to solve. But, wecan obtain a solution under two additional assumptions: The velocity fieldis in steady state and it is symmetric with respect to rotations about thecentral axis of the pipe. In other words, the partial derivatives of u withrespect to t and ✓ vanish. With these assumptions and taking into accountthat @uz/@z = 0, it su�ces to solve the single equation

1

Re

⇣1r

@

@r

�r@uz

@r

�⌘=@p

@z.

Because @p/@r = 0 and @p/@✓ = 0, @p/@z depends only on z and the left-hand side of the last equation depends only on r. Thus, the functions onboth sides of the equation must have the same constant value, say �.

It follows that p = ↵ + �z and

d(ru0z(r))

dr= (� Re)r

with the initial condition uz(a) = 0. The general solution of this ordinarydi↵erential equation has a free parameter because there is only one initialcondition given for a second-order di↵erential equation. The term with thefree parameter contains the factor ln r, which blows up as r approaches zero(that is, at the center of the pipe). With the free parameter set to zero, thesolution

uz(r) =1

4� Re (r2 � a2).

is continuous in the pipe and physically realistic. This steady state veloc-ity field, called Poiseuille flow, predicted from the Navier-Stokes model isparabolic with respect to the radial coordinate with the fastest flow at thecenter of the pipe and flow velocity zero at the pipe wall. Poiseuille flowis a close approximation to physical flow in a pipe for small Reynolds num-bers. As the Reynolds’s number is increased, a critical value is reached atwhich the radial symmetry hypothesis used to obtain the Poiseuille flow isviolated. Above this critical value, the steady state physical flow as measuredin experiments and the velocity field given by the Navier-Stokes model be-come more complex as the Reynolds’s number is increased. The flow regimechanges from laminar flow to turbulent flow. The causes of the onset of tur-bulent flow, the transition from laminar to turbulent flow, and the nature offully turbulent flow are not fully understood; they are important unsolvedproblems in physics and applied mathematics.

Page 21: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 277

Exercise 6.6. Consider Poiseuille flow in a section of length L of an infiniteround pipe with radius a. Suppose that the pressure is p in at the inlet of thesection and the flow speed at the center of the pipe is v in. Determine the pressureat the outlet. What happens in the limit as the Reynolds number grows withoutbound? Compare with the prediction of plug flow.

Exercise 6.7. (a) Write a detailed derivation of the gradient, divergence, andLaplacian in cylindrical coordinates (see display (6.27)). There are several methodsthat can be employed. One method begins by defining f c(r, ✓, z) = f(r cos ✓, r sin ✓, z)so that f(x, y, z) = f c(

px2 + y2, arctan(y/x), z) and continues by di↵erentiating

the latter equality with respect to x, y, and z. (b) Derive the expressions for thegradient, divergence, and Laplacian in spherical coordinates.

Exercise 6.8. [Hagen-Poiseuille law] Return to unscaled variables and reconsidersteady-state Poiseuille flow in a round pipe of radius a.. WIth the z coordinate inthe direction of the pipe axis, the equations of motion are

px = 0, py = 0, pz = µ(vxx + vyy), vz = 0

with the no slip boundary condition at the pipe wall. (a) Write out the full phys-ically realistic general solution for the pressure and and velocity. (b) Suppose thepressure at z = 0 is p

0

and the pressure at some z = ` > 0 is p1

with p0

> p1

. Thepressure drop on this section of pipe is p

0

� p1

. Determine the relation betweenthe pressure drop and the volumetric flow rate in the pipe (volume per time offluid passing a given cross section of the pipe). This relation is called the Hagen-Poiseuille law. (c) Determine the average velocity through a cross section.

Exercise 6.9. Determine the profile for Stokes flow in a round pipe of radius aunder the assumption that the flow is symmetrical with respect to rotations aroundthe central axis of the pipe.

Exercise 6.10. Consider flow over a flat infinite surface that is tilted ✓ radianswith respect to the horizontal. Ignore the banks of the river, assume that the flowvelocity is all in the downstream direction, and the body force driving the flowis gravity. (a) Show that the velocity at z units above the plate (river bottom)is u = 1

2µ⇢gz(2h � z) sin ✓. (b) The Mississippi River drops 2.5 inches per mile.Suppose its average depth is 30 feet. Compute the surface velocity of the river?Does the velocity computed using these assumptions agree with observation? Whichif any of the assumptions is not realistic? Discuss.

