24.02.2015 Views

Concise Fluid Mechanics - WVU Mechanical and Aerospace ...

Concise Fluid Mechanics - WVU Mechanical and Aerospace ...

Concise Fluid Mechanics - WVU Mechanical and Aerospace ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Concise</strong> <strong>Fluid</strong> <strong>Mechanics</strong><br />

A.V.Smirnov<br />

c Draft date September 12, 2004


Contents<br />

Contents<br />

Preface<br />

Nomenclature<br />

i<br />

v<br />

vii<br />

1 Properties <strong>and</strong> Variables 1<br />

1.1 Kinematic variables . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.1.1 Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.1.2 Substantial derivative . . . . . . . . . . . . . . . . . . . . . 3<br />

1.1.3 Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.1.4 Strain rate <strong>and</strong> vorticity . . . . . . . . . . . . . . . . . . . . 4<br />

1.2 Thermodynamic Variables . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.2.1 Equations of state . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.2.2 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.3 <strong>Fluid</strong> Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

1.3.1 Thermodynamic properties . . . . . . . . . . . . . . . . . . 8<br />

1.3.2 Transport Properties . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.3.3 Other properties . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

1.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

2 Fundamental Laws 15<br />

2.1 Conservation of Mass . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

i


ii<br />

CONTENTS<br />

2.1.1 General formulation . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.1.2 Constant density flow . . . . . . . . . . . . . . . . . . . . . 16<br />

2.1.3 Stream function . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2 Conservation of Momentum . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.2.1 General formulation . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.2.2 Constant density flow . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.3 Vorticity formulation . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.4 Potential flow . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.2.5 2D limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.2.6 Viscous limit . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

2.2.7 Inviscid limit . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.2.8 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . 29<br />

2.3 Pressure Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

2.3.1 General formulation . . . . . . . . . . . . . . . . . . . . . . 32<br />

2.3.2 Constant density flow . . . . . . . . . . . . . . . . . . . . . 32<br />

2.3.3 Viscous limit . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

2.3.4 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . 33<br />

2.4 Energy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

2.4.1 General formulation . . . . . . . . . . . . . . . . . . . . . . 34<br />

2.4.2 Constant density flow . . . . . . . . . . . . . . . . . . . . . 37<br />

2.4.3 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . 39<br />

2.5 Curvilinear Coordinates . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

2.5.1 Invariant forms . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

2.5.2 Non-inertial coordinate systems . . . . . . . . . . . . . . . 40<br />

2.6 The Law of Similarity . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

2.6.1 PI-Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

2.6.2 Non-dimensional formulations . . . . . . . . . . . . . . . . . 46<br />

2.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3 Laminar flows 53


CONTENTS<br />

iii<br />

3.1 Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

3.2 Confined flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.2.1 Flow between parallel plates . . . . . . . . . . . . . . . . . 54<br />

3.2.2 Axially moving concentric cylinders . . . . . . . . . . . . . . 56<br />

3.2.3 Rotating concentric cylinders . . . . . . . . . . . . . . . . . 57<br />

3.2.4 Poiseuille flow through ducts . . . . . . . . . . . . . . . . . 59<br />

3.2.5 Combined Couette-Poiseuille flows . . . . . . . . . . . . . 63<br />

3.2.6 Non-circular ducts . . . . . . . . . . . . . . . . . . . . . . . 64<br />

3.3 Unsteady flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

3.3.1 <strong>Fluid</strong> oscillation above infinite plate . . . . . . . . . . . . . . 66<br />

3.3.2 Unsteady flow between infinite plates . . . . . . . . . . . . 68<br />

3.4 Creeping flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

3.4.1 Stokes flow around a sphere . . . . . . . . . . . . . . . . . 72<br />

3.4.2 2D Creeping flows . . . . . . . . . . . . . . . . . . . . . . . 76<br />

3.4.3 Lubrication theory . . . . . . . . . . . . . . . . . . . . . . . 76<br />

3.5 Boundary layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

3.5.1 Flat plate integral analysis . . . . . . . . . . . . . . . . . . 81<br />

3.5.2 Laminar boundary layer equations . . . . . . . . . . . . . . 85<br />

3.5.3 Blasius solution . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

3.5.4 Reynolds analogy . . . . . . . . . . . . . . . . . . . . . . . 92<br />

3.5.5 Free shear flows . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

3.6 Integral methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

3.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

4 Turbulent flows 105<br />

4.1 Transition to turbulence . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

4.2 Turbulence Modeling . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

4.2.1 LES models . . . . . . . . . . . . . . . . . . . . . . . . . . . 106<br />

4.2.2 RANS models . . . . . . . . . . . . . . . . . . . . . . . . . 107


iv<br />

CONTENTS<br />

Bibliography 113<br />

A Introduction to Tensor Calculus 115<br />

A.1 Coordinates <strong>and</strong> Tensors . . . . . . . . . . . . . . . . . . . . . . . 116<br />

A.2 Cartesian Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

A.2.1 Tensor Notation . . . . . . . . . . . . . . . . . . . . . . . . 120<br />

A.2.2 Tensor Derivatives . . . . . . . . . . . . . . . . . . . . . . . 126<br />

A.3 Curvilinear coordinates . . . . . . . . . . . . . . . . . . . . . . . . 128<br />

A.3.1 Tensor invariance . . . . . . . . . . . . . . . . . . . . . . . 128<br />

A.3.2 Covariant differentiation . . . . . . . . . . . . . . . . . . . . 132<br />

A.3.3 Orthogonal coordinates . . . . . . . . . . . . . . . . . . . . 134<br />

A.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140<br />

B Curvilinear coordinate systems 143<br />

C Solutions to problems 147<br />

D Midterm Exam Topics: Laminar Flow Solutions 193<br />

E Final Exam Topics 197<br />

E.1 Fundamental Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . 197<br />

E.2 Analytical Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . 198<br />

E.3 Boundary Layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200<br />

E.4 Turbulence Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . 202<br />

Index 203


Preface<br />

The idea behind this book is to provide a formal but concise introduction to theoretical<br />

fluid mechanics.<br />

The book covers the traditional topics of fluid mechanics, such as the fundamental<br />

equations of motion, compressible <strong>and</strong> incompressible forms, <strong>and</strong> invariant<br />

formulations, special analytical solutions, laminar flows, elements of boundary<br />

layer theory, main aspects of turbulence modeling <strong>and</strong> numerical methods. The<br />

emphasis is on viscous flow phenomena.<br />

The author tried to put more emphasis on mathematical rigor rather than<br />

on lengthy narrative. Tensor notation is used extensively throughout the book.<br />

However, the knowledge of tensor calculus is not required of the reader, since<br />

enough introductory material is provided in the appendix.<br />

v


!<br />

"<br />

<br />

)<br />

<br />

¦<br />

'<br />

is defined as ¤ , or ¡ is equivalent to ¤<br />

¦<br />

<br />

¦<br />

<br />

<br />

Nomenclature<br />

. Note: ¡§¦¨¤©¦¡¤<br />

partial derivative over time: <br />

¡£¢¥¤ ¡<br />

¡§¦¨¤©¦<br />

¦ ¡©¦¨¤§¦ ¢<br />

¡ <br />

¡ ¦<br />

partial derivative over<br />

control volume<br />

time<br />

:<br />

<br />

¡ ¢¡ <br />

¦/ .<br />

021¨3<br />

6 ¦7<br />

"54<br />

-th component of a coordinate ( # =0,1,2), or <br />

fluid velocity:<br />

strain tensor<br />

any variable of coordinates <strong>and</strong> time<br />

stress tensor<br />

viscosity<br />

8<br />

¢ 8(:


¦<br />

¦ "<br />

'<br />

Chapter 1<br />

Properties <strong>and</strong> Variables<br />

Consider =?>@ a dimensional space of real ACBEDGF numbers representing physical<br />

space of = (= dimension =2,3) <strong>and</strong> time. The state of the fluid will be represented<br />

, @JIKIL= ), continuous with their<br />

#H<br />

by real functions of coordinates <strong>and</strong> ( time<br />

derivatives up to the second order. This is an Eulerian description in which both<br />

<strong>and</strong> represent a set of independent variables. In an alternative Lagrangian<br />

description the fluid is specified by a set of moving particles. These fluid particles<br />

form a continuum <strong>and</strong> their coordinates are themselves functions of time. Consequently,<br />

time becomes the only independent variable in this case. Our objective<br />

"<br />

will be to formulate the laws of fluid motion in Eulerian, fixed-space coordinates.<br />

The set of independent variables can be extended beyond space coordinates<br />

<strong>and</strong> time by introducing properties of the fluid. These properties describe<br />

different physical processes <strong>and</strong> are classified accordingly as thermodynamic<br />

properties, transport properties, etc.<br />

Dependent variables are functions of the independent variables implicitly<br />

expressed in a physical law. Dependent variables can also be classified according<br />

to the physical process they describe, <strong>and</strong> we shall consider only two types:<br />

kinematic variables <strong>and</strong> thermodynamic variables.<br />

1.1 Kinematic variables<br />

1.1.1 Velocity<br />

Let’s define a fluid particle as an infinitely small element of the fluid.<br />

1


(1.1) )<br />

¦<br />

'<br />

¦<br />

R<br />

"<br />

<br />

<br />

F<br />

'<br />

<br />

B<br />

)<br />

R<br />

"<br />

R<br />

Y<br />

R<br />

¢<br />

¦<br />

"<br />

'<br />

)<br />

¦<br />

'<br />

'<br />

2 CHAPTER 1. PROPERTIES AND VARIABLES<br />

Definition 1.1.1 <strong>Fluid</strong> velocity<br />

<strong>Fluid</strong> velocity, )<br />

¦ 1M3<br />

, 3 ¢N% "54<br />

<br />

at a given point in space <strong>and</strong> time, 1¨3 '<br />

IOIL<br />

is equal to the velocity of a fluid particle<br />

-<br />

:<br />

QP<br />

¦<br />

"54<br />

1¨3<br />

"54<br />

¦ 1M3<br />

"54 ¢SR QP<br />

The definition above can be inverted to define a particle trajectory.<br />

Definition 1.1.2 Particle trajectory<br />

For a given vector field of fluid ) velocities,<br />

a solution to the following problem:<br />

¦ 1¨3<br />

"54<br />

, particle P<br />

¦ 1 "54<br />

trajectory, , is<br />

R QP<br />

¦ 1¨3<br />

"54<br />

¦ 1 "54<br />

P '<br />

QP<br />

¦ 1UT 4 <br />

WV<br />

¦<br />

From these definitions it follows that the vector of velocity is tangential to<br />

particle trajectory at each point in the fluid. Particle trajectory is also called a<br />

streamline.<br />

In what follows we shall drop the super-index X , <strong>and</strong> also adopt the following<br />

dot-notation for particle velocity:<br />

(1.2)<br />

¦¢<br />

meaning that a time derivative of fluid particle position is taken along the particle’s<br />

trajectory at a space point . This notation should not create a confusion since<br />

space <br />

coordinates do not depend on time, while the fluid-particle coordinates<br />

do. Hence when using time derivatives of coordinates, as in (1.2), we will mean<br />

fluid particle coordinates.<br />

We shall use the same dot notation for partial time derivatives of other fluid<br />

variables at a fixed point in space. For example, for any variable, 021¨3 '<br />

"54<br />

(1.3)<br />

0 ¢SY <br />

021¨3<br />

"54<br />

"


¦<br />

¦<br />

¦ "<br />

'<br />

[<br />

¦ "<br />

'<br />

V<br />

<br />

"<br />

<br />

[<br />

[<br />

¦<br />

<br />

0<br />

"<br />

<br />

¦<br />

'<br />

<br />

¦<br />

'<br />

<br />

0<br />

¦<br />

'<br />

)<br />

¦<br />

[<br />

¦ "<br />

'<br />

"<br />

[<br />

R<br />

0<br />

"<br />

><br />

<br />

><br />

><br />

¦<br />

"<br />

Y<br />

¦<br />

0 <br />

<br />

¦<br />

'<br />

<br />

<br />

¦<br />

'<br />

¦<br />

'<br />

¦<br />

'<br />

¦ "<br />

'<br />

<br />

)<br />

¦<br />

'<br />

¦<br />

[<br />

¦<br />

¦<br />

"<br />

><br />

><br />

Y<br />

T<br />

<br />

Y<br />

¦<br />

'<br />

¦<br />

'<br />

[<br />

"<br />

"<br />

1.1. KINEMATIC VARIABLES 3<br />

1.1.2 Substantial derivative<br />

Definition 1.1.3 <strong>Fluid</strong> element<br />

By fluid element we shall underst<strong>and</strong> a finite volume of a fluid, which is small<br />

enough that the velocity of all it’s points can be approximated by the velocity of a<br />

single fluid particle inside the element.<br />

Partial derivatives (1.3) describe changes in a variable at a fixed space point<br />

attributed to the explicit time dependence of the variable. There are also changes<br />

brought about solely by the motion of the fluid, i.e. due to the fact that different<br />

fluid elements cross the given space point, changing the fluid variable at that<br />

point. To account for all the changes we introduce a substantial derivative. It is<br />

equal to the rate of change of fluid variable inside a fluid element moving with the<br />

velocity of the fluid.<br />

Consider a change of a fluid variable <br />

fluid element moved to a nearby position 1 0Z1 <br />

"54<br />

inside a fluid element as the<br />

:<br />

"54<br />

>\[]<br />

>\[<br />

(1.4)<br />

021<br />

"54<br />

021<br />

"54<br />

021<br />

"54 0Z1<br />

"54<br />

>\[]<br />

>\[<br />

¦ []<br />

" [<br />

Y <br />

where we used a Taylor expansion up to the first order. We can rewrite it in a<br />

more compact tensor notation (Sec.A):<br />

(1.5)<br />

021 <br />

<br />

021<br />

"54 021<br />

"54<br />

0 ¦ 1<br />

"54<br />

"54<br />

[^ ¦_<br />

[]<br />

[<br />

Considering that the displacement follows the fluid element, it should be<br />

a product of velocity <strong>and</strong> time: . Then we can rewrite the expression<br />

above in terms of variable change, , as follows:<br />

>^[^<br />

>\[<br />

[^<br />

0 ¢ 021<br />

"54a`<br />

0Z1<br />

"54<br />

(1.6)<br />

>\[]<br />

>\[<br />

021<br />

"54<br />

0 ¦ 1<br />

"54<br />

And dividing both sides by [<br />

, <strong>and</strong> taking the limit of [<br />

we have:<br />

"cb<br />

(1.7)<br />

dfefg Oi h<br />

¢ R<br />

¦ 0 ¦<br />

>^)


. ¦/k @<br />

l<br />

(1.9)<br />

(1.10) n<br />

)<br />

1<br />

)<br />

R<br />

¦<br />

"<br />

<br />

¦<br />

)<br />

1<br />

)<br />

1<br />

)<br />

><br />

1<br />

)<br />

<br />

)<br />

)<br />

4 CHAPTER 1. PROPERTIES AND VARIABLES<br />

this expression represents a substantial derivative of a fluid variable, which describes<br />

the change of that variable in the coordinate system moving with the fluid<br />

, in (1.7) is also called a convective derivative.<br />

element. The last term 1 )<br />

¦ 0 ¦ 4<br />

1.1.3 Acceleration<br />

Applying (1.7) to velocity itself, we have the relation for flow acceleration:<br />

(1.8)<br />

R )<br />

¦f<br />

j<br />

>^)Wj)<br />

where we used the dummy index rule (A.2.16).<br />

1.1.4 Strain rate <strong>and</strong> vorticity<br />

We can formally represent a velocity ) derivative<br />

an asymmetric parts:<br />

¦f<br />

as a sum of a symmetric <strong>and</strong><br />

@<br />

l<br />

¦f `<br />

m ¦ 4<br />

¦fk<br />

@<br />

l<br />

¦f<br />

m ¦ 4<br />

>^)<br />

¦7<br />

where these parts becomes the newly introduced strain rate ( . ¦/<br />

¦/<br />

(n<br />

) tensors.<br />

) <strong>and</strong> vorticity<br />

>on<br />

. ¦/<br />

Definition 1.1.4 Strain rate tensor<br />

The strain rate tensor, . ¦/<br />

, is defined as:<br />

m ¦ 4<br />

¦f<br />

Definition 1.1.5 Vorticity tensor<br />

The vorticity tensor, n<br />

¦/<br />

is defined as<br />

>^)<br />

¦7k<br />

¦f `<br />

m ¦ 4<br />

@<br />

l


(1.11) n<br />

(1.12) n<br />

(1.13) X<br />

¦<br />

n<br />

n<br />

F<br />

<br />

<br />

1<br />

1<br />

)<br />

1<br />

)<br />

<br />

F<br />

j<br />

rq<br />

F<br />

X<br />

1<br />

)<br />

`<br />

`<br />

1 ; '5s<br />

)<br />

)<br />

<br />

j<br />

`<br />

F<br />

4<br />

F<br />

4<br />

1.2. THERMODYNAMIC VARIABLES 5<br />

Definition 1.1.6 Vorticity vector<br />

The vorticity vector, n<br />

, is defined as<br />

¦Gp5¦/<br />

j n<br />

From this definition, <strong>and</strong> the definition of vorticity tensor (1.10), we have:<br />

¦$<br />

p5¦/<br />

4<br />

@<br />

l<br />

m<br />

j<br />

)Wj<br />

Using the definition (A.23) of the permutation tensor pq¦/ j we can write the<br />

components of (1.11) explicitly as:<br />

r `<br />

@<br />

l<br />

rq<br />

<br />

) <br />

r<br />

@<br />

l<br />

<br />

<br />

) <br />

r 4<br />

@<br />

l<br />

n <br />

1.2 Thermodynamic Variables<br />

Classical thermodynamics was formulated for equilibrium states. Even though<br />

fluid flow is not generally in equilibrium, we can apply thermodynamical concepts<br />

using a quasi-equilibrium approximation, which assumes that the flow changes<br />

slowly enough, so that at each point a local thermodynamic equilibrium is reached.<br />

1.2.1 Equations of state<br />

The equation of state relates important thermodynamic variables, such as pressure,<br />

X , temperature, s , <strong>and</strong> density, ; :<br />

4


(1.14) X<br />

u<br />

†<br />

<br />

l<br />

P<br />

(1.15) R „<br />

(1.16) R ˆ<br />

<br />

<br />

<br />

<br />

<br />

<br />

; Ats<br />

`<br />

P<br />

`<br />

R<br />

u<br />

†<br />

><br />

l<br />

P<br />

<br />

@<br />

6 CHAPTER 1. PROPERTIES AND VARIABLES<br />

such as an ideal gas law:<br />

A where is the gas constant. Depending on the number of parameters in the<br />

system, there can be several equations of state so as to keep the number of<br />

independent uv5wyx variables, to comply to the Gibbs rule [1]:<br />

uzU{}| where is the u<br />

P ~<br />

number of components <strong>and</strong> is the number of<br />

1}<br />

phases<br />

r,‚<br />

in the<br />

system. For example, for the<br />

4<br />

uz¨{€|<br />

water-vapor mixture we have <strong>and</strong><br />

. Thus the state of water-vapor mixture @ can be completely<br />

described by the pressure or temperature only.<br />

, then uƒv5wyx<br />

P ~<br />

uvmwmx<br />

uzU{}|<br />

P ~<br />

1.2.2 Energy<br />

The first law of thermodynamics expresses the principle of energy conservation<br />

related to the thermodynamic variables: internal energy 1 , „ , heat, … , <strong>and</strong> work,<br />

:<br />

that is, the change of energy of the fluid element is equal to the heat inflow plus<br />

the work done on that element. Rewriting this in terms of specific values, i.e.<br />

values related to the unit of mass (ˆ‰'qŠ'*‹ ), we have:<br />

R …‡><br />

R ŠŒ> R ‹<br />

Considering only the mechanical work, we have<br />

! :< <br />

@ : ;<br />

where is the specific volume:<br />

, <strong>and</strong> the minus sign signifies<br />

that the work done on the system is positive when it is compressed <strong>and</strong> negative<br />

when it is inflated. From the definition of entropy [1] we have:<br />

R ‹<br />

X RŽ<br />

1 In thermodynamics books it is usually denoted by ‘ , but we reserve this symbol for fluid velocity


(1.17) R ˆ<br />

(1.18) R ˆ<br />

(1.19) ˆ<br />

(1.20) ’<br />

(1.21) R ’<br />

`<br />

<br />

¢<br />

<br />

<br />

<br />

<br />

ˆ<br />

`<br />

<br />

<br />

1.2. THERMODYNAMIC VARIABLES 7<br />

Thus (1.16) becomes:<br />

R Š<br />

s R .<br />

s R . `<br />

RQ X<br />

which can be expressed in terms of density:<br />

s R . ><br />

X<br />

;<br />

r R ;<br />

This relation implies that ˆ is a function of . <strong>and</strong> ; only:<br />

1 . ' ; 4<br />

Relation (1.19) constitutes the equation of state as dictated by the energy conservation<br />

law.<br />

as:<br />

Another form of energy identified in thermodynamics is enthalpy. It is defined<br />

X<br />

;<br />

ˆ“>”X<br />

ˆ“><br />

<strong>and</strong> analogously to (1.17), we obtain:<br />

X R<br />

;<br />

s R . ><br />

“R X<br />

s R . ><br />

so-called s<br />

Expressing the entropy, , from relations (1.17) <strong>and</strong> (1.21) we obtain the<br />

.<br />

relations, well known in thermodynamics:<br />

.<br />

s R . <br />

R ˆ•>”X RŽ<br />

R ’<br />

“R X<br />

s R .