Page 22: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

278 Fluids

6.6 Bernoulli’s Form of Euler’s Equations

6.6.1 Isentropic Flow

In this section we will assume that our fluid is incompressible. In particular,we will assume the density is constant. More generally, the same results arevalid for isentropic flow; that is, there is some function q called the enthalpy,such that

grad q =1

⇢grad p.

In case the density is constant, q := p/⇢.Using the vector identity

1

2grad(u · u) = u⇥ curlu+ (u ·r)u,

we may rewrite Euler’s equation of motion (6.13) in the form

ut � u⇥ curlu = b+ grad(�1

2(u · u)� p

⇢).

The definition

↵ := �1

2|u|2 � p

⇢,

is used to obtain Bernoulli’s form of Euler’s equations

ut = u⇥ curlu+ grad↵ + b,

div u= 0,

u · ⌘ = 0 in @R. (6.30)

6.6.2 Potential Flow

An important specialization of Bernoulli’s form of the incompressible Euler’sequations, called potential flow, is the case where the velocity field u is thegradient of a potential � so that u = grad� and the body force is the gradientof a potential B so that b = gradB. By substitution into system (6.30), theidentity curl(gradu) = 0, and some rearrangement, the equations of motioncan be recast into the form

grad(@�

@t+

1

2| grad�|2 + p

⇢� B) = 0, �� = 0. (6.31)

Page 23: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 279

As a result, we see immediately that there is a number C, constant withrespect to the space variable, such that

@�

@t+

1

2| grad�|2 + p

⇢� B = C. (6.32)

In particular, if u is a steady state velocity field, then

p

⇢+

1

2|u|2 � B = C. (6.33)

This is Bernoulli’s law. In case the body force is gravity, Bernoulli’s lawstates that (at a height h above a reference surface)

p+1

2⇢|u|2 + g⇢h = C (6.34)

An important consequence of Bernoulli’s law is Bernoulli’s principle: Atconstant height, if the velocity of an incompressible fluid increases, thenthe pressure decreases. For example, flow over an airplane wing (which ismore curved on top) produces lift because the velocity of air relative to thewing is greater for flow over the top of the wing compared to flow over itsbottom. Exactly why the flow is faster over the top of an airplane wing canbe explained, but not simply (see, for example, [43]).

Exercise 6.11. Consider an open tank containing water with a round hole inthe bottom of the tank. Show that the velocity of the fluid in the drain is (approxi-mately) proportional to the square root of the depth of the fluid in the tank. Moreprecisely, this velocity is approximately

p2gh, where h is the depth of the water in

the tank and g is the acceleration due to gravity. What assumptions are requiredto make the velocity exactly

p2gh.

Potential Flow in Two Dimensions

In view of the second equation of system (6.31), the potential � is a harmonicfunction. Therefore, by considering a hypothetical flow on a two-dimensionalplane with Cartesian coordinates (x, y) and velocity field u = (x, y), thepotential � is locally the real part of a holomorphic function, say f = �+ i .Moreover, the pair �, satisfies the Cauchy–Riemann equations

@�

@x=@

@y,

@�

@y= �@

@x.

Page 24: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

280 Fluids

Thus, the assumption that u = grad� implies the fluid motions are solutionsof an ordinary di↵erential equation that can be viewed in two di↵erent ways:as the gradient system

x =@�

@x, y =

@�

@y; (6.35)

or the Hamiltonian system

x =@

@y, y = �@

@x. (6.36)

The function , a Hamiltonian function for system (6.36), is called thestream function. The orbits of system (6.36), called stream lines, all lie onlevel sets of . Because the stream lines are also orbits of the gradient sys-tem (6.35), there are no periodic fluid motions in a region where the function� is defined.

It should be clear that function theory can be used to study planar po-tential flow. For example, if is a harmonic function defined in a simplyconnected region of the complex plane such that the boundary of the regionis a level set of , then is the imaginary part of a holomorphic functiondefined in the region, and therefore is the stream function of a steady stateflow. This fact can be used to find steady state solutions of Euler’s equations.