(1.22) – v<br />

(1.23) R ˆ<br />

(1.24) –<br />

(1.25) R ’<br />

P<br />

Y ˆ ¢˜—<br />

sZ v Y<br />

<br />

’ ¢˜—§Y<br />

sZ Y<br />

<br />

–<br />

P<br />

P<br />

8 CHAPTER 1. PROPERTIES AND VARIABLES<br />

1.3 <strong>Fluid</strong> Properties<br />

<strong>Fluid</strong> properties are given as free parameters in a physical law. Along with the<br />

space coordinates <strong>and</strong> time they form the set of independent variables of the<br />

system.<br />

1.3.1 Thermodynamic properties<br />

Specific heat at constant volume – v determines the amount of thermal energy<br />

that is needed to be transfered to the substance to heat it up by one degree,<br />

while keeping it at a constant volume:<br />

For ideal gases the internal energy, ˆ depends only on temperature, <strong>and</strong> the<br />

relation above can be rewritten as:<br />

– v R s<br />

Specific heat at constant – pressure, , is defined as the amount of energy<br />

needed to be supplied to the substance to heat it up by one degree at constant<br />

pressure:<br />

P<br />

where the enthalpy, ’ , is used instead of internal energy, ˆ , since it accounts for<br />

the work done by the pressure forces to extend/compress the substance. For an<br />

ideal gas this translates to:<br />

R s


(1.26) š$›M)œ<br />

(1.27)<br />

6 ¦/<br />

¢<br />

¥<br />

R<br />

¥<br />

1.3. FLUID PROPERTIES 9<br />

1.3.2 Transport Properties<br />

A general form for the transport law is presumed to obey the gradient approximation<br />

in a form:<br />

1 ! 4 £Ÿž¡ ¢J£<br />

1 ! 4<br />

where is the kinematic or thermodynamic property, which represents a dependent<br />

variable of the problem, flux is the amount of property passed through a<br />

unit area per unit time <strong>and</strong> gradient is the direction of maximum change in property.<br />

Both flux <strong>and</strong> gradient are a vectors. The coefficient of proportionality,<br />

!<br />

represents the transport property controlling the transport process (1.26).<br />

The coefficient of viscosity<br />

The coefficient of viscosity is introduced to quantify the process of momentum<br />

transport.<br />

In elasticity theory the resistance force is proportional to defor-<br />

Newtonian fluid:<br />

mation:<br />

¤¥<br />

¢ £ " #<br />

R ˆ§š<br />

=<br />

Similarly in fluid mechanics the resistance to fluid motion is proportional to<br />

the velocity change in the direction normal to the fluid motion (strain).<br />

–ˆt¦<br />

¨ " ¢<br />

. " ¢J£Ž#<br />

=<br />

Since generally the stresses <strong>and</strong> strains are tensors (Sec.1.1.4, 2.2.1), the<br />

relation above is usually written as:<br />

ˆ . . ¦<br />

¦ . ¦7<br />

Definition 1.3.1 Newtonian fluid<br />

A fluid with the linear relationship between stresses <strong>and</strong> strains, like (1.27)<br />

is called a Newtonian fluid.


%<br />

6<br />

r 8 )<br />

F<br />

F<br />

¢S©ª<br />

The kinematic viscosity is defined as 9<br />

(1.28)<br />

6<br />

¦<br />

…<br />

<br />

1<br />

<br />

…<br />

¦<br />

8 R ) <br />

® R<br />

R<br />

s<br />

ˆ<br />

…<br />

<br />

)<br />

¢<br />

ˆ<br />

4<br />

r<br />

10 CHAPTER 1. PROPERTIES AND VARIABLES<br />

Definition 1.3.2 Coefficient of viscosity<br />

The coefficient of viscosity is introduced as a proportionality constant between<br />

the shear stress <strong>and</strong> strain. Considering one component of stress tensor<br />

at: i=1, j=2, we have:<br />

r<br />

¦_«%<br />

¦­<br />

Using a more common notation for vector components: )<br />

('q®G',+- we have:<br />

'*‹¬-Q' )('<br />

,¯<br />

Viscosity usually decreases with temperature for liquids <strong>and</strong> increases for<br />

rarefied gases [2].<br />

Non-newtonian fluid: For non-Newtonian fluids the relation between the stress<br />

<strong>and</strong> the strain is non-linear, for example<br />

6 ¦/<br />

¦ . B<br />

¦/<br />

Thermal Conductivity<br />

Following the gradient approximation (1.26), we presume the heat transport to<br />

obey the relation:<br />

£ "54<br />

¢J£<br />

1 "<br />

¢


(1.30) º<br />

P<br />

j ¦ <br />

¡<br />

<br />

r<br />

ˆ<br />

‹<br />

¥<br />

£ . .<br />

"<br />

’<br />

`§½¿¾ ¦ ; j `]½Å¾ ¦ 1 ; Ä<br />

ˆ<br />

<br />

4<br />

"<br />

’<br />

4<br />

±<br />

°<br />

P<br />

R<br />

"<br />

ˆ<br />

±<br />

1<br />

"<br />

’<br />

.§. £<br />

"<br />

Ä<br />

®<br />

4<br />

)<br />

¢<br />

ˆ<br />

Ä<br />

j<br />

1.3. FLUID PROPERTIES 11<br />

° ²<br />

conductivity. Dimensionality of , which we shall ± denote as °<br />

by the dimensions of the units in the equation above:<br />

, will be determined<br />

¢ °<br />

¢J£ "<br />

° ²<br />

XWˆ<br />

" #<br />

ž 1<br />

›¨ˆ¡=<br />

›¨ˆ¡=<br />

r <br />

° ²³§·k †<br />

± : 1 ža¸ 4 ° ¹: 1 . ¸ 4<br />

then in SI units: .<br />

Pr<strong>and</strong>tl number: momentum transport / heat transport<br />

¢» 8 –<br />

where –<br />

is specific heat at constant pressure (Sec.1.3.1).<br />

Mass Diffusivity<br />

The gradient approximation (1.26) for the mass transport can be written as:<br />

1 £ . . 4 ¼ ¢


where –<br />

È : –,v<br />

P<br />

–<br />

Ç<br />

<br />

ˆ<br />

¨<br />

<br />

–<br />

9 <br />

½<br />

– ;<br />

P<br />

X —ÃÈ“Y<br />

; Y<br />

°<br />

½<br />

Ê<br />

r<br />

Ä<br />

12 CHAPTER 1. PROPERTIES AND VARIABLES<br />

(1.32)<br />

j ¦ <br />

¡<br />

j 4<br />

`§½ ; j ¾ ¦ dfÆ 1<br />

<strong>and</strong> relating the mass flux to the fluid velocity across the boundary:<br />

we have:<br />

j ¦ ; j ! ¦§¡<br />

,<br />

j ¦ `§½¿¾ ¦ dfÆ 1<br />

!<br />

Ä<br />

j 4<br />

There are two non-dimensional numbers relating the momentum-to-mass<br />

<strong>and</strong> mass-to-heat transport processes:<br />

Schmidt number:<br />

momentum transport / mass transort<br />

Lewis number:<br />

heat transport / mass transport<br />

1.3.3 Other properties<br />

Speed of sound<br />

The speed of sound for compressible flow is defined as the rate of propagation of<br />

small pressure perturbations, <strong>and</strong> it is found to be equal to [3]:<br />

F}É


(1.33) Ë<br />

<br />

ˆ<br />

À<br />

<br />

— Y X ;<br />

; Ê Y<br />

Y ; —<br />

s Y<br />

È<br />

r<br />

À<br />

À<br />

T<br />

<br />

T<br />

1.4. PROBLEMS 13<br />

Bulk modulus<br />

The bulk modulus expresses the change of density with increasing pressure at a<br />

constant temperature:<br />

¸¢<br />

; –<br />

<strong>and</strong> is used in acoustic problems.<br />

Coefficient of thermal expansion<br />

The Coefficient of thermal expansion relates density to temperature changes:<br />

@ `<br />

;<br />

P<br />

<strong>and</strong> is used in the problems of natural convection.<br />

1.4 Problems<br />

Problem 1.4.1 Mass diffusivity in terms of concentration<br />

Show how to obtain (1.31) from (1.32).<br />

Problem 1.4.2 Lubrication<br />

A plate of mass<br />

incline at Î angle<br />

with viscosity 8<br />

l TÅÍ<br />

with an area<br />

°<br />

slides down a long<br />

–<br />

, on which there is a film of oil of ¡Ì<br />

IO@ ¿ ’ thickness ,<br />

. Assuming the plate does not deform the oil<br />

{ T<br />

: 1 ž . 4 `^Ñ•°<br />

ÐÏ<br />

film estimate (1) the terminal sliding velocity , <strong>and</strong> (2) the time required for the<br />

plate to accelerate from rest to Ó T !GÒ of the terminal velocity.<br />

IÕÔ‰Ô


14 CHAPTER 1. PROPERTIES AND VARIABLES


¦<br />

R<br />

"<br />

R<br />

; <br />

R<br />

; R<br />

R<br />

"<br />

R<br />

!<br />

><br />

<br />

T<br />

Ö ; )<br />

–<br />

¥<br />

¦<br />

=<br />

¦<br />

R<br />

£<br />

T<br />

Chapter 2<br />

Fundamental Laws<br />

2.1 Conservation of Mass<br />

2.1.1 General formulation<br />

The conservation of mass dictates that:<br />

Ö<br />

! <br />

= . "<br />

which also means that<br />

The total change of mass inside the control volume will consist of changes<br />

of mass inside the volume because of density changes that may occur at each<br />

point of the flow, <strong>and</strong> the influx or out-flux of mass through the boundary. This can<br />

be expressed as:<br />

Ö<br />

£<br />

= where is the unit normal vector to the R boundary, is the surface area element,<br />

<strong>and</strong> the last integral spans all the boundary of the control volume.<br />

15


(2.1) R<br />

R<br />

"<br />

R<br />

; <br />

R<br />

!<br />

><br />

Ö ; )<br />

Ö 1 ; )<br />

(2.4) )<br />

=<br />

¦ £<br />

R<br />

¦ £Ž¦$ Ö×1 ; ) R<br />

; R<br />

" > ; )<br />

R<br />

; R<br />

"<br />

R<br />

R<br />

¥Ö 1 ; !<br />

><br />

<br />

T<br />

T<br />

R<br />

!<br />

1 ; )<br />

R<br />

16 CHAPTER 2. FUNDAMENTAL LAWS<br />

Let’s define the surface area vector, R<br />

£Ž¦<br />

as:<br />

£‰¦$¢<br />

By Gauss theorem we can convert the last integral to the volume integral:<br />

¦ 4 ¦<br />

<strong>and</strong> finally:<br />

Ö<br />

¦ 4 ¦<br />

¦ 4 ¦ 4<br />

! T<br />

Considering the arbitrary nature of the control volume selection, we conclude:<br />

(2.2)<br />

; <br />

><br />

1 ; )<br />

¦ 4 ¦$ T<br />

This is a general relation of mass conservation valid for both compressible<br />

<strong>and</strong> incompressible flows. Differentiating the second term by parts, <strong>and</strong> using the<br />

relation of substantial differentiation (1.7) the latter can be rewritten as:<br />

(2.3)<br />

¦f ¦G<br />

2.1.2 Constant density flow<br />

For a constant density flow<br />

; T <br />

, <strong>and</strong> from (2.2) it follows:<br />

¦f ¦G<br />

which is also called the continuity equation or incompressibility condition. Vector<br />

field satisfying (2.4) is also called solenoidal or divergence-free.<br />

Another form of this relation can be obtained by combining (2.3) <strong>and</strong> (2.4):<br />

(2.5)<br />

T


(2.6)<br />

; )<br />

(2.7)<br />

; )<br />

<br />

Ø<br />

F<br />

¦<br />

; )<br />

F<br />

Ø<br />

Ø<br />

Ø<br />

F<br />

<br />

<br />

F<br />

<br />

`<br />

F<br />

)<br />

`<br />

`<br />

'<br />

Ø<br />

Ø<br />

r<br />

j<br />

T<br />

Ø<br />

Ø<br />

<br />

F<br />

F<br />

4<br />

<br />

4 <br />

F<br />

F<br />

F<br />

<br />

4 ; r <br />

<br />

)<br />

R ;<br />

¦<br />

R<br />

R<br />

¦<br />

<br />

2.1. CONSERVATION OF MASS 17<br />

2.1.3 Stream function<br />

Let’s introduce the stream function, which is closely related to the mass flow rate.<br />

The stream Ø function , which is a vector in 3D, also called the streamlinevorticity<br />

function is defined such as to satisfy the relation<br />

or, using the nabla operator (A.32):<br />

¦Gp5¦/<br />

<br />

jÙØ­j<br />

¦G£p5¦/<br />

¾ <br />

Øaj<br />

which analogously to (1.13) is<br />

1<br />

r `<br />

rq<br />

Ø <br />

(2.8)<br />

) ;<br />

) ;<br />

r• 1<br />

Ø <br />

1<br />

rq<br />

r 4<br />

In two dimensions the streamline function is defined as<br />

¦$Ú% T<br />

'*ØŒi.e.<br />

a vector normal to the plane, <strong>and</strong> therefore (2.8) are reduced to:<br />

(2.9)<br />

; )<br />

; )<br />

r<br />

r<br />

The following relation between the stream function <strong>and</strong> the mass flow rate<br />

can be shown for a two dimensional case:<br />

£Ž¦<br />

(2.10)<br />

<br />

R <br />

F<br />

r<br />

r“<br />

R Ø<br />

>^Ø<br />

R <br />

; 1 `<br />

>])<br />

R <br />

R <br />

£‰¦<br />

where is the element of the surface normal to the ) velocity R<br />

also be proved more rigorously for a 3D space.<br />

. This relation can<br />

!


(2.11) )<br />

)<br />

)<br />

F<br />

Ö<br />

)<br />

<br />

F<br />

R<br />

R<br />

R<br />

R<br />

"<br />

"<br />

R<br />

Ø<br />

¦<br />

¤<br />

F<br />

¦<br />

`<br />

¤<br />

T<br />

Ø<br />

r<br />

F<br />

T<br />

18 CHAPTER 2. FUNDAMENTAL LAWS<br />

Remark 2.1.1 Existence of stream-function<br />

It should be noted that sometimes, instead of definition (2.6), the streamfunction<br />

is defined from a simpler relation:<br />

¦Gp5¦7<br />

<br />

jØaj<br />

where the density is omitted. In this case the stream function can only be used to<br />

describe incompressible flow. This can be shown by computing the divergence of<br />

velocity ) vector :<br />

¦K ¦<br />

¦f ¦Gp5¦7<br />

y¦G<br />

¦7<br />

which is true due to the symmetric identity (A.27) <strong>and</strong> ØÛj<br />

the symmetry of with<br />

respect to the order of differentiation. This becomes especially obvious in a 2D<br />

case:<br />

jØaj<br />

¦f ¦G<br />

rq r•<br />

r<br />

Thus, in terms of definition (2.11) the stream function can only exist for incompressible<br />

flows.<br />

>^)<br />

2.2 Conservation of Momentum<br />

2.2.1 General formulation<br />

According to Newton’s law a particle of mass<br />

force as:<br />

is accelerated by the action of a<br />

1 ¦ 4 <br />

which will apply to a particle of both constant <strong>and</strong> a variable mass. Applying this<br />

in a small control volume, we have<br />

to a fluid particle of density ; <strong>and</strong> velocity )<br />

(2.12)<br />

1 ; )<br />

¦ 4<br />

! <br />

¦


¦<br />

¤<br />

Ö<br />

Y<br />

"<br />

Y<br />

¤<br />

R —<br />

R<br />

¦<br />

"<br />

R<br />

R<br />

><br />

"<br />

Ö<br />

¦ !<br />

R š<br />

j<br />

><br />

¦ !<br />

R š<br />

j<br />

`<br />

š<br />

¦<br />

<br />

><br />

j<br />

j<br />

¦<br />

R<br />

!<br />

¦<br />

£<br />

j<br />

2.2. CONSERVATION OF MOMENTUM 19<br />

The forces acting on a fluid element come from the possible external forces,<br />

like gravity, electromagnetic fields, etc. (body forces), <strong>and</strong> forces caused by the<br />

interaction of this fluid element with neighboring fluid elements or boundaries<br />

(surface forces). Body forces relate to the unit of volume <strong>and</strong> surface forces relate<br />

to the unit of area.<br />

¤ƒÜ<br />

¤¬Ý<br />

(2.13)<br />

¤ Ü<br />

¦$<br />

¦ Ö<br />

Ö 6 ¦<br />

j R<br />

š where is the volumetric density of the body force . It corresponds to a force<br />

field like electromagnetic, gravity, etc. Generally it can serve as a source term<br />

connecting this equation to other equations.<br />

Definition 2.2.1 Stress tensor<br />

The surface force term 6 ¦/<br />

in (2.13) is called the stress tensor.<br />

Using this definition <strong>and</strong> applying Gauss theorem to the last term in (2.13),<br />

we have:<br />

(2.14)<br />

<br />

R j<br />

¦<br />

Ö 6 ¦<br />

Now, comparing (2.14) with (2.12), we have<br />

¦ 4a` 6 ¦<br />

! T<br />

1 ; )<br />

<br />

j<br />

Using the fact the control volume was chosen arbitrarily, the integral sign<br />

could be dropped, <strong>and</strong> we have:<br />

(2.15)<br />

1 ; )<br />

<br />

jÛ>Þš<br />

¦ 4 6 ¦<br />

Using the definition of substantial derivative (1.7), we have:<br />

(2.16)<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

<br />

jc>\š<br />

¦ 4<br />

6 ¦<br />

¦<br />

>])Wj


(2.17)<br />

6 ¦/k `§ß ¦/<br />

(2.18)<br />

6 ¦7Œ `§ß ¦/<br />

(2.19) ã 6 ¦/Œ¢ l 8¬. ¦/ `]ß ¦/Ùá<br />

ã 6 ¦7k l 8¬. ¦/ 8 1<br />

)<br />

(2.20)<br />

¦<br />

Y<br />

"<br />

Y<br />

Y<br />

"<br />

Y<br />

¦<br />

<br />

<br />

`<br />

X<br />

`<br />

X<br />

¦<br />

20 CHAPTER 2. FUNDAMENTAL LAWS<br />

Using the hypothesis of Newtonian fluid (1.27), <strong>and</strong> general considerations<br />

of symmetry for the case of isotropic <strong>and</strong> homogeneous fluid, a general relation<br />

can be written as [2, p.66]:<br />

between the stress tensor 6 ¦7<br />

<strong>and</strong> the strain tensor . ¦/<br />

<br />

j<br />

where X is the pressure, á is the coefficient of bulk viscosity, which is only important<br />

for compressible flows.<br />

Sometimes it is convenient to separate the stress tensor (2.17) into the<br />

pressure-related <strong>and</strong> viscous parts:<br />

X><br />

)Wj<br />

l 8¬. ¦/ `àß ¦/Ùá<br />

Xâ>Úã 6 ¦7<br />

where ã 6 ¦/<br />

is the viscous stress tensor defined as<br />

<br />

j<br />

Parameter á is the coefficient of bulk viscosity, which can only be important for<br />

variable density flows [2]. Thus, for incompressible flows (2.19) becomes:<br />

)Wj<br />

¦f<br />

m ¦ 4<br />

where we used the definition of strain rate tensor (1.9). Using the definition of the<br />

viscous stress (2.19) we can write (2.16) as:<br />

>^)<br />

(2.21)<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

<br />

j<br />

š When<br />

(2.21) as:<br />

represents the gravity forces: š<br />

¦ 4<br />

¦<br />

>])Wj<br />

>Þš<br />

¦H ; ¦<br />

we can rewrite equation<br />

6 ¦<br />

j >äã<br />

(2.22)<br />

¦<br />

¦ 4<br />

> ; ])Wj<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

6 ¦<br />

j ã<br />

<br />

j


)<br />

¦<br />

)<br />

ã<br />

j<br />

<br />

<br />

j<br />

<br />

j<br />

j<br />

p<br />

j<br />

j<br />

)<br />

9 )<br />

j<br />

j<br />

p<br />

j<br />

j<br />

j<br />

1<br />

)<br />

P<br />

1<br />

)<br />

P<br />

1<br />

)<br />

P<br />

<br />

æ<br />

1<br />

`<br />

<br />

`<br />

`<br />

æ<br />

`<br />

æ<br />

X<br />

T<br />

)<br />

¦ <br />

> ;<br />

)<br />

æ<br />

)<br />

)<br />

<br />

P<br />

æ<br />

æ<br />

4<br />

P<br />

4<br />

P<br />

¦<br />

j<br />

2.2. CONSERVATION OF MOMENTUM 21<br />

2.2.2 Constant density flow<br />

The viscous term in equation (2.22) can be further simplified for constant density<br />

flows. Using (2.20) we can rewrite it as:<br />

l 8(. ¦<br />

8 1<br />

¦K<br />

¦ 4 <br />

(2.23)<br />

<br />

j<br />

jc>^)Wj<br />

6 ¦<br />

¦f<br />

j*jÛ>^)Wj<br />

<br />

j<br />

4 ¦ 4 8 )<br />

8 1<br />

¦<br />

4 8 1<br />

¦f<br />

¦f<br />

j*jÛ><br />

)Wj<br />

j*j<br />

where we used the continuity )åj relation:<br />

Substituting this into (2.22), we have:<br />

<br />

j<br />

for incompressible flow (2.4).<br />

(2.24)<br />

¦f<br />

j<br />

¦f<br />

j*j<br />

>^)WjÙ)<br />

which is an incompressible form of momentum equation, also referred to as the<br />

Navier-Stokes equation (NS).<br />

2.2.3 Vorticity formulation<br />

Our objective will be to replace the velocity vector in a constant density NS equation<br />

(2.24) with a vorticity vector (1.11). For this purpose consider a cross product<br />

between nabla operator (A.32) <strong>and</strong> vorticity vector:<br />

(2.25)<br />

p5¦/<br />

¾ <br />

¼p5¦/<br />

<br />

n­j<br />

j n­j<br />

Using (1.12) the cross product (2.25) can be rewritten as:<br />

p5¦/<br />

“<br />

pm¦/<br />

1 p<br />

454 <br />

@<br />

l<br />

j n­j<br />

P æ<br />

p5¦/<br />

@<br />

l<br />

(2.26)<br />

Pæ<br />

¦/ p<br />

@<br />

l<br />

Using the tensor identity (A.29):<br />


p<br />

<br />

j<br />

j<br />

1<br />

)<br />

P<br />

ß ¦ 1<br />

P<br />

<br />

`<br />

æ<br />

@<br />

l<br />

æ<br />

)<br />

ß ¦ 1<br />

P<br />

P<br />

)<br />

p<br />

j<br />

æ<br />

j<br />

p5¦<br />

Pæ<br />

4 <br />

P<br />

1<br />

)<br />

P æ<br />

`]ß ¦<br />

P æ<br />

(2.29) )<br />

<br />

T<br />

<br />

j<br />

`<br />

æ<br />

)<br />

æ<br />

<br />

<br />

ß ¦ <br />

P<br />

æ<br />

ß ¦ 1<br />

P<br />

)<br />

æ<br />

P<br />

ß<br />

j<br />

æ<br />

æ<br />

æ<br />

4a`]ß ¦<br />

æ P<br />

`]ß ¦<br />

æ P<br />

@<br />

l<br />

1<br />

)<br />

)<br />

)<br />

<br />

)<br />

<br />

P<br />

)<br />

)<br />

)<br />

æ<br />

)<br />

æ<br />

æ<br />

ß<br />

P<br />

j<br />

P<br />

P<br />

1<br />

)<br />

P P<br />

<br />

æ<br />

1 <br />

) )<br />

<br />

><br />

)<br />

)<br />

)<br />

P<br />

`<br />

æ<br />

`<br />

æ<br />

æ<br />

P<br />

)<br />

)<br />

æ<br />

)<br />

æ<br />

4<br />

P<br />

454<br />

P<br />

)<br />

æ<br />

4<br />

P<br />

)<br />

T<br />

22 CHAPTER 2. FUNDAMENTAL LAWS<br />

p5¦/<br />

ß <br />

`]ß <br />

which can be rewritten as<br />

¦/ p<br />

ß <br />

`àß ¦<br />

ß <br />

P<br />

to match the indexes, we can simplify the cross product (2.25):<br />

P æ<br />

@<br />

l<br />

@<br />

l<br />

¦/p<br />

ß <br />

`]ß ¦<br />

ß <br />

4 1<br />

P æ<br />

ß <br />

ß <br />

ß <br />

ß <br />

ß <br />

ß ¦<br />

ß <br />

@<br />

l<br />

¦f} `<br />

m ¦/ `<br />

m ¦/<br />

¦f} 4<br />

(2.27)<br />

>^)<br />

¦f€ `<br />

m ¦/<br />

And finally (2.26) becomes:<br />

(2.28)<br />

p5¦/<br />

“<br />

¦f€ `<br />

m ¦/<br />

jn­j<br />

For constant density flows it follows from (2.4) that )<br />

, <strong>and</strong><br />

m ¦7 1<br />

m 4 ¦G<br />

j n­j<br />

¦f}¼p5¦/<br />

<br />

Thus we can replace the diffusive ) term<br />

above.<br />

¦f}<br />

in (2.24) with the cross-product<br />

¦f<br />

Now let’s consider the )<br />

convective term in (2.24). Using the constant<br />

()<br />

density assumption ) <strong>and</strong> now considering the cross product of the type<br />

¦f ¦c<br />

n­j , we can repeate the steps as in (2.26) <strong>and</strong> obtain 1 :<br />

pm¦/<br />

jÙ)<br />

(2.30)<br />

p5¦/<br />

çžEžEžQ<br />

¦f `<br />

4 ¦<br />

@<br />

l<br />

jÙ)<br />

n­j<br />

1 See Problem 2.7.1


j<br />

<br />

ë<br />

¦<br />

> )<br />

¦<br />

> )<br />

¦<br />

> )<br />

1 <br />

) )<br />

1 <br />

) )<br />

1 <br />

) )<br />

)<br />

<br />

)<br />

><br />

<br />

<br />

)<br />

<br />

><br />

<br />

<br />

<br />

ë<br />

1 <br />

) )<br />

j<br />

><br />

<br />

)<br />

<br />

4<br />

T<br />

2.2. CONSERVATION OF MOMENTUM 23<br />

Thus<br />

(2.31)<br />

@<br />

l<br />

¦Kk£p5¦7<br />

4 ¦<br />

jè)<br />

n­jc><br />

And substituting (2.29) <strong>and</strong> (2.31) into the momentum equation (2.24) we<br />

have:<br />

(2.32)<br />

4 ¦<br />

p5¦7<br />

<br />

¦<br />

@<br />

l<br />

¦ ` ;Ãé F X<br />

j)<br />

n­j<br />

j naj<br />

9 pm¦/<br />

Rearranging the terms, <strong>and</strong> using the relation <br />

¦f<br />

(A.34), we have:<br />

ß ¦/<br />

(2.33)<br />

9 p5¦/<br />

@ 1<br />

l<br />

4<br />

` ¡<br />

4 ¦$ ` p5¦/ <br />

<br />

> ;Ãé F X<br />

jÙ)<br />

n­jê><br />

jn­j<br />

This equation can also be rewritten as:<br />

(2.34)<br />

@<br />

X ;<br />

4a` è<br />

4 ¦$¼p5¦/<br />

`<br />

@ 1<br />

l<br />

1 9 n­j<br />

naj<br />

This is a NS equation in vorticity formulation for the incompressible flow 2 .<br />

2.2.4 Potential flow<br />

¦<br />

Let’s consider the irrotational flow where the vorticity vector (n is zero ). This<br />

flow is also called potential flow, since the velocity vector can be replaced by a<br />

gradient of a ë scalar function, , also called a velocity potential function:<br />

¦G<br />

¦<br />

This is possible, because the gradient of a scalar function also satisfies the condition<br />

of zero vorticity, which follows from the definition of the vorticity vector (1.12)<br />

<strong>and</strong> the symmetry of the second derivative ë of with respect to order of differentiation,<br />

ë<br />

<br />

<br />

j<br />

:<br />

2 Note that we achieved only a partial success in our objective to replacing the velocity vector with the<br />

vorticity vector, but that’s the best we can do.


(2.35) )<br />

p<br />

Ý<br />

n<br />

¦<br />

><br />

¦<br />

> )<br />

p<br />

j<br />

Ý<br />

1<br />

)<br />

1 <br />

) )<br />

`<br />

s<br />

Ý<br />

<br />

¦<br />

ë<br />

<br />

<br />

><br />

Ý<br />

s<br />

T<br />

j<br />

<br />

1<br />

ë<br />

–<br />

¥<br />

<br />

j<br />

`<br />

ë<br />

T<br />

24 CHAPTER 2. FUNDAMENTAL LAWS<br />

@<br />

l<br />

m<br />

j<br />

@<br />

l<br />

4 <br />

j<br />

¦G<br />

pm¦/<br />

4 <br />

p5¦/<br />

<br />

)åj<br />

In the case of steady state incompressible potential flow the continuity condition<br />

(2.4) translates into the Laplace equation for the velocity potential:<br />

¦f ¦G<br />

¦ì¦G<br />

Thus, the solution of the problem in this case is reduced to finding a single scalar<br />

function ë from equation (2.35). Another important relation in this case can be obtained<br />

from equation (2.34), which, after eliminating time derivatives <strong>and</strong> vorticity<br />

terms reduces to:<br />

(2.36)<br />

X<br />

;<br />

4í` è<br />

@<br />

l<br />

= . "<br />

which is a weak formulation of the Bernoulli’s Equation 3 .<br />

2.2.5 2D limit<br />

Let’s rewrite (2.33) as<br />

(2.37)<br />

¡ ¦G<br />

` p5¦/<br />

<br />

jè)<br />

n­jc><br />

j naj<br />

Where we denoted the term in parentheses by . The equality above is a first<br />

. Let’s now form a cross product between<br />

this equality <strong>and</strong> ¡ Ý<br />

¦<br />

of the type p î¦ ¾ ¦$p î¦ ¦f <br />

(see also (A.32)):<br />

j p5¦/<br />

rank tensor equality with terms of type s<br />

9 pm¦/<br />

(2.38)<br />

¾ î¦<br />

)<br />

î¦ ¾ ¨¡ ¦$£p<br />

î¦ ¾ 1 ` p5¦/<br />

4<br />

Using the definition of ¾ ¦<br />

(A.32), we get:<br />

jÙ)<br />

n­jê><br />

jn­j<br />

9 p5¦/<br />

3 See also (2.78)


p<br />

Ý<br />

Ý ¤<br />

ß<br />

Then the term ¤<br />

Ý<br />

j<br />

<br />

)<br />

><br />

<br />

p<br />

Ý<br />

j<br />

<br />

><br />

Ý<br />

)<br />

<br />

<br />

<br />

n<br />

j<br />

`<br />

j<br />

Ý<br />

Ý<br />

Ý<br />

Ý<br />

p<br />

Ý<br />

1 Ý<br />

)<br />

)<br />

<br />

n<br />

<br />

Applying the same manipulations to the Ä<br />

Ä<br />

Ý<br />

<br />

¤<br />

j<br />

<br />

Ý<br />

<br />

'<br />

T<br />

)<br />

Ý<br />

<br />

`<br />

<br />

n<br />

)<br />

<br />

Ý<br />

Ý<br />

ß<br />

Ý<br />

n<br />

n<br />

n<br />

j<br />

Ý<br />

Ý<br />

)<br />

`<br />

)<br />

<br />

n<br />

Ý<br />

n<br />

)<br />

Ý<br />

` ³<br />

)<br />

)<br />

j<br />

Ý<br />

Ý<br />

F<br />

n<br />

p<br />

Ý<br />

Ý<br />

Ý<br />

n<br />

n<br />

Ý<br />

j<br />

><br />

Ý<br />

Ý<br />

r<br />

'<br />

T<br />

)<br />

1<br />

)<br />

T<br />

T<br />

Ý<br />

Ý<br />

Ý<br />

<br />

<br />

<br />

n<br />

<br />

n<br />

<br />

n<br />

)<br />

Ý<br />

Ý<br />

Ý<br />

2.2. CONSERVATION OF MOMENTUM 25<br />

9 pm¦/<br />

¦f <br />

î¦M¡ ¦ì£p<br />

î¦ 1 ` p5¦7<br />

4 <br />

î¦<br />

)<br />

(2.39)<br />

n­j<br />

> 9 p5¦7<br />

n­j<br />

jè)<br />

n­jê><br />

j naj<br />

` p5¦7<br />

î¦ 1<br />

4 <br />

î¦<br />

y<br />

£¤<br />

Ý<br />

Ý<br />

where we used abbreviations<br />

is a symmetric tensor then by (A.2.22), we have p<br />

¤<br />

Noticing that by the definition of vorticity n vector:<br />

the equation (2.39) to the form:<br />

'*Ä<br />

for the last two terms on the RHS. Since<br />

.<br />

î¦M¡ƒ ¦ì<br />

£p<br />

î¦<br />

¦K <br />

¡¬ ¦/<br />

, we can reduce<br />

>\Ä<br />

(2.40)<br />

Using the permutation property (A.24): p Ý<br />

, we have:<br />

£¤<br />

>\Ä<br />

î¦$<br />

` p<br />

¦ì(pm¦<br />

4 <br />

` pm¦/<br />

p5¦<br />

1<br />

naj<br />

4 $ ` 1 ß <br />

`]ß y ß<br />

4 1<br />

naj<br />

4 <br />

`tß <br />

1<br />

naj<br />

4 <br />

ß m ß<br />

naj<br />

4 $ ` 1<br />

4 <br />

<br />

<br />

³ <br />

>])<br />

³ <br />

>^)<br />

where we used the incompressibility condition (2.4): )<br />

ï% T<br />

In a 2D n<br />

limit we have<br />

the 3-rd component of the equation above, i.e.<br />

'mnŒ- <strong>and</strong> )<br />

¦aï%<br />

¦f ¦G<br />

'q)<br />

.<br />

- . Let’s consider only<br />

<br />

`<br />

³ <br />

<br />

>^)<br />

(2.41)<br />

<br />

n<br />

¼¤<br />

>^Ä <br />

can be simplified as:<br />

<br />

<br />

term, we have:<br />

T ` T<br />

>^)<br />

(2.42)<br />

yÚžEžEžQ 9 1<br />

î 4<br />

9 p5¦/<br />

pm¦<br />

n­j<br />

And in the 2D limit:


(2.44) )<br />

<br />

@ ß ¦<br />

±<br />

P<br />

l<br />

<br />

ß ¦ 1<br />

P<br />

Ø<br />

æ<br />

æ<br />

æ<br />

`]ß ¦<br />

æ P<br />

(2.46) n<br />

<br />

F<br />

r<br />

n<br />

æ<br />

<br />

<br />

n<br />

<br />

FðF<br />

n<br />

<br />

P<br />

<br />

n<br />

P<br />

Ø<br />

<br />

4<br />

Ø<br />

<br />

æ<br />

@<br />

l<br />

)<br />

æ<br />

¦<br />

n<br />

`<br />

j<br />

n<br />

n<br />

`<br />

P<br />

òß<br />

P<br />

1<br />

Ø<br />

<br />

`<br />

)<br />

<br />

<br />

Ø<br />

j<br />

j<br />

@<br />

l<br />

j<br />

P<br />

F<br />

`<br />

`<br />

Ø<br />

n<br />

n<br />

` 9 n<br />

æ<br />

Ø<br />

r<br />

j<br />

æ<br />

j<br />

P<br />

Ø<br />

Ø<br />

`<br />

P<br />

æ<br />

<br />

P<br />

æ<br />

æ<br />

`<br />

P<br />

Ø<br />

<br />

p<br />

j<br />

j<br />

æ<br />

ß<br />

Ø<br />

j<br />

æ<br />

¦ ß<br />

P<br />

T<br />

T '<br />

¦ ß<br />

P<br />

Ø<br />

Ø<br />

4<br />

Ø<br />

æ<br />

Ø<br />

Ø<br />

æ<br />

æ<br />

<br />

P<br />

<br />

P<br />

æ<br />

<br />

P<br />

<br />

P<br />

j<br />

j<br />

4<br />

4<br />

j<br />

²<br />

j<br />

4<br />

j<br />

'<br />

T<br />

26 CHAPTER 2. FUNDAMENTAL LAWS<br />

9 1<br />

rðr 4<br />

Ä <br />

ð<br />

9 1}T `<br />

rðr 4<br />

FðF<br />

FðF<br />

` 9 1<br />

rðr 4 <br />

¦Õ¦<br />

After substituting ¡ , Ä into (2.41), it becomes:<br />

>on<br />

(2.43)<br />

9 n<br />

¦ `<br />

¦ì¦<br />

which is the 2D limit of (2.37).<br />

Stream-function formulation<br />

By the definition of the stream-function (2.6) we have:<br />

jØaj<br />

¦Gp5¦7<br />

<br />

Substituting (2.44) into (1.12) we can find relation between the vorticity vector<br />