As an example, let us consider plug flow in a round pipe with radius aand notice that every planar slice containing the axis of the pipe is invariantunder the flow. Thus, it seems reasonable to consider the two-dimensionalflow on the strip

S := {(x, y) : 0 < y < 2a},

viewed as such a slice, where we have taken x as the axial direction andthe center pipe as the line with equation y = a. The plug flow solution ofEuler’s equations in S is given by the velocity field u = (c, 0) and pressurep = p

0

, where c and p0

are constants. It is a potential flow, with potential�(x, y) = cx, stream function (x, y) = cy, and complex potential f(x, y) =cz = cx+ icy.

Suppose that Q is an invertible holomorphic function defined on S andthat R is the image of S under Q, then w 7! f(Q�1(w)) for w 2 R is aholomorphic function on R with real part w 7! �(Q�1(w)) and imaginarypart w 7! (Q�1(w). Thus, the function given by w 7! (Q�1(w)) is astream function for a steady state potential flow in R with potential w 7!

Page 25: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 281

(Q�1(w). In particular, stream lines of map to stream lines of w 7! (Q�1(w)). And, this flow is a solution of Euler’s equation in the domain R.

For example, let us note that w := Q(z) =pz has a holomorphic branch

defined on the strip S such that this holomorphic function maps S into theregion R in the first quadrant of the complex plane bounded above by thehyperbola {(�, ⌧) : �⌧ = a}. In fact, Q�1(w) = w2 so that

x = �2 � ⌧ 2, y = 2�⌧.

The velocity field of the new “flow at a corner” is

(2c�,�2c⌧).

The corresponding pressure is found from Bernoulli’s equation (6.33). Infact, there is a constant p

1

such that

p = p1

� 2c2(�2 + ⌧ 2). (6.37)

The stream lines for the flow at a corner are all hyperbolas.The flow near a wall is essentially plug flow. In fact, if we consider, for

example, the velocity field on a vertical line orthogonal to the �-axis, say theline with equation � = �

0

, then the velocity field near the wall, where ⌧ ⇡ 0,is closely approximated by the constant vector field (2c�

0

, 0). In other words,the velocity profile is nearly linear.

Exercise 6.12. Consider the plug flow vector field u = (c, 0) defined in a hori-zontal strip in the upper half plane of width 2a. Find the push forward of u intothe first quadrant with respect to the map Q(z) =

pz with inverse Q�1(w) = w2.

Is this vector field a steady state solution of Euler’s equations at the corner for anincompressible fluid? Explain.

6.7 Equations of Motion in Moving Coordi-nate Systems

Moving coordinate systems are discussed in this section in the context ofgeneral mechanical systems. The theory is applied to the equations of fluiddynamics and particle mechanics.

Page 26: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

282 Fluids

6.7.1 Moving Coordinate Systems

An inertial coordinate system is a coordinate system in which Newton’s lawsof motion are valid; in particular, a free particle moves along a (straight)line.

Consider a rectangular inertial coordinate system, with coordinates (⇠, ⌘, ⇣),and the motion of a particle whose position vector in this coordinate systemis R. Suppose there is a second (moving) rectangular coordinate system withrectangular coordinates (x, y, z) and the same origin as the inertial coordi-nates. Let Q denote the position of the particle with respect to the movingcoordinates. The symbols Q and R refer to the same point in space; but,Q 6= R when Q is given as the triple of numbers that specify its coordinatesin the moving coordinate system and R is given as the triple of numbersthat specify its coordinates in the inertial coordinate system. This can be aconfusing distinction. Suppose that

Q =

0

@abc

1

A

with respect to the moving coordinates. This means

Q = aex + bey + cez,

where ex, ey, and ez are the unit basis vectors (with respect to the usualinner product in the inertial coordinate frame) in the positive directions onthe coordinate axes of the rectangular moving coordinate system. Thesebasis vectors have coordinate representations in the inertial coordinates:

ex =

0

@a11

a21

a31

1

A , ey =

0

@a12

a22

a32

1

A , ex =

0

@a13

a23

a33

1

A .

Thus, the coordinate representation of Q in the inertial coordinates is

a

0

@a11

a21

a31

1

A+ b

0

@a12

a22

a32

1

A+ c

0

@a13

a23

a33

1

A .