<strong>and</strong> the stream-function:<br />

@<br />

l<br />

¦$<br />

p5¦/<br />

1 p<br />

pñ<br />

P æ<br />

Pæ<br />

1 p<br />

¦/ p<br />

¦/pñ<br />

@<br />

l<br />

Pæ<br />

Pæ<br />

ß <br />

`]ß ¦<br />

ß <br />

1 ß<br />

ß ¦<br />

`àß<br />

@<br />

l<br />

ß <br />

ß <br />

ß ¦<br />

jc><br />

m ¦/ `<br />

¦f} `<br />

¦f<br />

¦<br />

(2.45)<br />

j*jê>]Øaj<br />

m ¦/ `<br />

¦f}<br />

In the 2D limit:<br />

the first of the last two terms in (2.45) will Ø vanish:<br />

¦ò«%<br />

'q<br />

- , )<br />

¦ò«%<br />

'*)<br />

- , n<br />

¦òÐ% T<br />

y k<br />

<br />

, <strong>and</strong> we have:<br />

'5nŒ- , <strong>and</strong> Ø<br />

«% T<br />

'*،- ,<br />

¦ì¦


1<br />

<br />

¦<br />

'<br />

(2.48) Ø<br />

<br />

Ø<br />

<br />

)<br />

<br />

<br />

n<br />

)<br />

¦<br />

Ø<br />

<br />

T<br />

`<br />

T<br />

1<br />

<br />

¦<br />

'<br />

2.2. CONSERVATION OF MOMENTUM 27<br />

Now (2.43) can be rewritten in terms of the stream-function only:<br />

(2.47)<br />

<br />

jqj<br />

9 Ø<br />

Thus, the flow is now completely defined by a<br />

"54<br />

Ø<br />

scalar field , which<br />

is obtained as a solution of (2.47). Note that since (2.47) contains 4th order<br />

"54<br />

Ø<br />

derivatives of , it involves more complex boundary conditions, <strong>and</strong> poses<br />

higher differentiability ë requirements on .<br />

In the case of irrotational (n flow<br />

Laplace equation for the stream-function:<br />

), equation (2.46) reduces to the<br />

¦<br />

¦ì¦<br />

j*j<br />

j*j<br />

¦Õ¦<br />

with the boundary conditions derived from the relation between Ø <strong>and</strong> the velocity<br />

field (2.9). Solving the equation for stream function is usually preferred over solving<br />

an equation for the vorticity, since the velocity field can be obtained from the<br />

stream function by a simple differentiation of type (2.6) or (2.9), whereas obtaining<br />

the velocity field from the vorticity as in (1.12) would require a more laborious<br />

integration.<br />

2.2.6 Viscous limit<br />

Consider the incompressible NS equation (2.24). In the viscous limit we shall<br />

assume the viscous term to be much larger than the convective term. Thus in the<br />

viscous formulation we shall simply neglect the convective terms:<br />

(2.49)<br />

¦<br />

¦G 9 )<br />

¦f<br />

j*j<br />

` º<br />

;<br />

Using the expression of vorticity vector (1.12), we can express the above<br />

equation in terms of vorticity vector only:<br />

(2.50)<br />

9 n<br />

m<br />

j*j<br />

k<br />

which is an incompressible viscous limit of the NS equation in vorticity formulation<br />

(see Problem.2.7.3).


j<br />

)<br />

p<br />

)<br />

j<br />

Y<br />

"<br />

Y<br />

¾ R m¦<br />

R<br />

j<br />

¦<br />

)<br />

)<br />

R<br />

n<br />

R<br />

R<br />

¦<br />

"<br />

><br />

"<br />

p<br />

p<br />

j<br />

j<br />

<br />

)<br />

)<br />

T<br />

;<br />

<br />

j<br />

j<br />

`<br />

º<br />

º<br />

)<br />

T<br />

)<br />

28 CHAPTER 2. FUNDAMENTAL LAWS<br />

In the case of steady state flow this equation simplifies to a Laplace equation<br />

for the vorticity vector:<br />

m<br />

j*j<br />

2.2.7 Inviscid limit<br />

The fluid with a zero viscosity is called an ideal fluid, <strong>and</strong> the flow of such a fluid<br />

is called inviscid. Consider equation (2.22) in the limit of inviscid flow, when the<br />

viscous tensor, 6 ¦<br />

j , vanishes:<br />

(2.51)<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

¦ 4<br />

¦<br />

>])Wj<br />

In the incompressible limit it will simplify to:<br />

(2.52)<br />

¦<br />

` º<br />

¦f<br />

j<br />

>])åj)<br />

;<br />

This is Euler equation for inviscid incompressible flow. It can also be rewritten<br />

in terms of substantial derivative (1.8):<br />

(2.53)<br />

¦<br />

R )<br />

` º<br />

Conservation of vorticity<br />

It can be shown that the inviscid flow preserves vorticity. For this purpose let’s<br />

form a vector product between the nabla operator <strong>and</strong> equation (2.53):<br />

m¦<br />

¦/k<br />

¦G<br />

m¦<br />

¦Kp<br />

where the last equality is due to the symmetry º of<br />

according to:<br />

y¦<br />

transform the j term p<br />

¦f<br />

¦/<br />

<strong>and</strong> identity (A.27). We can<br />

" )<br />

(2.54)<br />

¦/<br />

m ¦åp<br />

m¦ 1<br />

¦f `<br />

y ¦ 4<br />

l p<br />

m¦<br />

¦fk¼p<br />

m¦<br />

¦f


T<br />

Y<br />

"<br />

Y<br />

ó ¦<br />

(2.56)<br />

j<br />

Y<br />

"<br />

Y<br />

><br />

R<br />

R<br />

T<br />

¦<br />

T<br />

1 ; )<br />

ó¦<br />

j<br />

¦<br />

<br />

–<br />

`<br />

4<br />

¥<br />

º<br />

2.2. CONSERVATION OF MOMENTUM 29<br />

where we used the skew-symmetric property of pq¦/ j (A.24).<br />

definition of vorticity vector (1.12), we obtain:<br />

Finally, using the<br />

(2.55)<br />

¦<br />

¦$<br />

which means that the vorticity (n = . "<br />

is conserved ). In particular, this<br />

¦G<br />

means<br />

that if the flow (n<br />

was irrotational ), it will remain so 4 . In this case the problem<br />

of inviscid flow can be solved using velocity potential function (Sec.2.2.4).<br />

" n<br />

The momentum flux<br />

Since )åj<br />

for incompressible flow (2.4), we can rewrite (2.51) as:<br />

<br />

j<br />

1 ; )<br />

1 ; )<br />

4<br />

j<br />

¦ 4<br />

¦<br />

)Wj<br />

And introducing the momentum flux as<br />

¢ ß ¦<br />

j躇><br />

)åj<br />

we have the Euler equation in momentum-flux formulation:<br />

(2.57)<br />

1 ; )<br />

¦ 4 `<br />

<br />

j<br />

2.2.8 Boundary conditions<br />

Equation system (2.2), (2.21), (2.19) <strong>and</strong> (1.13) may not have a unique solution<br />

for any boundary conditions. Generally, the character of the equation system,<br />

i.e. hyperbolic, parabolic or eliptic [4, 5], may change depending on the boundary<br />

conditions <strong>and</strong> the region of space <strong>and</strong> time. However, there are several types<br />

of boundaries that are typically considered, <strong>and</strong> that usually lead to well posed<br />

problems.<br />

4 See Problem 2.7.4


j<br />

<br />

)<br />

F<br />

<br />

)<br />

T<br />

F<br />

r<br />

30 CHAPTER 2. FUNDAMENTAL LAWS<br />

Inlet<br />

At the inlet boundary the value of the velocity is usually specified . This boundary<br />

condition is known as Dirichlet boundary condition. Note, that this is not always<br />

the case, since a pressure can be prescribed as the inlet condition instead, when<br />

the Poisson equation for pressure (2.60) is used.<br />

Outlet<br />

Depending on the character of the equation system the boundary conditions may<br />

or may not need be specified at the outlet boundary. The most common outlet<br />

boundary condition is the condition of the zero boundary-normal velocity derivative<br />

(Neuman boundary).<br />

In more complex flow situations there may not be a clear distinction between<br />

the inlet <strong>and</strong> the outlet, since the flow may reverse. These types of situations are<br />

hard to solve in a consistent manner <strong>and</strong> should be avoided by repositioning the<br />

inlet/outlet of the domain so as to comply to either Dirichlet or Neuman boundary<br />

conditions.<br />

<strong>Fluid</strong>-solid interface (Wall)<br />

In the case of a fluid-solid boundary the flow velocity is set equal to the velocity of<br />

the wall, which covers the cases of both stationary <strong>and</strong> moving boundaries. This<br />

boundary condition is called a no-slip boundary condition.<br />

In some cases a finite velocity jump may be imposed at the boundary, in<br />

which case this is called a slip boundary condition.<br />

The specification of the velocity alone at the boundary may not be enough,<br />

since the momentum equation (2.21) contains second order velocity derivatives<br />

in the viscous (ã 6 ¦<br />

j term ), which means that the first order derivatives should be<br />

given at the boundary. However, with the no-slip condition at the wall, the velocity<br />

derivatives at the wall can be considered zero. This follows from the continuity<br />

relationship (2.4). For example, if we consider velocity )<br />

components ) <strong>and</strong> as<br />

being parallel to the wall, <strong>and</strong> ) normal to the wall, then from the no-slip condition<br />

we have:<br />

r<br />

<strong>and</strong> consequently:


(2.58) X<br />

)<br />

<br />

F<br />

<br />

F<br />

`<br />

<br />

)<br />

F<br />

)<br />

<br />

F<br />

1 @ £ö<br />

A<br />

`<br />

<br />

)<br />

><br />

T<br />

A<br />

@<br />

¯<br />

4<br />

T<br />

2.2. CONSERVATION OF MOMENTUM 31<br />

rq r“<br />

Thus from continuity (2.4) we must have:<br />

rq r“<br />

<br />

<br />

) <br />

In cases when the forces on the wall need to be estimated they can be<br />

related to the boundary normal forces due to pressures, <strong>and</strong> shear forces that<br />

can be related to the stress tensor via (2.14).<br />

It should be noted that if the boundary is moving with acceleration additional<br />

non-inertial terms should be introduced into the boundary conditions (Sec.2.5.2).<br />

<strong>Fluid</strong>-fluid interface (Free surface)<br />

The gas-liquid or liquid-liquid boundary conditions are also called free-boundary<br />

conditions. They consist of the requirements that the pressure, velocities <strong>and</strong><br />

the fluxes of mass <strong>and</strong> momentum be continuous functions across the interface.<br />

This means that these quantities should have the same values on both sides of<br />

the interface. The position of the interface surface will then be determined as a<br />

solution to the flow equations subjected to the free surface boundary conditions.<br />

In cases where surface tension effects are important, they should enter into<br />

the pressure boundary condition, namely the extra boundary pressure should be<br />

added on both sides of the interface. This pressure should be inversely proportional<br />

to the local surface curvature. Since the surface curvature generally<br />

depends on the direction selected on the surface to measure the curvature, one<br />

form of its estimate may be to set it proportional to the sum of inverse curvature<br />

radii in two orthogonal directions:<br />

ÝUô<br />

xyõ<br />

where ö is the coefficient of surface tension. Note, that the forces resulting from<br />

the pressure terms should always act normal to the surface. Shear forces at the<br />

boundary are usually considered to be zero.


¦ Y ¾<br />

Y<br />

(2.60) X<br />

"<br />

><br />

Y<br />

Y<br />

"<br />

¦ì¦G ` Y <br />

1<br />

Y<br />

><br />

1<br />

X<br />

><br />

X<br />

`<br />

><br />

¦<br />

32 CHAPTER 2. FUNDAMENTAL LAWS<br />

2.3 Pressure Equation<br />

2.3.1 General formulation<br />

In the solution of the equation system (2.2), (2.21), (2.19), <strong>and</strong> (1.13), is complicated<br />

by the fact that the equation of mass conservation (2.2) does not contain<br />

pressure, <strong>and</strong> the equation of momentum conservation (2.21) contains both velocity<br />

<strong>and</strong> pressure [6]. In the case of compressible flow the equation of state<br />

(1.13) can be used as an additional relation between pressure <strong>and</strong> density. However,<br />

for the incompressible flow the continuity equation (2.4) has velocity only,<br />

<strong>and</strong> can not be effectively used in combination with the momentum equation.<br />

To make the equation system better conditioned, the continuity equation (2.4) is<br />

usually replaced by the Poisson equation for pressure. To obtain the<br />

¦÷¢<br />

equation<br />

for pressure, let’s apply the divergence operator, , (A.2.32) to the<br />

momentum equation (2.22):<br />

¾<br />

Y : Y <br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

6 ¦<br />

j ã<br />

<br />

j<br />

4 ¾ ¦ 1 ¦<br />

4 `C¾ ¦<br />

¦<br />

¾ ¦ 1 ; ¦ 4<br />

¾ ¦<br />

)Wj<br />

(2.59)<br />

¦ 4 ¦<br />

4 ¦ `<br />

¦ì¦<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

> ; ¦f ¦<br />

6 ¦<br />

j >äã<br />

¦<br />

j<br />

)Wj<br />

After substituting the first term on the LHS from (2.2), considering that J¦ is<br />

a constant, <strong>and</strong> rearranging terms, we have:<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

6 ¦<br />

j ã<br />

<br />

j<br />

¦ 4 ¦ ` 1<br />

4 ¦<br />

"54<br />

)Wj<br />

which is a Poisson equation for pressure. It has to be solved together with the<br />

momentum equation (2.21) <strong>and</strong> the relation of the state law (1.13).<br />

2.3.2 Constant density flow<br />

As it was pointed out the pressure equation is mainly used for incompressible<br />

flows where it replaces the continuity equation (2.4). In this case we can simplify<br />

the pressure equation (2.60) by applying the continuity ) condition to (2.60):<br />

¦K ¦


(2.61) X<br />

(2.62) X<br />

<br />

T<br />

X<br />

)<br />

)<br />

¦<br />

; `<br />

)<br />

` ; <br />

)<br />

j<br />

T<br />

`<br />

)<br />

j<br />

<br />

j<br />

j<br />

j<br />

¦<br />

j<br />

¦ <br />

j<br />

1<br />

2.3. PRESSURE EQUATION 33<br />

¦ 4 <br />

4 ¦<br />

jÙ)Wj<br />

¦ì¦$ ` 1q1 ; )<br />

ã 6 ¦<br />

` ; 1<br />

¦f ¦ 4 <br />

¦f<br />

¦<br />

jÙ)Wj<br />

j )Wj<br />

>Úã<br />

6 ¦<br />

¦K<br />

¦<br />

jÙ)Wj<br />

>Úã<br />

Using the incompressible form of the viscous stress tensor ã 6<br />

, (2.20), <strong>and</strong> the<br />

continuity relation, )åj<br />

<br />

j<br />

, it can be shown that the last term will be zero:<br />

6 ¦<br />

6 ¦<br />

j ã<br />

¦G l 8¬. ¦<br />

j j<br />

<br />

j<br />

¦K<br />

j*j<br />

¦ 4 ¦f<br />

j*jc><br />

<br />

j<br />

<strong>and</strong> finally we obtain:<br />

¦W 8 1<br />

¦<br />

¦ 4 8 151<br />

4 ¦Õ¦ 4 T<br />

>])åj<br />

)Wj<br />

¦<br />

¦ì¦G ` ; )<br />

¦f<br />

j)Wj<br />

This is the incompressible form of the pressure equation, also called the Poisson<br />

equation for pressure. It should be considered together with the momentum<br />

equation (2.24).<br />

2.3.3 Viscous limit<br />

Considering the viscous limit (2.49), <strong>and</strong> taking divergence of this equation, we<br />

have:<br />

¦ì¦$<br />

which is the Laplace equation for pressure.<br />

2.3.4 Boundary conditions<br />

Pressure equation is an elliptic second order PDE. As such it requires the specification<br />

of two sets of boundary conditions, which usually are the values of the<br />

pressure <strong>and</strong> it’s boundary normal derivatives.<br />

At the solid walls the boundary-normal derivative of pressure is usually set<br />

to zero, which is a Neuman boundary condition. The value of pressure at the wall


"<br />

R<br />

R<br />

ˆ<br />

R<br />

R<br />

Ö R <br />

R<br />

<br />

<br />

"<br />

R<br />

R<br />

R<br />

R<br />

"<br />

¦<br />

†<br />

¦<br />

)<br />

><br />

)<br />

¦<br />

)<br />

l<br />

¦<br />

<br />

¦<br />

<br />

; R<br />

"Qù R<br />

R<br />

R<br />

!<br />

!<br />

34 CHAPTER 2. FUNDAMENTAL LAWS<br />

comes out as the solution which satisfies this condition. However, because the<br />

Poisson equation is a second order equation, the Neuman boundary condition<br />

alone will result in an indeterminate solution, when adding any constant to the<br />

pressure will still satisfy the equation. Fixing the value for the pressure in at least<br />

one point will remove this uncertainty. Thus, other conditions at the inlet/outlet<br />

boundaries are usually applied. At the inlet/outlet pressure values are given <strong>and</strong><br />

the boundary normal derivatives are usually set to zero.<br />

In the case when the velocity is also specified at the inlet/outlet, both pressure<br />

<strong>and</strong> velocity specifications should be consistent so as not to create a overdefined<br />

problem. Specifying either pressure or velocity alone will be enough in<br />

many cases. However, this will depend on the character of the flow <strong>and</strong> the discretization<br />

scheme used to solve the equations [4].<br />

2.4 Energy Equation<br />

2.4.1 General formulation<br />

Usually the flow field is a carrier for the transport of other variables of the continuum<br />

media. One important variable is energy.<br />

The balance of energy in a control volume can be written as<br />

(2.63)<br />

" „<br />

" …‡><br />

where the LHS is the rate of energy change inside the control volume, <strong>and</strong> the<br />

two terms on the RHS represent total heat inflow into the control volume <strong>and</strong> work<br />

done on it. This relation is also known as the second law of thermodynamics.<br />

Note that both heat <strong>and</strong> work are transported into the control volume through its<br />

boundary, so they can be represented by flux-vectors.<br />

Energy:<br />

¦ `


Š<br />

¦<br />

R<br />

R<br />

¦<br />

<br />

Š<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

"<br />

s<br />

s<br />

<br />

¦<br />

R<br />

R<br />

!<br />

s<br />

!<br />

2.4. ENERGY EQUATION 35<br />

where the minus sign in front of the gravity term means the potential energy increases<br />

as we move against the gravity force.<br />

Heat: The total heat change inside the volume can be related to the heat flux<br />

through the boundary of the volume:<br />

(2.65)<br />

" …<br />

` Ö ¦<br />

R Š<br />

£Ž¦<br />

where is the heat flux, which is the rate of heat inflow through a unit area per<br />

unit of time, <strong>and</strong> <br />

£‰¦<br />

the vector-element of the boundary introduced in (2.1). Minus<br />

R<br />

sign occurs because of the convention of surface normal vectors to point outside<br />

of the volume, meaning Š that is actually the ”out-flux” of heat … whereas was<br />

defined as an incoming heat by the virtue of (2.63). Applying the Gauss theorem<br />

to (2.65) we have:<br />

(2.66)<br />

` Ö ¦<br />

R Š<br />

` Ö £‰¦$ ¦K ¦<br />

R Š<br />

" …<br />

It is postulated that the heat flux is proportional to temperature gradient:<br />

¦$ `t°ú¾ ¦<br />

`C°<br />

¦<br />

where the minus sign signifies that the heat flows from higher to lower temperatures.<br />

This relation is known as the Fourier’s law. Substituting it into (2.66), we<br />

have:<br />

(2.67)<br />

Ö×1 °<br />

¦ 4 ¦<br />

" …<br />

Work: In analogy to (2.65) we introduce the flux of external work through the<br />

fluid element:<br />

(2.68)<br />

£‰¦<br />

¥Ö †<br />

‹<br />

The work flux vector ‹<br />

<br />

through the area can be computed as


š<br />

¦<br />

R<br />

R<br />

"<br />

)<br />

<br />

‹<br />

<br />

R<br />

R<br />

"<br />

š<br />

)<br />

¥<br />

¢<br />

<br />

‹<br />

R ž<br />

R<br />

¦/m ¦G 6 y¦f ¦G R 6<br />

R<br />

)<br />

— R <br />

R<br />

)<br />

1òR <br />

R<br />

"<br />

"<br />

)<br />

"<br />

. # £<br />

XW›<br />

)<br />

1 ; )<br />

)<br />

<br />

<br />

R<br />

!<br />

)<br />

)<br />

"<br />

)<br />

!<br />

¦<br />

36 CHAPTER 2. FUNDAMENTAL LAWS<br />

–ˆ<br />

" R<br />

–ˆ<br />

ˆ¡=<br />

We saw in (2.2.1) that the surface forces are described by the stress tensor<br />

¦/<br />

. The time derivative of the displacement of fluid element is given by velocity )<br />

6<br />

(1.1). Thus 5 :<br />

(2.69)<br />

¦G<br />

Applying Gauss theorem to (2.68) <strong>and</strong> combining it with (2.69) we obtain:<br />

6 ¦/<br />

(2.70)<br />

† Ö1 6 ¦/<br />

4 ¦<br />

Differentiating by parts, we have:<br />

(2.71)<br />

1 6 ¦/<br />

4 ¦G<br />

6 ¦/y ¦ 6 ¦/<br />

><br />

We can eliminate the derivative of : 6 ¦7m ¦<br />

replaced by the gravity force ` ; J¦<br />

:<br />

6<br />

m ¦<br />

, by expressing it from (2.15) with<br />

; è<br />

where we used the symmetry of 6 : 6 ¦/k<br />

. Now we can rewrite (2.71) as:<br />

4a`<br />

6 m¦<br />

1 6 ¦/<br />

4 ¦G<br />

4a` ; ¡ 4<br />

m ¦<br />

Substituting the above into (2.70), we have:<br />

1 ; )<br />

> 6 ¦/<br />

(2.72)<br />

1 ; )<br />

† Ö ø<br />

m ¦<br />

4a`<br />

; è<br />

> 6 ¦/<br />

ù R<br />

Combining (2.64), (2.67) <strong>and</strong> (2.72), we have<br />

5 Note that the positive signs in (2.68) <strong>and</strong> (2.69) follow the convention that the work done on the system<br />

is positive when the direction of external force coincides with the direction of displacement.


"<br />

1 ; ˆ<br />

4a` ¦ ; )<br />

(2.74) – v<br />

; R ’<br />

(2.76) "<br />

R<br />

"<br />

"<br />

Ö<br />

¦<br />

1 ; ˆ“>ûX<br />

X üR<br />

" ><br />

R<br />

s<br />

"<br />

"<br />

"<br />

R<br />

1 ; ˆ<br />

¦<br />

ˆ<br />

"<br />

1 ; ’<br />

s<br />

)<br />

R <br />

R<br />

¦<br />

)<br />

l<br />

<br />

"<br />

s<br />

¦<br />

1 ; )<br />

; R<br />

"<br />

R<br />

s<br />

)<br />

s<br />

R<br />

)<br />

s<br />

)<br />

¦<br />

><br />

><br />

)<br />

)<br />

¦<br />

)<br />

l<br />

¦<br />

)<br />

l<br />

)<br />

¦<br />

)<br />

l<br />

¦<br />

¦<br />

)<br />

"<br />

)<br />

; R<br />

"ù R<br />

R<br />

; R<br />

"<br />

R<br />

¦<br />

1 ; )<br />

; R<br />

"<br />

R<br />

><br />

!<br />

!<br />

)<br />

¦<br />

)<br />

l<br />

)<br />

¦<br />

–<br />

; R<br />

"<br />

R<br />

¥<br />

)<br />

2.4. ENERGY EQUATION 37<br />

øR<br />

R<br />

1 ; 1<br />

` ; )<br />

" ><br />

4a`<br />

; ¡<br />

y ¦<br />

¥Ö ø 1 °<br />

¦ 4 ¦<br />

> 6 ¦/<br />

>\)<br />

ù R<br />

which after rearranging terms <strong>and</strong> dropping the integration sign due to the arbitrariness<br />

of our choice of the control volume becomes:<br />

¦,R )<br />

¦ 4 ¦<br />

¦ 4a`<br />

¦ ; ¦<br />

m ¦<br />

R<br />

R<br />

¦$R<br />

R<br />

1 °<br />

> ; )<br />

" ><br />

And after canceling the same terms on both sides, we have:<br />

> 6 ¦/<br />

>])<br />

(2.73)<br />

R<br />

R<br />

> 6 ¦/<br />

m ¦<br />

4 1 °<br />

¦ 4 ¦<br />

which is the equation for the rate of change of energy density valid for both compressible<br />

<strong>and</strong> incompressible fluid.<br />

In an ideal gas approximation we can express the LHS in terms of temperature<br />

using the thermodynamic relation (1.23):<br />

y ¦<br />

R<br />

R<br />

4 1 °<br />

¦ 4 ¦<br />

It is useful to express (2.73) in terms of enthalpy (1.20). For this purpose we<br />

to both sides of (2.73) <strong>and</strong> obtain:<br />

can add R X : R<br />

1 ; s<br />

> 6 ¦/<br />

(2.75)<br />

R<br />

R<br />

R 4<br />

R<br />

ïR X 4<br />

" ><br />

R<br />

> 6 ¦/<br />

m ¦<br />

1 °<br />

¦ 4 ¦<br />

which is the equation for the rate of change of enthalpy valid for both compressible<br />

<strong>and</strong> incompressible fluid.<br />

2.4.2 Constant density flow<br />

Let’s substitute 6 ¦<br />

:<br />

j from (2.17) into (2.73), <strong>and</strong> consider that ; <br />

= . "<br />

1 °<br />

¦ 4 ¦<br />

m ¦ 1 `tß ¦/<br />

l 8¬. ¦/ `]ß ¦/á<br />

4<br />

<br />

j<br />

>])<br />

XH><br />

)Wj


R ’ ; "<br />

R<br />

; R ’<br />

(2.77) "<br />

R<br />

X üR<br />

"<br />

R<br />

(2.78) ’><br />

; –<br />

P<br />

(2.79)<br />

s R<br />

"<br />

R<br />

<br />

<br />

X R<br />

" ><br />

R<br />

`<br />

X R<br />

" ><br />

R<br />

; –<br />

(2.80)<br />

P<br />

¦<br />

)<br />

°<br />

¦<br />

s<br />

s R<br />

"<br />

R<br />

><br />

s<br />

–<br />

<br />

s<br />

¥<br />

s<br />

–<br />

¥<br />

)<br />

38 CHAPTER 2. FUNDAMENTAL LAWS<br />

Using the definition of viscous stress (2.19), we can rewrite this as:<br />

ã 6 ¦/<br />

which after applying the continuity relation (2.4) this reduces to:<br />

¦f ¦<br />

1 °<br />

¦ 4 ¦<br />

m ¦<br />

X©)<br />

>^)<br />

ã 6 ¦/<br />

1 °<br />

¦ 4 ¦<br />

m ¦<br />

>^)<br />

where we should use the incompressible form of viscous stress (2.20):<br />

¦f<br />

m ¦ 4<br />

Equation (2.77) can be used in combination with the momentum equation<br />

(2.21) to derive a strong form of the Bernoulli’s equation 6 :<br />

6 ¦7k l 8¬. ¦/ 8 1<br />

)<br />

ã<br />

>^)<br />

@<br />

) l<br />

= . "<br />

> +<br />

We can also rewrite (2.77) in terms of temperature change, <strong>and</strong> velocity. Using<br />

the definition of specific heat (1.25), substituting the velocity from the relation<br />

, we have:<br />

of viscous stress (2.20), <strong>and</strong> assuming ° <br />

= . "<br />

¦ì¦<br />

m ¦ 1<br />

¦K<br />

y ¦ 4<br />

This is an incompressible heat conduction equation.<br />

> 8 )<br />

>])<br />

Heat dominated flow<br />

In the cases with small pressure <strong>and</strong> velocity gradients, or when fluid viscosity<br />

is small compared to the heat conductivity, which corresponds to small Pr<strong>and</strong>tl<br />

number (1.30), equation (2.79) simplifies to:<br />

°<br />

¦ì¦<br />

6 See Problem 2.7.5


; –<br />

(2.81)<br />

P<br />

¦<br />

°<br />

s<br />

=<br />

¦<br />

s<br />

Š<br />

s<br />

2.5. CURVILINEAR COORDINATES 39<br />

where we also presumed that the heat conduction coefficient, ° , is a constant.<br />

This is a limit case of (2.79) for the case of heat dominated flow. Note that since<br />

the substantial derivative is used for T, we still have the convective terms present:<br />

¦ 4 °<br />

¦ì¦<br />

1<br />

s]>])<br />

which is also called a heat convection equation.<br />

2.4.3 Boundary conditions<br />

Generally temperature may experience a jump at the boundary [2]. Usually this<br />

jump is small <strong>and</strong> the temperature of the fluid at the wall is considered to be equal<br />

to the temperature of the wall.<br />

Since the equation (2.81) is a second order differential equation, we would<br />

need to specify temperature derivatives in addition to specifying temperature values<br />

at the boundary. These conditions can be obtained from the consideration<br />

of energy conservation. In particular, heat flux across the boundary should be<br />

conserved. Thus from (2.81) we obtain<br />

(2.82)<br />

¦<br />

¦ò<br />

is a boundary-normal unit vector, <strong>and</strong> Š is the heat flux across the bound-<br />

= where<br />

ary.<br />

2.5 Curvilinear Coordinates<br />

Physical laws should not depend on the choice of a coordinate system. This is<br />

expressed in the terminology of tensor calculus as coordinate invariance. Tensors<br />

are designed to be invariant under coordinate transformations (Remark A.3.3).<br />

Therefore, tensor relations provide a consistent way of writing physical laws.<br />

There are two aspects of expressing physical laws in tensor forms: identifying<br />

, physical components, <strong>and</strong> forming invariant expressions.


Y<br />

"<br />

Y<br />

R<br />

R<br />

0<br />

"<br />

) ¦<br />

6<br />

<br />

0 <br />

¦<br />

¦ )<br />

<br />

j 6 j<br />

<br />

<br />

¦<br />

`<br />

º<br />

40 CHAPTER 2. FUNDAMENTAL LAWS<br />

2.5.1 Invariant forms<br />

The scalar product (Definition A.3.4) was constructed to be invariant. By virtue<br />

of its invariance it represents a physical entity. Using the invariant forms of the<br />

scalar product (Corollary A.3.5), we can rewrite the expression for the substantial<br />

derivative (1.7) in invariant form:<br />

(2.83)<br />

0 ¦<br />

>^)<br />

Correspondingly, the mass conservation equation (2.2) will be expressed as<br />

(2.84)<br />

; <br />

><br />

1 ; )<br />

4 ¦$ T<br />

<strong>and</strong> the momentum equation (2.22) becomes:<br />

(2.85)<br />

1 ; )<br />

j 1 ; )<br />

4 ¦<br />

j<br />

6 j ¦f<br />

j >äã<br />

¦ 4<br />

¦<br />

>])<br />

where the covariant <strong>and</strong> contravariant velocities <strong>and</strong> stress tensors are linked by<br />

the conjugate tensor relations (A.42), (A.43):<br />

¦‡


ý<br />

þ<br />

¡<br />

is<br />

<br />

F<br />

r<br />

><br />

¡<br />

¡<br />

can<br />

2.5. CURVILINEAR COORDINATES 41<br />

non-inertial coordinate systems. This treatment is used in relativistic fluid dynamics<br />

[7]. This approach would also make sense in the treatment of general<br />

moving <strong>and</strong> deforming coordinate systems. However, in a variety of applications<br />

it makes sense to treat space <strong>and</strong> time as separate variables. In this case one<br />

has to distinguish between inertial <strong>and</strong> non-inertial coordinate systems. For any<br />

moving coordinate system one has to formulate an explicit dependence of coordinates<br />

on time. Generally, this time dependence can reveal itself in motions <strong>and</strong><br />

deformations. In this section we shall only consider the case of a non-deforming<br />

<strong>and</strong> moving coordinate system. And in particular, we shall focus attention on an<br />

important case of rotating coordinate systems.<br />

Rotating coordinate systems<br />

Consider an inertial coordinate ý system <strong>and</strong> a non-inertial coordinate system<br />

, which is moving with respect ý to . Let the coordinates of a particular point in<br />

be described by a vector 7 3 Ú% <br />

'q - :<br />

'q<br />

(2.86)<br />

3 ÿ<br />

> ¡<br />

where is the position vector the origin of system with respect to þ ý<br />

ÿ ¡1 ÿ 1 "54<br />

is the coordinate vector of the point in "54 system .<br />

þ <strong>and</strong>¡<br />

Differentiating (2.86) over time, we get<br />

(2.87)<br />

3 ÿ <br />

where the time derivative with respect to the inertial frame of ý reference .<br />

We shall consider the motion of as composed of displacement determined by<br />

<br />

1 þ<br />

ÿ "54<br />

<strong>and</strong> rotation, given by angular n velocity vector . Then be represented<br />

as<br />

<br />

(2.88)<br />

n<br />

Í ¡<br />

><br />

is the velocity of the point as measured in the non-inertial coordi-<br />

£¢ ¡<br />

¤<br />

nate system . Since the rotation does not affect the motion of the origin, , we<br />

ÿ þ<br />

7 Here we are using vector notation, since only the vector quantities are involved<br />

where¢ï¢¥¤§¦


3 <br />

(2.90) <br />

T<br />

<br />

R<br />

<br />

R<br />

R©<br />

R<br />

ÿ<br />

¢ ©<br />

R<br />

R<br />

<br />

"<br />

¡<br />

n<br />

R<br />

R©<br />

<br />

Í ¢<br />

l<br />

R<br />

R<br />

"<br />

Í ¡<br />

><br />

n<br />

<br />

><br />

¢<br />

"<br />

–<br />

<br />

Í ¢<br />

¥<br />

R<br />

R<br />

n<br />

><br />

r<br />

ÿ<br />

r<br />

"<br />

Í<br />

n<br />

¡<br />

n<br />

<br />

–<br />

¥<br />

42 CHAPTER 2. FUNDAMENTAL LAWS<br />

ÿ <br />

¤<br />

have<br />

¢© ¤§¨ <br />

<strong>and</strong> (2.87) becomes<br />

(2.89)<br />

" ><br />

¢<br />

>Ìn<br />

,<br />

3 R <br />

Í ¡<br />

R<br />

" >Ìn<br />

R¡<br />

><br />

Assuming that the rotation is constant:<br />

©<br />

1 "54 <br />

= . "<br />

we can differentiate (2.89) further to obtain<br />

where<br />

©<br />

><br />

n<br />

(2.91)<br />

<strong>and</strong> using 2.88 we have<br />

(2.92)<br />

1¢<br />

><br />

¢ç <br />

Í ¡4<br />

n<br />

Substituting (2.88), (2.91) <strong>and</strong> (2.92) to (2.90) we get<br />

3 R©<br />

Í ¡4<br />

(2.93)<br />

R<br />

R¢<br />

" ><br />

" > n<br />

Í]1¢<br />

>àn<br />

Í]1<br />

Í ¡4<br />

R<br />

" ><br />

R¢<br />

" ><br />

The extra acceleration terms involving n arise due to rotation <strong>and</strong> are interpreted<br />

as being the result of the Coriolis force.<br />

An important special case of a non-inertial coordinate systems is a coordinate<br />

system undergoing a pure rotation with a constant angular velocity. Assume<br />

that the moving coordinate system is undergoing a rotation with n = . "<br />

<strong>and</strong><br />

¤§¨<br />

þ<br />

<br />

. Then we can align the origins of the two coordinate system to ¤<br />

¤ ¤¨


F<br />

r<br />

(2.94) <br />

(2.95) )<br />

<br />

'<br />

T<br />

4<br />

R<br />

R<br />

)<br />

r<br />

"<br />

B<br />

'<br />

<br />

Ç<br />

B<br />

R<br />

"<br />

F<br />

B<br />

2.6. THE LAW OF SIMILARITY 43<br />

eliminate<br />

plane þ<br />

<br />

altogether. Let’s also assume that the axis of rotation is normal to the<br />

, that n is . With these assumptions (2.93) becomes<br />

1UT<br />

'q<br />

'mn<br />

3 R<br />

>à><br />

are the additional acceleration vectors (Problem 2.7.7). The<br />

corresponding Coriolis forces are introduced into the equations of motion in a rotating<br />

coordinate<br />

r¡<br />

Where<strong>and</strong><br />

Ü<br />

w÷><br />

system:<br />

with being the displaced mass. In computations of continuum media dynamics<br />

is replaced with mass R element :<br />

à> <br />

; ¡<br />

where is the density of the fluid <strong>and</strong> is the face-normal velocity across the<br />

face of area of a control volume (see Problem 2.7.8).<br />

¡ ;<br />

2.6 The Law of Similarity<br />

The law of similarity [7, 8] enables in some situations to use a single solution to the<br />

equations of fluid motion to represent a whole family of different cases. Consider<br />

an example of a steady flow past a solid body, where the flow velocity upstream<br />

of the body is. Consider also several cases of such flows when the body has<br />

the same shape but different sizes, . Now, if the only fluid property affecting<br />

this process is the kinematic viscosity, Ç , then in all these cases the distribution<br />

9<br />

, <strong>and</strong> of at least<br />

-<br />

of velocity should be a function of space coordinates Ú% <br />

three additional parameters, ( Ç 3 9<br />

): '“'<br />

IKI/<br />

¦$<br />

¦ 1¨3<br />

'“' 9 4<br />

The number of parameters 8 can be reduced by considering the dimensions<br />

of physical units in which they are measured:<br />

8 A parameter can be looked at as just another independent variable, like space coordinate or time.<br />

However, we treat them separately, since parameters are specific for each physical law, whereasare<br />

not.