Or, equivalently, 0

@↵��

1

A =

0

@a11

a12

a13

a21

a22

a23

a31

a32

a33

1

A

0

@abc

1

A

Page 27: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 283

are the inertial coordinates of the point R. More compactly, we will write

R = AQ,

where A is the matrix with components aij. In the general case, where thetwo coordinate systems might not share a common origin, the origin of themoving frame has a position vector � in the inertial frame. Also there is anotion of parallel transport in Euclidean space. The inertial frame may bemoved parallel to itself until its origin coincides with the origin of the movingframe. As before, there is a rotation matrix A that changes the coordinatesof the particle in the moving frame to the coordinates of this point in thetransported inertial frame. Thus, the coordinates with respect to the originalinertial frame are obtained by adding �; that is,

R = � + AQ. (6.38)

The matrix A is an orthogonal transformation; that is, A�1 = AT withrespect to the usual inner product, which will be denoted by angled bracketsh i (see Exercise 6.13). Here, A gives the inertial coordinates of the positionvector Q and its inverse gives the moving coordinates of a point expressed inthe coordinates of the inertial frame. Because A is orthogonal,

AAT = ATA = I.

By Newton’s second law, the equation of motion of the particle withposition R moving in the inertial frame is

mR = F, (6.39)

where m is the mass of the particle and F is the (vector) sum of the forcesacting on the particle.

Our goal is to express the equation of motion for our particle in the movingcoordinate system. A convenient method is to first describe the motion usingthe position, velocity, and acceleration of the origin of the moving coordinatesystem in inertial coordinates and the inertial representations r := AQ ofthe particle’s position Q, v := AQ of its velocity Q, and a := AQ of itsacceleration. In view of equation (6.38), the velocity of the particle can beexpressed in the form

R= � + AQ+ AQ,

= � + AAT (AQ) + v

= � + AAT r + v. (6.40)

Page 28: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

284 Fluids

The transformation ⌦ := AAT has a special property: it is skew-symmetric;that is, ⌦T = �⌦. This fact follows simply by di↵erentiation of the identityAAT = I with respect to the temporal parameter. It follows that

AAT + AAT = 0; (6.41)

therefore,⌦T = (AAT )T = AAT = �AAT = �⌦.

By using the definition of skew-symmetry, the matrix representation of ⌦must have the form 0

@0 �!

3

!2

!3

0 �!1

�!2

!1

0

1

A . (6.42)

The action of ⌦ on a vector W (that is, ⌦W ) can also be expressed usingthe vector cross product (with respect to the usual orientation of space) by

⌦W = ! ⇥W, (6.43)

where ! is the vector with components (!1

,!2

,!3

). Thus, the velocity maybe expressed in either of the forms

R= � + ⌦r + v,

R= � + ! ⇥ r + v. (6.44)

Using the first equation in display (6.44) for the velocity of our particle,we find its acceleration to be

R= � + v + ⌦r + ⌦r. (6.45)

Since, by definition, r = AQ and v = AQ, we have that

r = AAT (AQ) + AQ

= ⌦r + v, (6.46)

and

v = AAT (AQ) + AQ

= ⌦v + a. (6.47)

Page 29: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 285

These formulas for r and v are substituted into equation (6.45) to obtain theacceleration in the form

R= � + a+ 2⌦v + ⌦r + ⌦2r; (6.48)

or, equivalently,

R= � + a+ 2! ⇥ v + ! ⇥ r + ! ⇥ (! ⇥ r). (6.49)

The equation of motion (6.39) may be recast in the forms

ma = �m(� + 2⌦v + ⌦r + ⌦2r) + F (6.50)

orma = �m(� + 2! ⇥ v + ! ⇥ r + ! ⇥ (! ⇥ r)) + F, (6.51)

where the vectors a, v, and r give the acceleration, velocity and positionof the particle, moving under the influence of the force F , with respect toan observer in the moving coordinate system. Here, the acceleration andthe state vectors r and v are expressed in the inertial coordinates. Theircoordinates in the moving frame are simply obtained from multiplication onthe left by the orthogonal matrix AT . For example, using equation (6.50),the equation of motion in the inertial coordinates is

mAQ = �m(� + 2⌦AQ+ ⌦AQ+ ⌦2AQ) + F ; (6.52)

and, in the moving coordinates it is

mQ = �m(AT � + 2AT⌦AQ+ AT ⌦AQ+ AT⌦2AQ) + ATF. (6.53)

The last equation can also be written in the form

mQ = �m(AT � + 2�Q+ �Q+ �2Q) + ATF, (6.54)

where � := AT⌦A; that is, � and ⌦ are names for the same linear transfor-mation expressed in di↵erent bases.