±<br />

"<br />

’<br />

ã <br />

(2.96)<br />

(2.97) &1<br />

(2.98) ± ý<br />

"<br />

’<br />

" ²<br />

’<br />

ý<br />

Ç<br />

<br />

A!<br />

¢<br />

¦<br />

F<br />

š<br />

1<br />

ˆ<br />

ˆ<br />

T<br />

" ²<br />

r<br />

’<br />

F<br />

ˆ<br />

44 CHAPTER 2. FUNDAMENTAL LAWS<br />

² é F<br />

The only non-dimensional combination of these parameters is provided by the<br />

Reynolds number:<br />

±/›¨ˆ¡=<br />

±/›¨ˆ¡=<br />

±/›Uˆ¡=<br />

Ç ²$<br />

²<br />

±²$<br />

² é F<br />

±<br />

" #<br />

9 ²G<br />

±<br />

" #<br />

±<br />

9<br />

We can non-dimensionalize other variables by scaling them with the appropriate<br />

length <strong>and</strong> velocity scales:<br />

Ç<br />

¦$¢ <br />

¦G¢ )<br />

Since units dimensions should be preserved in an expression of a physical<br />

law, a law formulated in dimensionless variables can only contain dimensionless<br />

parameters. Hence the new dimensionless variables (2.96) should enter into a<br />

only,<br />

<br />

relation with since it’s the only dimensionless parameter derived from<br />

the properties of the system. Following the convention that the velocity is the<br />

dependent variable (2.95), we can write this relation as:<br />

A!<br />

) ã<br />

¦G<br />

Thus, using simple considerations of physical dimensions, we reduced the<br />

number of parameters from three, '“' 9<br />

( ) to one, Similar considerations<br />

allow to reduce the number of parameters in a more general case, which is proved<br />

in a so-called PI-theorem.<br />

(A# Ç<br />

) ã<br />

ã<br />

3<br />

',A"<br />

4<br />

2.6.1 PI-Theorem<br />

).<br />

Let’s consider a physical law formulated for a set of u<br />

variables, ý<br />

IKI/ý%$:<br />

Suppose that the law requires that each variable can be expressed in units<br />

of length, time an mass, which we call the primary dimensions:<br />

IOILý%$4<br />

¦f²å<br />

" # ²('*),+±<br />

².-/),+±<br />

£ .§. ² ©),+<br />

±/›Uˆ¡=


²<br />

ã ý<br />

(2.99)<br />

(2.100) ý<br />

(2.101) &1<br />

<br />

ý<br />

F<br />

Y<br />

Y<br />

<br />

Y<br />

Y<br />

Y<br />

F<br />

ý<br />

Ç')5+ã 1<br />

Y<br />

Ç<br />

ý<br />

ý<br />

ý<br />

F<br />

r<br />

ý<br />

á0;2Ç')5+é F<br />

á


60=2ý<br />

V<br />

ý<br />

ý<br />

ý<br />

F<br />

F<br />

<br />

<br />

F<br />

<br />

<br />

<br />

<br />

T<br />

T<br />

T<br />

=<br />

<br />

=<br />

`<br />

`<br />

46 CHAPTER 2. FUNDAMENTAL LAWS<br />

We can multiply the equality (2.102) by Ç <strong>and</strong> get:<br />

(2.103)<br />

Y&1<br />

IOILý%$4<br />

á0;2ý<br />

This is an extra relation imposed in addition to our physical law (2.97) by virtue<br />

of scale invariance or homogeneity of our law with respect to length-scaling [9].<br />

In a similar manner we can arrive at two more relations imposed because of<br />

homogeneity with respect to other two primary dimensions: time <strong>and</strong> mass:<br />

Y ý<br />

(2.104)<br />

Y&1<br />

IOI/ý:$4<br />

Y ý<br />

(2.105)<br />

Y&1<br />

IOI/ý:$4<br />

80=2ý<br />

Thus we have three more relations in addition to our physical law (2.97),<br />

Y ý<br />

which means that the number of variables can be reduced u from u À to .<br />

If we use a more complex law that involves an additional primary dimension,<br />

such as temperature, then we can reduce the number of variables of the problem<br />

by 4. Generally, if we have R primary dimensions <strong>and</strong> = independent variables in<br />

the problem, then the independent variables can be reduced ã R to nondimensional<br />

parameters. These parameters can be different depending on the<br />

choice of scaling factors used in transformation (2.100). Generally, normalization<br />

(2.100) does not have to be done by primary dimensions, but can be used with<br />

respect to any group of variables that do not form a so called PI-group, i.e. their<br />

products of the type (2.98) can not be reduced to a non-dimensional number, no<br />

matter what powers are used [10, 11]. This constitutes the essence of the PItheorem<br />

[11]. It lays a more rigorous foundation for the law of similarity [7, 8],<br />

which means that the same solution can be reused by rescaling the variables.<br />

2.6.2 Non-dimensional formulations<br />

To formulate a physical law in dimensionless variables we should introduce dimensional<br />

scales for each variable. The scale, representing the variable, will be<br />

denoted with the same symbol, but with the subscript 0. Scales can be introduced<br />

for both scalar, vector, <strong>and</strong> general tensor variables. Thus, the scale for a vector<br />

variable will be denoted as 10 . In some situations there can be different scales<br />

¦<br />

10 Subscript 0 shouldn’t be confused with the vector component, since vector components are numbered<br />

with one


"<br />

"<br />

V<br />

<br />

<br />

V<br />

V<br />

<br />

<br />

¦<br />

<br />

<br />

V<br />

¾ <br />

<br />

X<br />

X<br />

¦ "<br />

'<br />

"<br />

; V<br />

V<br />

<br />

)<br />

)<br />

¦<br />

"<br />

V<br />

)<br />

`<br />

¦<br />

)<br />

V<br />

2.6. THE LAW OF SIMILARITY 47<br />

for different components, in which case we shall use a different notation.<br />

Space <strong>and</strong> time derivatives should also be scaled. Thus, if we consider<br />

space <strong>and</strong> time as independent variables: , <strong>and</strong> as dependent variables velocity,<br />

density, <strong>and</strong> pressure we can introduce the following non-dimensional variables:<br />

(2.106)<br />

¦$<br />

¦G<br />

ã " <br />

V ã ¾<br />

ã <br />

V<br />

ã ; ;<br />

;<br />

V<br />

<br />

ã ; <br />

V<br />

<br />

ã )<br />

V<br />

¦G )<br />

where we use the Nabla operator to denote the space derivative.<br />

After the dimensional variables are replaced withe the dimensionless ones<br />

by means of (2.106), one should look into the physics of the problem <strong>and</strong> see if<br />

some extra relations between the scales can be applied. For example, in some<br />

problems the characteristic velocity scale can be related to length <strong>and</strong> time scales<br />

) : "<br />

as: . This can be the case in the problem of a steady flow around a fixed<br />

object. However, a steady flow around a rotating object will have an independent<br />

time scale related to the period of rotation.<br />

After all possible eliminations of scales were done, one should try to construct<br />

dimensionless combinations of scales, or non-dimensional parameters.<br />

There can be several different ways in which these parameters can be selected.<br />

This process can be formalized somewhat [10], but there is still a room for subjective<br />

judgment on which dimensionless combinations of scales are most appropriate<br />

as parameters for the problem at h<strong>and</strong>. No matter how these parameters are<br />

selected the PI-theorem states that their minimum number can be as low as =<br />

where = is the number of dimensional scales <strong>and</strong> R is the number of primary dimensions<br />

of the problem. If all the dimensionless parameters have been correctly<br />

identified, it should be possible to replace all the dimensional scales with these<br />

parameters, thereby rendering the physical law in a dimensionless formulation<br />

with the minimum set of independent parameters.<br />

Let’s consider several cases of non-dimensional formulations <strong>and</strong> of application<br />

of the PI-theorem.<br />

R ,<br />

X ã<br />

;<br />

ã<br />

Mass conservation law<br />

Let’s write a non-dimensional formulation of the mass conservation law (2.2):<br />

(2.107)<br />

; <br />

><br />

¾ ¦ 1 ; )<br />

¦ 4 T


X<br />

(2.108) )åj<br />

V<br />

V<br />

V<br />

r<br />

V<br />

)<br />

r<br />

V<br />

)<br />

<br />

V<br />

V<br />

V<br />

r<br />

¾<br />

9 <br />

)<br />

V<br />

<br />

<br />

V<br />

rV<br />

V<br />

j<br />

¾<br />

<br />

<br />

V<br />

V<br />

<br />

l<br />

X<br />

ã<br />

;<br />

V<br />

><br />

V<br />

V<br />

<br />

V V r<br />

)<br />

V<br />

V<br />

ã<br />

¦<br />

V<br />

V<br />

V<br />

48 CHAPTER 2. FUNDAMENTAL LAWS<br />

where we used the nabla operator (A.32) to simplify further analysis. There are<br />

three primary dimensions in this (R À case ): [length], [time] <strong>and</strong> [mass]. Using<br />

the scaling transformations (2.106), the non-dimensional form of (2.107) is:<br />

; > ã ¾ ¦ 1<br />

ã ; ã ) ã<br />

¦ 4 T<br />

As can be seen, all the dimensional parameters canceled out from the equation.<br />

Momentum equation<br />

Consider the steady-state limit of the incompressible momentum equation given<br />

by the Navier-Stokes equation (2.24):<br />

¾ ¦<br />

¦G<br />

¦ `<br />

<br />

Since this is a constant density formulation, density becomes a<br />

<br />

parameter<br />

of the<br />

;<br />

problem: . Considering this, <strong>and</strong> transforming to the non-dimensional<br />

variables according to (2.106) we obtain:<br />

;<br />

(2.109)<br />

ã ¾ ¦<br />

ã ¾<br />

¾ ¦<br />

9 )<br />

¦ ` X<br />

X ã<br />

) ã<br />

)Wj jÃã<br />

) jã<br />

j ã ¾<br />

To simplify things, let’s select for the pressure X scale the dynamic<br />

<br />

pressure:<br />

. By doing this we state that pressure scale, X , is not an independent<br />

;<br />

parameter of our problem, but is related to the density <strong>and</strong> velocity scales. Now,<br />

let’s make each term of (2.109) dimensionless, by multiplying the whole equation<br />

by :<br />

: )<br />

) ã<br />

)Wj jÃã<br />

) júã<br />

X> ã<br />

ã ¾ ¦<br />

ã ¾<br />

j ã ¾<br />

¦ ` ã ¾ ¦<br />

ã<br />


ã )<br />

(2.112)<br />

¤<br />

ã s<br />

(2.113)<br />

ô<br />

(2.114) u<br />

°<br />

<br />

@<br />

¤<br />

¾<br />

ô<br />

x<br />

=<br />

<br />

°<br />

¦<br />

<br />

9 V<br />

V<br />

<br />

V<br />

r<br />

r<br />

Ç<br />

V<br />

¦<br />

¤<br />

@<br />

¦<br />

x<br />

¦<br />

ã<br />

`<br />

R<br />

2.6. THE LAW OF SIMILARITY 49<br />

the number of independent parameters can be as low = as<br />

introduce two non-dimensional numbers: Reynolds number:<br />

Ñ»` l l<br />

. If we<br />

(2.110) A!<br />

)<br />

<strong>and</strong> Froude number, relating the forces of inertia to gravity:<br />

¢<br />

(2.111)<br />

¢ )<br />

V<br />

then we obtain the non-dimensional form of the momentum equation:<br />

)Wj jÃã<br />

) júã<br />

X> ã<br />

ã ¾ ¦<br />

A!<br />

ã<br />

j ã ¾<br />

¦ ` ã ¾ ¦<br />

with the four non-dimensional ã ' ã<br />

variables:<br />

parameters: x (see also Problem 2.7.9).<br />

A!<br />

'<br />

ºâ' ã<br />

ã ) <strong>and</strong><br />

, <strong>and</strong> two non-dimensional<br />

Boundary conditions<br />

Some non-dimensional parameters arise from the boundary conditions. For example,<br />

non-dimensionalizing the boundary condition of the energy equation (2.82)<br />

leads to:<br />

¦<br />

¦$<br />

u )<br />

where u<br />

is the Nusselt number:<br />

¢ …<br />

[^s<br />

²“ ´<br />

(±L… : 1<br />

…<br />

. 4 [^s<br />

where is the wall heat flux ), the characteristic length-scale,<br />

the heat conduction coefficient, (1.29), <strong>and</strong> the characteristic temperature<br />

difference between the wall <strong>and</strong> the fluid.


†<br />

Ü<br />

Ç<br />

)<br />

ö V V<br />

T<br />

¦<br />

r<br />

¦<br />

–<br />

¥<br />

50 CHAPTER 2. FUNDAMENTAL LAWS<br />

Boundary conditions at the free surface give rise to additional parameters,<br />

such as Froude number, relating inertia forces to gravity, (2.111), Weber number,<br />

relating inertia to surface tension:<br />

(2.115)<br />

¢ ;<br />

where<br />

(2.58).<br />

ö<br />

is the coefficient of surface tension entering the boundary condition<br />

Other non-dimensional parameters may appear as new phenomena are<br />

added into the physical law [2].<br />

2.7 Problems<br />

Problem 2.7.1 Derivation of the vorticity equation<br />

Obtain the result outlined in (2.30).<br />

Problem 2.7.2 2D vorticity limit<br />

Perform the missing steps in (2.42).<br />

Problem 2.7.3 Incompressible viscous limit<br />

Derive (2.50) from (2.49).<br />

Problem 2.7.4 Conservation of circulation<br />

The velocity circulation is defined as<br />

(2.116) ><br />

@?<br />

)<br />

R <br />

where the integration is over any closed loop inside the fluid.<br />

Show that for irrotational flow (n<br />

¦G<br />

.<br />

Problem 2.7.5 Bernoulli’s equation<br />

Using the energy equation (2.77):<br />

):><br />

<br />

= . "


¦<br />

Y<br />

"<br />

Y<br />

in (2.94) in terms of n©'q®<br />

F<br />

r<br />

R ’ ;<br />

"<br />

R<br />

<br />

R<br />

R<br />

!<br />

"<br />

F<br />

'<br />

X R<br />

" ><br />

R<br />

R<br />

R<br />

ÿ<br />

"<br />

¦<br />

)<br />

<br />

r<br />

¦<br />

¦<br />

Ä<br />

<br />

Ö<br />

n<br />

s<br />

<br />

)<br />

`<br />

<br />

¦<br />

=<br />

X<br />

–<br />

¦ £<br />

R<br />

'<br />

T<br />

¥<br />

¦<br />

£<br />

'<br />

'<br />

£<br />

'BAF<br />

'BAr<br />

'CA<br />

2.7. PROBLEMS 51<br />

ã 6 ¦/<br />

1 °<br />

¦ 4 ¦<br />

m ¦<br />

<strong>and</strong> momentum equation (2.21):<br />

>])<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

6 ¦<br />

j >äã<br />

<br />

j<br />

¦ 4<br />

¦<br />

derive the strong formulation of the Bernoulli’s equation:<br />

>^)Wj<br />

>\š<br />

’><br />

> +<br />

@<br />

) l<br />

= . "<br />

<strong>and</strong> formulate it’s applicability limits.<br />

Problem 2.7.6 Volume change inside a moving boundary<br />

Suppose that a region of space is enclosed by a moving boundary. The<br />

velocity of motion of the ) boundary, , is given at each point on the boundary.<br />

Show that the rate of change of the volume, , of that region will be equal to:<br />

!<br />

(2.117)<br />

= where is the unit normal vector to the boundary R <strong>and</strong><br />

<strong>and</strong> find the Ä coefficient .<br />

surface area element,<br />

Problem 2.7.7 Rotating coordinates<br />

£Žr<br />

Obtain explicit relations for the components of acceleration vectors in<br />

F<br />

£ . '<br />

'*®<br />

'*® '<br />

Problem 2.7.8 Rotation with separated coordinate origins<br />

Consider a simple rotation with n<br />

T<br />

˜%<br />

n<br />

ý<br />

origin of the rotating coordinate system rotate with the same around the<br />

origin of :<br />

'mnŒ- as in (2.94), but now let the<br />

Í ÿ<br />

Derive the expression for<br />

3<br />

in this case.