Since the equation of motion (6.50) (or (6.54)) is valid for the non-inertialobserver, this observer treats the terms on the right-hand side of the equationof motion as additional forces. The term � is the acceleration of the originof the non-inertial frame, ⌦r = ! ⇥ r is the acceleration due to the rotationof the moving frame, 2m⌦v = 2m! ⇥ v is the Coriolis force, and m⌦2r =m!⇥ (!⇥ r) is the centrifugal force. Of course, from Newtonian mechanics,F is the only force acting on the particle. The remaining fictitious forces areartifacts of reference to a non-inertial coordinate system.

Page 30: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

286 Fluids

Exercise 6.13. Prove that the coordinate transformation in display (6.38) isorthogonal.

6.7.2 Rotation

To consider pure rotation, imagine a rotating disk whose motion is measuredwith respect to a rectangular rotating coordinate system (with coordinates(Q

1

, Q2

, Q3

)), whose third coordinate-axis coincides with the axis of rotationof the disk, whose origin is at the intersection of the disk with this axis,and whose first two coordinates are fixed in the disk. In addition, supposewithout loss of generality, that an inertial frame with coordinates (⇠, ⌘, ⇣)has the same origin. Of course, we could choose the inertial coordinates sothat the ⇣-axis corresponds with the axis of rotation; but, we will not makethis assumption to illustrate an important feature of three-dimensional space:Each rotation is determined by a vector (specifying the axis of rotation) andthe rotation angle about this direction. Thus, under our assumption thatthere is a fixed axis of rotation, there is an orthogonal transformation B,which does not depend on time, taking the unit vector in the direction ofthe ⇣-axis of the inertial frame to one of the two unit vectors in the directionof the axis of rotation. The matrix A, which determines the coordinatetransformation R = AQ from the moving frame to the inertial frame, takesthe form BTRB, where

R :=

0

@cos ✓ � sin ✓ 0sin ✓ cos ✓ 00 0 1

1

A (6.55)

is the rotation matrix (in inertial coordinates) for the motion about the Q3

-axis in the moving frame.

By a direct computation, it is easy to check that

⌦ = AAT =

0

@0�✓ 0✓ 0 00 0 0

1

A , (6.56)

which is exactly the angular velocity in matrix form; that is, the action ofthis matrix on a vector W is

� ⇥W,

where � is the vector with components (0, 0, ✓).

Page 31: Chapter 6 Conservation of Momentum: Fluids and Elastic Solids

Fluids 287

For rotations in three-dimensional space, we have the useful identity

AAT = AT A. (6.57)

Thus, ⌦ = AT A.

6.8 Fluid Motion in Rotating Coordinates

The equation of motion (6.53) of a particle in a moving coordinate system isvalid for a particle of fluid. We will obtain the transformation of the Eulerianequations of motion for a fluid to a moving coordinate system.

Let us consider the equations of motion (conservation of momentum andmass) for a fluid in a gravitational field

⇢Du

Dt=�rp+ µ�u+ ⇢g

D⇢

Dt=�⇢r · u, (6.58)

where it is important to note that the derivatives on the left-hand sides arethe material derivatives of the velocity and the density respectively; that is,

Du

Dt=@u

@t+ u ·ru,

D⇢

Dt=@⇢

@t+ u ·r⇢.

Also, as a remark, note that the equations of motion do not form a closed sys-tem; there are three state variables u, p, and ⇢ and only two equations. Moreprecisely, there are five state variables (if we include as separate variables thethree components of u) and four equations (if we view the conservation ofmomentum as three scalar equations). To close the system we need one addi-tional scalar equation, for example, an equation of state that relates densityand pressure or the statement that the density is constant.

For simplicity, let us consider a moving coordinate system whose originis fixed at the origin of the inertial coordinates and the transformation fromthe moving coordinates to the inertial coordinates given by

R = AQ.