V<br />

V<br />

)<br />

– ;<br />

P<br />

s R<br />

"<br />

R<br />

X üR<br />

" ><br />

R<br />

(2.118) „©z<br />

r<br />

V<br />

°<br />

s<br />

–<br />

P<br />

V<br />

s<br />

r<br />

)<br />

52 CHAPTER 2. FUNDAMENTAL LAWS<br />

Problem 2.7.9 Nondimesionalizing energy equation<br />

Write a non-dimensional form of the heat convection equation (2.79):<br />

¦ì¦<br />

m ¦ 1<br />

¦K<br />

y ¦ 4<br />

;<br />

selecting for the pressure scale. Determine the minimum number of<br />

X<br />

dimensionless parameters. Write the equation using the Eckert number (3.114)<br />

as one of the parameters:<br />

> 8 )<br />

>])<br />

¢ )<br />

V


where º<br />

¦<br />

)<br />

(3.4) º<br />

; é F ã 6 ¦<br />

j<br />

(3.5)<br />

P<br />

¦<br />

s<br />

<br />

<br />

j<br />

s<br />

<br />

j<br />

<br />

)<br />

¦<br />

T<br />

Chapter 3<br />

Laminar flows<br />

3.1 Assumptions<br />

Flow equations discussed in Chapter 2 provide analytical solutions only in some<br />

special cases. In this chapter we shall consider the equations for incompressible<br />

flow: (2.4), (2.24) <strong>and</strong> (2.74), assuming that all the coefficients are constant:<br />

(3.1)<br />

(3.2)<br />

(3.3)<br />

; –<br />

>])WjÙ)<br />

1<br />

s]>])<br />

¦f<br />

j<br />

F ã 6 ¦ ;Ãé<br />

¦ 4 °<br />

¦G ¦f<br />

;Ãé F<br />

`<br />

º<br />

¦<br />

¦ì¦<br />

>])<br />

y ¦ 6 ¦/<br />

is the Hydrostatic pressure:<br />

; <br />

For Newtonian incompressible fluids the viscous stress ã 6 ¦<br />

j term, , has the form<br />

(2.23):<br />

<br />

j<br />

9 )<br />

¦f<br />

j*j<br />

Definition 3.1.1 Laminar flow<br />

Let’s make an assumption of laminar flow which states that the time scale<br />

of changes in the flow can not be lower than the time-scale of the motion of the<br />

53


F<br />

r<br />

<br />

)<br />

54 CHAPTER 3. LAMINAR FLOWS<br />

boundary or any external sources. In other words, if there is any repeatability in<br />

the motion of the boundary or in the external forces then the frequencies associated<br />

with either factors can not be lower than the frequencies of the flow motion.<br />

It means that neither the boundaries not external forces can induce any additional<br />

frequencies in the flow. In the limit case of non-moving boundaries <strong>and</strong><br />

non-changing forces the flow should not depend on time, which means that all<br />

the dependent variables should become functions of spatial coordinates only.<br />

The conditions of laminar flow defined by (3.1.1) are realized when the contribution<br />

of the non-linear term in the momentum equation (3.2) is small, or<br />

¦f<br />

when<br />

dominates. This is usually the case<br />

j*j<br />

when non-dimensional Reynolds number (2.110):<br />

the contribution of the viscous term 9 )<br />

(3.6) A!<br />

<br />

9<br />

is small. In practical situations the ”smallness” of<br />

Ç<br />

A#<br />

to A!<br />

corresponds<br />

Navier-Stokes equation, (3.2) is known to have very few analytical<br />

¦f<br />

solutions.<br />

This is mainly due to the non-linear )<br />

convective term , which is the main<br />

cause for the rich dynamical features of fluid flow. For this reason, most of the<br />

cases that provide analytical solution do not include the convective term. Below<br />

we shall consider several such cases.<br />

ED@<br />

T‰T<br />

.<br />

3.2 Confined flows<br />

Probably the simplest of confined flows are the flows between moving surfaces,<br />

which belong to the category of Couette flows [2].<br />

3.2.1 Flow between parallel plates<br />

Let’s consider a flow between two parallel plates, one of which is moving relative<br />

to the other with a constant velocity¦<br />

(Fig. 3.1).<br />

We are looking for a two-dimensional solution, since by the assumption of<br />

laminar flow (3.1.1) <strong>and</strong> from the symmetry of the problem we do not expect<br />

any changes in the transverse direction. We are also looking for a steady-state<br />

solution, thus all the variables will be the functions of axial <strong>and</strong> vertical coordinates<br />

- only:<br />

, <strong>and</strong> only two velocity components need to be considered<br />

3 %<br />

'q


)<br />

)<br />

F<br />

r<br />

F<br />

(3.9) )<br />

r<br />

s<br />

r<br />

> 8 —íR ) r<br />

®í R<br />

6 8 R )<br />

(3.10)<br />

® R<br />

(3.11) Äkõ<br />

¢<br />

s<br />

V<br />

1<br />

®<br />

6 l<br />

;r<br />

<br />

r<br />

)<br />

r<br />

l 8<br />

<br />

T<br />

T<br />

s<br />

F<br />

l<br />

F<br />

3.2. CONFINED FLOWS 55<br />

Figure 3.1: Flow between parallel plates: the lower plate is at rest, the upper plate is<br />

moving with velocity.<br />

¦§%<br />

- . Since the plates are considered to be infinite no variable should<br />

'*)<br />

change in direction either. Thus, the only independent variable of the problem<br />

becomes the vertical <br />

coordinate , which we shall denote ® as . Likewise, from<br />

the symmetry of the problem the only non-zero component of velocity ) is , which<br />

we shall denote ) by . In addition to this we can also assume the pressure to be<br />

constant. This can be explained by the absence of normal stresses in this flow.<br />

With these assumptions the momentum <strong>and</strong> energy equations (3.2), (3.3) reduce<br />

to:<br />

(3.7)<br />

8 R<br />

r <br />

(3.8)<br />

R ®<br />

° R<br />

Equations (3.7) <strong>and</strong> (3.8) can be<br />

1UT<br />

solved<br />

4<br />

with<br />

<br />

the<br />

1U<br />

boundary<br />

4<br />

conditions<br />

<br />

u(0)<br />

s<br />

= s<br />

0 <strong>and</strong> u(H) = U, <strong>and</strong> <strong>and</strong> , with the solution (See Problem<br />

3.7.2):<br />

R ®<br />

4 <br />

®<br />

<strong>and</strong> the shear stress:<br />

The non-dimensional friction coefficient, ÄŒõ , becomes inversely proportional<br />

to the Reynolds number:<br />

8<br />

A!<br />

;


(3.12) º­{<br />

'<br />

T<br />

s<br />

1<br />

®<br />

¤<br />

®<br />

Y<br />

¢<br />

Y<br />

x<br />

r<br />

<br />

¢<br />

Y —W¢<br />

¢<br />

Y<br />

)GH<br />

Ä<br />

F<br />

s<br />

F<br />

F<br />

`<br />

`<br />

<br />

s<br />

s<br />

<br />

V<br />

l<br />

4<br />

T<br />

<br />

<br />

56 CHAPTER 3. LAMINAR FLOWS<br />

<strong>and</strong> the Poiseuille number:<br />

Solution to the temperature equation (3.8) produces a quadratic dependence<br />

on ® (See Problem 3.7.2):<br />

— s<br />

ÄkõEA!<br />

<br />

` 8 4<br />

°<br />

8<br />

° l<br />

l ><br />

V ><br />

®»>às<br />

V<br />

The dimensionless Brinkman number is introduced as a relative measure of viscous<br />

forces to thermal fluxes:<br />

(3.13)<br />

8r<br />

° 1<br />

In the momentum equation (3.7) the effect gravity was neglected under the<br />

assumption that the gravity force acts normal to the direction of the flow. Generally<br />

it may not be the case, but the solution procedure remains essentially the same<br />

(see Problem 3.7.3).<br />

3.2.2 Axially moving concentric cylinders<br />

¦GÚ%<br />

In this case only the axial component of ) )WxÙ'*)GFè'q)IHÙ-<br />

% T<br />

velocity vector is<br />

¢<br />

non-zero:<br />

, <strong>and</strong> it only depends on<br />

4 <br />

: )IH1 ¢ 4<br />

. The appearance of any<br />

)IH1<br />

',Î'*+<br />

'*)GHÙother<br />

velocity component, or a dependence on other coordinates will lead to the<br />

violation of the assumption of a laminar flow (Definition 3.1.1). ¢ Substituting this<br />

form of the solution into the momentum equations in cylindrical coordinates, we<br />

find that only the an axial momentum equation takes a non-trivial form:<br />

(3.14)<br />

@<br />

¢<br />

A solution that satisfies this equation is:<br />

(3.15) )GH1<br />

r<br />

4 <br />

dfÆ 1 ¢ 4<br />

¢<br />

>\Ä


¢<br />

F<br />

r<br />

Ä<br />

F<br />

<br />

F<br />

A<br />

4 <br />

V<br />

F<br />

<br />

A<br />

A<br />

`<br />

V<br />

4 <br />

V<br />

V : A<br />

V F<br />

Ä<br />

Ä<br />

V<br />

F<br />

A<br />

A<br />

X R<br />

¢<br />

R<br />

T<br />

F<br />

F<br />

Y<br />

Î Y<br />

'<br />

' )GH1K<br />

<br />

V<br />

)GH1<br />

A<br />

V<br />

4 >JF<br />

T<br />

Fr<br />

T<br />

F<br />

4 <br />

F<br />

T<br />

`<br />

Ä<br />

1 ¢ : dfÆ<br />

A<br />

A<br />

F<br />

F<br />

V<br />

V<br />

A<br />

4<br />

4<br />

- , <strong>and</strong> )GF<br />

V<br />

4<br />

)GF1 ¢<br />

3.2. CONFINED FLOWS 57<br />

',Ä<br />

Substituting this into (3.15) we have:<br />

)GH1<br />

r“<br />

dKÆ 1<br />

<strong>and</strong> the solution becomes:<br />

dfÆ 1<br />

4 ' Ä<br />

(3.16) )IH1<br />

: ¢ 4<br />

dKÆ 1<br />

¢<br />

dKÆ 1<br />

dKÆ 1<br />

Remark 3.2.1 Pulling an infinite rod<br />

Consider the problem above with the boundary conditions:<br />

: A<br />

: A<br />

T<br />

Then, applying this conditions to (3.15), we have<br />

)GH1<br />

4 <br />

V<br />

4<br />

dKÆ 1K<br />

4<br />

r“<br />

rk<br />

from which it Ä<br />

follows that . Thus the problem of pulling an infinite rod<br />

does not have a steady-state solution.<br />

>\Ä<br />

3.2.3 Rotating concentric cylinders<br />

In this case we ) have:<br />

Continuity:<br />

¦$Ú%<br />

are the constants, which can be determined from the boundary con-<br />

Ä where<br />

ditions:<br />

)œx¡'*)GFÙ'*)GHè-<br />

Ú% T<br />

'*)GFè'<br />

4<br />

.<br />

-momentum:<br />

(3.17)<br />

; )<br />

¢


R<br />

¢<br />

R<br />

A<br />

A<br />

R —<br />

R<br />

(3.19) )IF<br />

(3.20) )IF<br />

V<br />

<br />

T<br />

A<br />

n<br />

V<br />

V<br />

A<br />

A<br />

F<br />

V<br />

F<br />

¢<br />

R<br />

R<br />

r<br />

A<br />

A<br />

s<br />

4<br />

<br />

V<br />

n<br />

F<br />

n<br />

V<br />

F<br />

R<br />

R<br />

'<br />

'<br />

: ¢ ` ¢ : A<br />

F : F `<br />

A : A A<br />

V<br />

V<br />

Ä<br />

¢<br />

F<br />

F<br />

n<br />

s<br />

s<br />

><br />

¢<br />

F<br />

1<br />

A<br />

1<br />

A<br />

V<br />

F<br />

r<br />

Ä<br />

¢<br />

n<br />

F<br />

`<br />

F<br />

T<br />

A<br />

¢ Mr<br />

)GF<br />

s<br />

s<br />

V<br />

F<br />

: ¢<br />

A<br />

A :<br />

F<br />

V<br />

V<br />

<br />

`<br />

`<br />

T<br />

A<br />

A<br />

V : A<br />

V<br />

F<br />

F<br />

r<br />

58 CHAPTER 3. LAMINAR FLOWS<br />

Î -momentum (Problem 3.7.5):<br />

(3.18)<br />

)GF<br />

<br />

¢NM ¢EL)IF<br />

¢ r ><br />

1 ¢<br />

°<br />

¢<br />

8 L)IF<br />

x ><br />

Boundary conditions:<br />

)IF1<br />

4 <br />

4 <br />

)IF1<br />

4 <br />

4 <br />

The equation that determines the velocity<br />

satisfy this equation has the form:<br />

)IF1<br />

4 ¢<br />

is (3.18). A solution that will<br />

Substituting it into the boundary conditions, we find the Ä constants<br />

<strong>and</strong> the final solution becomes:<br />

, Ä<br />

,<br />

: ¢<br />

>\A<br />

Remark 3.2.2 Flow inside a rotating cylinder<br />

If we set A<br />

the solution above will become:<br />

)GF<br />

which is a solid body rotation. Thus the steady-state solution for the flow inside a<br />

rotating cylinder is a solid body rotation.<br />

Remark 3.2.3 Rotating an infinite rod<br />

If we solve the problem above with the boundary conditions:


(3.23) X<br />

F<br />

<br />

T<br />

<br />

Ä<br />

X<br />

A<br />

F<br />

X R<br />

¢<br />

R<br />

X<br />

V<br />

V<br />

Ä<br />

><br />

<br />

V<br />

><br />

Ä<br />

A<br />

¢<br />

V<br />

r<br />

r<br />

V<br />

V<br />

Fr<br />

r<br />

' )GFK<br />

A<br />

<br />

Ä <br />

K<br />

<br />

r<br />

r<br />

r<br />

V<br />

r<br />

; n<br />

1<br />

@<br />

r<br />

`<br />

V<br />

T<br />

r<br />

A<br />

r<br />

r<br />

4<br />

F<br />

r<br />

1<br />

A<br />

V<br />

X<br />

V<br />

3.2. CONFINED FLOWS 59<br />

T<br />

we can obtain the following system for the coefficients Ä<br />

:<br />

)GFqA<br />

nÅA<br />

',Ä<br />

<br />

(3.21)<br />

FK<br />

><br />

n?A<br />

from which we have: Ä<br />

, <strong>and</strong> the solution becomes:<br />

r<br />

',Ä<br />

n?A<br />

(3.22) )IF<br />

n?A<br />

¢<br />

Substituting it into (3.17) <strong>and</strong> integrating it, we can obtain the pressure distribution:<br />

 A<br />

V<br />

; )<br />

¢ <br />

<strong>and</strong> the pressure distribution is<br />

1 ¢ 4 ` ; n<br />

 A<br />

V<br />

where the Ä constant is found from the boundary X condition:<br />

the pressure distribution is:<br />

4 <br />

. Thus,<br />

>^Ä<br />

l ¢ r<br />

@ ; n l<br />

It is useful to compute the rotational momentum (torque) that arises in a<br />

system of two rotating cylinders (Problem 3.7.6).<br />

: ¢<br />

3.2.4 Poiseuille flow through ducts<br />

Let’s consider a case of a straight duct with a constant cross-sectional area, .<br />

Since the area does not change its form, the length-scale of the problem will be<br />

the characteristic size of the duct 1 : ¡ .<br />

½<br />

1 See Sec.3.2.6 for another measure of duct diameter<br />

¼¡<br />

F}É


(3.24) )<br />

<br />

)<br />

F<br />

¢<br />

)<br />

F<br />

<br />

F<br />

)<br />

1<br />

<br />

<br />

r<br />

º<br />

1<br />

<br />

4<br />

F<br />

4<br />

º<br />

`<br />

<br />

F<br />

º<br />

F<br />

<br />

<br />

F<br />

Y º <br />

Y<br />

)<br />

F<br />

)<br />

1<br />

<br />

F<br />

F<br />

1<br />

<br />

)<br />

r<br />

r<br />

R º <br />

R<br />

4<br />

F<br />

4<br />

'<br />

T<br />

'<br />

4<br />

T<br />

º<br />

<br />

)<br />

T<br />

F<br />

)<br />

r<br />

<br />

F<br />

F<br />

r<br />

4<br />

60 CHAPTER 3. LAMINAR FLOWS<br />

Compared to the case of infinite plates, a duct has an entrance <strong>and</strong> it has a<br />

finite cross-sectional area. The existence of an entrance causes entrance effects,<br />

such that the flow is three dimensional over some distance from the<br />

¦íÌ%<br />

entrance,<br />

that is all three components of velocity ) '*) - are non-zero: , <strong>and</strong> each<br />

component also depends on all three )<br />

spatial coordinates: .<br />

However, we assume that this transition region will end at some distance<br />

after the entrance <strong>and</strong> be replaced by a fully developed flow region, where axial<br />

velocity does no longer depend on the axial coordinate, i.e.<br />

¦G<br />

¦ 1 '*)<br />

'5<br />

'q <br />

'5 <br />

It should be noted that the existence of a fully developed laminar regime<br />

is only an assumption, but it happens so, that there is a solution satisfying this<br />

assumption. However, it also happens that there are other solutions, which do not<br />

satisfy this assumption, i.e. unsteady turbulent regime. Which solution is realized<br />

in reality depends on the magnitude of the Reynolds number. At the low Reynolds<br />

numbers the fully developed regime occurs in circular ducts at distances between<br />

30 <strong>and</strong> 100 duct diameters from the entrance.<br />

In a fully developed flow in addition to (3.24), the velocity components<br />

¦t%<br />

normal<br />

to the axis should ) 'q - be zero: . This is because the appearance<br />

of any velocity component normal to the axis can not be sustained for a<br />

long period of time since there is no pressure gradient imposed in that direction.<br />

On the other h<strong>and</strong>, any short time appearance of such components will violate<br />

the assumption of a steady-state laminar nature of the flow in a fully developed<br />

region. Thus, we can use only the axial component, which we shall denote as<br />

. With these assumptions, in a fully developed flow the axial momentum<br />

equation (3.2) can be simplified to:<br />

(3.25)<br />

T <br />

rðr<br />

> 8 1<br />

<br />

ð<br />

The other two momentum equations simplify º to<br />

only on º : , <strong>and</strong> we can write:<br />

r»<br />

<br />

<br />

. Thus, º depends<br />

>^)<br />

Since )<br />

'q <br />

the second term on the RHS of (3.25) does not depend on<br />

, <strong>and</strong> then it follows from (3.25) º that should not depend on either. But<br />

<br />

F


ã<br />

F<br />

<br />

¦<br />

F<br />

r<br />

<br />

ã<br />

@<br />

¢<br />

ã<br />

R<br />

¢<br />

ã R<br />

ã<br />

) ã<br />

—<br />

<br />

ã<br />

'<br />

@<br />

T<br />

)<br />

¢<br />

<br />

1<br />

@ ) ã<br />

<br />

@<br />

½<br />

F<br />

r<br />

F<br />

<br />

)<br />

3.2. CONFINED FLOWS 61<br />

since º<br />

depends only on is a constant. Thus, dividing<br />

, <strong>and</strong> introducing dimensionless variables:<br />

equation (3.25) by R º : R <br />

, it follows that R º : R <br />

¾ ¦G ½Å¾ ¦<br />

¦$<br />

8 )<br />

where ½<br />

is the appropriate measure of the duct cross-sectional size, we obtain a<br />

the following boundary value problem for the dimensionless ã ) velocity :<br />

½ ' ã<br />

) ã<br />

R º : R <br />

(3.26)<br />

¾ ¦ ã<br />

¾ ¦<br />

)POQ<br />

where ‹ subscript st<strong>and</strong>s for ”wall” <strong>and</strong> index spans only the variables that )<br />

#<br />

. This is a Poisson equation in a confined 2D domain with the<br />

'q<br />

depends on:<br />

no-slip velocity at the boundary (Dirichlet boundary conditions).<br />

Remark 3.2.4 Reynolds number independence<br />

Note that the Reynolds number does not enter the equation (3.26), <strong>and</strong><br />

therefore, should not affect the solution. This is because we excluded non-steady<br />

solutions <strong>and</strong> turbulence from our consideration.<br />

The circular pipe<br />

Consider a fully developed flow region in a circular pipe of radius A . The natural<br />

coordinate system for this case is cylindrical: ¢ ',Î',+ , with + being the axial coordinate.<br />

Following the discussion of the previous section, only the axial velocity<br />

ã<br />

component, )IHwill be non zero, which we shall denote for simplicity ) as . As it<br />

was shown, it can only depend on ¢ <strong>and</strong> Î . However, because of the symmetry of<br />

the duct, Î the dependence on will inevitably lead to<br />

1<br />

time-dependent<br />

¢<br />

solution <strong>and</strong><br />

violate the assumption of fully developed flow.<br />

4<br />

)<br />

Thus, we should have ,<br />

<strong>and</strong> using the Laplacian operator in cylindrical coordinates, (3.26) reduces to:<br />

(3.27)<br />

ã ¾<br />

¢ R ã<br />

) ã<br />

R ã<br />

4 T<br />

which gives the parabolic solution:


(3.28) )<br />

(3.29) …<br />

8 —ÛR ) `<br />

¢<br />

R<br />

(3.31) ÄŒõ<br />

(3.32) º­{<br />

<br />

<br />

<br />

xZYV<br />

l 6Q<br />

º R<br />

R¢<br />

R<br />

r<br />

¢<br />

º R<br />

R<br />

)<br />

r<br />

á<br />

Ñ<br />

r<br />

)<br />

`<br />

r<br />

¢<br />

<br />

<br />

R<br />

`<br />

A<br />

@<br />

)œ|­w<br />

<br />

l<br />

º R<br />

R<br />

UAÂ `<br />

W8<br />

<br />

º R<br />

R<br />

62 CHAPTER 3. LAMINAR FLOWS<br />

¢ 4 @ 1<br />

ÑSRã<br />

<strong>and</strong> in dimensional units:<br />

) ã<br />

@T<br />

¢ 4 @ 1<br />

8 Ñ<br />

This solution is called the Poiseuille flow.<br />

The Volumetric flow rate through the pipe can be computed as<br />

rT<br />

Ö Ö<br />

¡¼<br />

¢»<br />

The mean duct velocity is defined as:<br />

V<br />

l§UÖJV<br />

) R<br />

` A<br />

… ¢<br />

¡<br />

)<br />

X<br />

The wall shear stress:<br />

W8<br />

(3.30)<br />

6Q<br />

Ñ<br />

)<br />

A<br />

8X<br />

A `<br />

l<br />

<br />

Darcy friction factor:<br />

<br />

áÅ<br />

Skin-friction coefficient<br />

;X W6Q<br />

r <br />

)<br />

where the Reynolds number is based on duct diameter:<br />

;X A! @*[<br />

A\<br />

Poiseuille number<br />

)<br />

X<br />

½ :‰9<br />

.<br />

<br />

@[<br />

Äkõ§A!


(3.33)<br />

8 R<br />

<br />

<br />

ã )<br />

(3.35)<br />

V<br />

<br />

ã )<br />

(3.36)<br />

T<br />

F<br />

<br />

@<br />

`<br />

Ä<br />

r<br />

<br />

<br />

R<br />

r<br />

) ã<br />

®<br />

Ä<br />

r<br />

1<br />

r<br />

1<br />

@<br />

r<br />

)<br />

R º r<br />

R<br />

r<br />

F<br />

T<br />

@<br />

º<br />

R<br />

`<br />

Ä<br />

V<br />

<br />

Ä<br />

r<br />

<br />

º<br />

1<br />

<br />

4<br />

3.2. CONFINED FLOWS 63<br />

3.2.5 Combined Couette-Poiseuille flows<br />

Couette flows are driven by shear boundary motion, <strong>and</strong> Poiseuille flows - by<br />

the axial pressure gradient. The equations describing both types of flows do not<br />

have a nonlinear convective terms, which makes these equations linear, <strong>and</strong> thus<br />

enables linear superpositions of solutions.<br />

Consider the Couette flow between parallel plates (Sec.3.2.1) but with a<br />

constant pressure gradient applied in the axial direction. The momentum equation<br />

in this case will be a combination of (3.7) <strong>and</strong> (3.25) in the form:<br />

R ®<br />

where we are using for the axial ® <strong>and</strong> for the vertical coordinates. Since the<br />

LHS does not depend on <strong>and</strong> the pressure is a function of º only: , we<br />

conclude R º : R that must be a constant. Introducing dimensionless coordinates:<br />

)<br />

:, we have:<br />

® ã<br />

: '^Ä<br />

®»>^Ä ã<br />

Applying the boundary conditions:<br />

1UT 4 <br />

) ã<br />

4 <br />

we Ä have:<br />

form:<br />

'*Ä<br />

, <strong>and</strong> after renaming Ä<br />

, the solution has the<br />

) ã<br />

4<br />

® ã<br />

®»>¼@ Ä¥ã


(3.37) Ä<br />

®<br />

<br />

T<br />

½<br />

(3.40)<br />

X<br />

X R<br />

R<br />

½<br />

<br />

~<br />

<br />

<br />

@<br />

l 8<br />

¢<br />

r<br />

º<br />

R<br />

›<br />

64 CHAPTER 3. LAMINAR FLOWS<br />

We can find the constant Ä after substituting this solution into (3.34):<br />

From (3.36) we can see, that Ä @ when the velocity changes sign at the lower<br />

(ã wall ). This is called flow separation point, <strong>and</strong> according to (3.37) it corresponds<br />

to the pressure gradient:<br />

R<br />

(3.38)<br />

l<br />

r<br />

8<br />

A pressure gradient greater than this will cause the flow at the lower wall to<br />

reverse 2 .<br />

3.2.6 Non-circular ducts<br />

Because the problem of the fully developed duct flow was reduced to a well posed<br />

<strong>and</strong> well studied boundary value problem based on the Poisson equation there<br />

are variety of analytical solutions obtained for ducts of various shapes [2].<br />

There are several convenient measures that are introduced for ducts of arbitrary<br />

shapes.<br />

The cross-sectional length-scale of the duct is called the Hydraulic diameter,<br />

which is introduced as a generalization of relation for a diameter of a circle.<br />

For a circle of diameter the relation between it’s area <strong>and</strong> a perimeter]<br />

¡ Ñ ¡ ½ <br />

is:<br />

. Thus, for any non-circular duct of º perimeter <strong>and</strong> area the<br />

¡ :]<br />

hydraulic diameter is defined as:<br />

(3.39)<br />

Ñ ¡<br />

The mean wall shear stress is defined as:<br />

]<br />

6Q<br />

<br />

] ? 6QR<br />

2 See Problem 3.7.7


)<br />

1<br />

<br />

r<br />

? 6QR<br />

<br />

)<br />

¢<br />

› R <br />

X<br />

4<br />

<br />

`<br />

¡]<br />

R<br />

º<br />

R<br />

Ö ;Ãé F º<br />

<br />

F<br />

<br />

R<br />

F<br />

"<br />

3.3. UNSTEADY FLOWS 65<br />

where the integration is done over the perimeter of the duct.<br />

In a fully developed flow each element of the fluid between cross-sectional<br />

planes at <strong>and</strong> ><br />

that element should be zero. This means that the friction force at the wall should<br />

exactly balance the axial force pushing the fluid element due to the pressure drop<br />

in the duct. Thus, we have the following relation between the mean wall shear<br />

stress <strong>and</strong> the pressure gradient:<br />

R moves with a constant velocity. Thus the sum of forces on<br />

` ¡<br />

R º<br />

<strong>and</strong> using the definition of the wall mean shear stress (3.40), we have:<br />

3.3 Unsteady flows<br />

6Q<br />

<br />

Some unsteady flows in the ducts can still be solved analytically. To introduce<br />

unsteadiness we formally use the same assumptions that lead to (3.26), but now<br />

we shall reinstall the time derivative from (3.2), which with these assumptions<br />

becomes:<br />

(3.41)<br />

9 )<br />

¦ì¦ ` ; é F º<br />

where as in ) '5 <br />

Sec.3.2.4, is the axial component of velocity in the duct,<br />

which is the only non-zero velocity component. Following the same reasoning<br />

as in Sec.3.2.4, we conclude that the last term can not depend on any spatial<br />

coordinate. Since we consider unsteady flows, this term can still depend on time.<br />

However, in this problem we can combine velocity <strong>and</strong> pressure into a single joint<br />

variable:<br />

)<br />

^<br />

)â><br />

And the final equation becomes:<br />

(3.42)<br />

9<br />

¦ì¦<br />

)<br />

^<br />

)<br />

^


(3.43) P`_w<br />

<br />

1 4<br />

®<br />

1<br />

Y<br />

Y<br />

)<br />

^<br />

"<br />

<br />

š<br />

1 4<br />

ˆ ®<br />

r<br />

n<br />

r<br />

)<br />

^<br />

<br />

)<br />

^<br />

1<br />

<br />

><br />

r<br />

)<br />

^<br />

<br />

4<br />

66 CHAPTER 3. LAMINAR FLOWS<br />

This equation has a parabolic character, which means that one independent<br />

variable - time - is asymmetric with respect to direction. Specifically, at any point<br />

in time the solution will depend only on the previous points on the time axis, but<br />

not on the subsequent points. This difference in the directions of time: the future<br />

<strong>and</strong> the past, is the result of the first order time derivative in the equation (3.42).<br />

In contrast, the second order Laplacian space derivative,<br />

directions equivalent. It should be noted that since<br />

operator in (3.42) includes only two components:<br />

)<br />

^<br />

¦ì¦G<br />

)<br />

^<br />

rðr<br />

)<br />

^<br />

¦ì¦<br />

makes all space<br />

, the Laplacian<br />

'q<br />

. ð<br />

Equation (3.42) is still simple enough to provide analytical solutions in several<br />

cases. The important cases include:<br />

1. Starting flow in a duct.<br />

2. Pipe flow due to oscillating pressure gradient.<br />

3. <strong>Fluid</strong> oscillating above an infinite plate.<br />

4. Unsteady flow between infinite plates.<br />

3.3.1 <strong>Fluid</strong> oscillation above infinite plate<br />

Suppose that the plate is oscillating in direction<br />

, <strong>and</strong> the velocity of oscillation is:<br />

®<br />

, which we shall denote as<br />

r<br />

<br />

ba/ced1<br />

"54<br />

In this case the solution for the fluid velocity will depends on only one ® direction :<br />

. Then equation (3.42) can be rewritten as:<br />

)<br />

^<br />

)<br />

^<br />

(3.44)<br />

9 Y<br />

)<br />

r^<br />

Y ®<br />

We can note that in this case the motion in time is periodic, while the changes<br />

in space are aperiodic. Correspondingly, we may look for a solution form which<br />

is a periodic function in time <strong>and</strong> a decaying function of space. One form of the<br />

solution that will satisfy this equation is:<br />

(3.45)<br />

¦,f$<br />

"54 <br />

)<br />

^<br />

®å'<br />

>\Ä


where Ë<br />

1 4<br />

®<br />

1<br />

<br />

T<br />

r<br />

1 4<br />

®<br />

1<br />

'<br />

1<br />

1<br />

1<br />

#<br />

š<br />

1<br />

®<br />

n<br />

9<br />

ˆ énp¯<br />

<br />

n<br />

ˆ<br />

r<br />

#<br />

¥<br />

n<br />

¥<br />

n<br />

n<br />

r<br />

á<br />

®<br />

n<br />

4<br />

r<br />

4<br />

4 :nm<br />

3.3. UNSTEADY FLOWS 67<br />

where š<br />

is an unknown function ® of . If we substitute the latter into (3.44), we’ll<br />

get the equation š for :<br />

(3.46)<br />

n<br />

9<br />

¢×á<br />

šg¢<br />

where<br />

has a solution:<br />

š<br />

šGgg<br />

R š : R ® , <strong>and</strong> the parameter á was introduced for brevity. This equation<br />

4 ¼¡<br />

where ¡<br />

is a constant. Thus, the solution for<br />

) , (3.45), is:<br />

^<br />

ˆ é'¯<br />

"54 ¡ih`jlk1<br />

#<br />

"­`<br />

Using the definition of á , (3.46), <strong>and</strong> the identity: # F}É<br />

, we can write:<br />

)<br />

^<br />

®å'<br />

>\Ä<br />

1 #<br />

l<br />

>¼@<br />

áÅ —<br />

F}É<br />

1 #<br />

n<br />

l 9<br />

4o<br />

<strong>and</strong> for<br />

we obtain:<br />

>@<br />

"54<br />

®å'<br />

)<br />

^<br />

¦0f$<br />

)<br />

^<br />

®å'<br />

"54 £¡<br />

>\Ä<br />

. Considering only the real solution, we have:<br />

énp¯2énp¯<br />

1<br />

F}É<br />

n : l 9 4<br />

Ë®<br />

)<br />

^<br />

®å'<br />

"54 ¡<br />

>\Ä<br />

Ä ¡ Constant <strong>and</strong> can be found from the initial conditions. In case of a moving<br />

plate <strong>and</strong> a stagnant flow-field at infinity (3.43), we have:<br />

. 1<br />

"a`<br />

–<br />

Thus ¡£<br />

, Ä<br />

, <strong>and</strong> the final solution is:<br />

)<br />

^<br />

>\Ä<br />

1}T<br />

"54 <br />

qarced1<br />

"54 ¼¡<br />

a/ced1<br />

"54<br />

(3.47)<br />

"54 <br />

. 1<br />

"a`<br />

4<br />

)<br />

^<br />

®å'<br />

Cˆ énp¯<br />

–<br />

Ë®


1<br />

¥<br />

'<br />

® ts¬<br />

"<br />

V<br />

<br />

<br />

n<br />

r<br />

ˆ énp¯<br />

Y 9r<br />

Y<br />

Y<br />

Y<br />

)<br />

<br />

)<br />

¥<br />

<br />

n<br />

)<br />

¥<br />

n<br />

n<br />

68 CHAPTER 3. LAMINAR FLOWS<br />

In case of a stationary plate <strong>and</strong> a flow-field oscillating asÒ 1 "54 <br />

we can obtain the solution by subtracting the equation above fromÒ<br />

(t):<br />

ba/ced1<br />

"54<br />

,<br />

(3.48)<br />

)<br />

^<br />

®å'<br />

"54 <br />

H±/–<br />

. 1<br />

"54a`<br />

–<br />

. 1<br />

"a`<br />

Ë®<br />

4 ²<br />

It can be checked by a direct substitution that the above equation satisfies (3.44)<br />

<strong>and</strong> the boundary conditions: , <strong>and</strong><br />

.<br />

1}T<br />

"54 T<br />

"54 <br />

"54<br />

)<br />

^<br />

)<br />

^<br />

1K<br />

'<br />

t–<br />

. 1<br />

3.3.2 Unsteady flow between infinite plates<br />

Consider two parallel plates separated by a distance , <strong>and</strong> a fluid with viscosity<br />

initially at rest is filling up the space between the plates. Suppose that the<br />

upper plate starts moving with velocity. We are looking for the solution for the<br />

unsteady flow-field between the plates. This case is similar to the one described<br />

9<br />

above, but now the solution should be aperiodic in time, <strong>and</strong> in fact it can be periodic<br />

in space, since any periodic function with a spatial period equal to the plate<br />

separation, , will be suitable. The equation (3.44) is still valid in this case. However,<br />

now we have an explicit length-scale, , <strong>and</strong> using dimensionless variables<br />

becomes more attractive. Let’s define the non-dimensional variables as:<br />

(3.49)<br />

"<br />

"<br />

)<br />

^<br />

È <br />

ã )<br />

V<br />

Then equation (3.44) can be rewritten in the non-dimensional variables as:<br />

ã ) Y<br />

È r<br />

r^<br />

Ys<br />

<br />

And selecting the time scale as " V<br />

:J9<br />

, we have:<br />

ã ) Y<br />

È r<br />

<br />

This equation should be solved for the unknown function ã<br />

Ys<br />

conditions:<br />

r^<br />

1MÈ<br />

's4<br />

with the boundary


)<br />

)<br />

1<br />

'<br />

) ã<br />

1<br />

) ã<br />

) ã<br />

@<br />

`<br />

'<br />

) ã<br />

1<br />

) ã<br />

) ã<br />

š<br />

šg<br />

'<br />

)<br />

T<br />

@<br />

T<br />

È<br />

š<br />

@<br />

`<br />

@<br />

Y<br />

Y<br />

<br />

r<br />

T<br />

3.3. UNSTEADY FLOWS 69<br />

(3.50)<br />

(3.51)<br />

(3.52)<br />

(3.53)<br />

1¨È<br />

1UT<br />

1MÈ<br />

<br />

's4<br />

<br />

@‰'s4<br />

'K 4<br />

T 4 <br />

<br />

where (3.50) describes the fixed lower plate, (3.51) describes the moving upper<br />

plate, (3.52) is the initial <strong>and</strong> (3.52) the final velocity distributions. Note, that the<br />

final velocity distribution was obtained before as the steady state solution for this<br />

case (Sec.3.2.1). As can be seen, boundary condition (3.51) is non-zero. To<br />

simplify our search for the right solution, it would be nice if we could look for a<br />

function which is zero at the boundaries. To make the boundary conditions both<br />

zero let’s look for a solution that is represented by the difference between the<br />

steady-state ã solution<br />

at both plates:<br />

1MÈ<br />

4<br />

, (3.53), <strong>and</strong> ã<br />

'K<br />

1¨È<br />

's4<br />

, since this function will be zero<br />

) ã<br />

1¨È<br />

Now, if ã 's4<br />

we replace with the new unknown function:<br />

following boundary value problem:<br />

) ã<br />

1MÈ<br />

's4<br />

, we obtain the<br />

1MÈ<br />

È `<br />

1¨È<br />

's4 <br />

's4<br />

(3.54)<br />

Y©<br />

(3.55)<br />

(3.56)<br />

(3.57)<br />

1}T<br />

<br />

's4<br />

<br />

@‰'s4<br />

4 T<br />

T `<br />

1UT<br />

È r<br />

<br />

T ` T T<br />

's4<br />

Ys<br />

<br />

@‰'s4<br />

T 4 È ` T È<br />

1¨È<br />

1¨È<br />

È `<br />

4 4<br />

where we have all the boundary values zero. We can look for a solution in form<br />

of separated variables:<br />

'K 'K<br />

1¨È<br />

<br />

È `<br />

1¨È<br />

È ` È T<br />

) ã<br />

(3.58)<br />

's4 ‡ 1¨È 4<br />

1¨È<br />

1s4<br />

Substituting this into (3.54) we obtain relationship between š <strong>and</strong> :<br />

gg<br />

r<br />

¢ ` á


È<br />

¡<br />

B<br />

1<br />

'<br />

¡<br />

B<br />

<br />

<br />

@<br />

š<br />

4<br />

š<br />

Ö F <br />

F é<br />

4<br />

1<br />

r<br />

š<br />

š<br />

š<br />

Ò<br />

wBYV<br />

<br />

><br />

<br />

¡<br />

4<br />

<br />

¤<br />

a/ced1<br />

1UT<br />

1MÈ<br />

BdefÆ 1<br />

ù<br />

B/z<br />

é<br />

B/z<br />

R<br />

È<br />

@<br />

`<br />

<br />

4<br />

T<br />

@<br />

4<br />

B<br />

70 CHAPTER 3. LAMINAR FLOWS<br />

where as in (3.46) we introduced a constant á . This time, however, á is not known<br />

in advance, <strong>and</strong> should be determined from the boundary conditions. The relation<br />

above is identically satisfied by:<br />

(3.59)<br />

(3.60)<br />

1MÈ 4 ¼¡<br />

defÆ 1 á È 4<br />

1s4 <br />

á<br />

È 4<br />

Ä ˆ é'u<br />

with , Ä <strong>and</strong> unknown constants in addition to . Out of these the constant<br />

can be absorbed into <strong>and</strong> , since ¡ ¤<br />

¤ <strong>and</strong> enter as a product in (3.58). So,<br />

¡ á š Ä<br />

without loss of generality we can set Ä<br />

(3.55) - (3.58) can be translated to 1¨È 4 <strong>and</strong> š<br />

. The boundary conditions on<br />

@<br />

as:<br />

1MÈ 's4<br />

1s4<br />

(3.61)<br />

(3.62)<br />

(3.63)<br />

(3.64)<br />

1¨È<br />

1MÈ<br />

1}T<br />

T ‡ 1UT 4<br />

's4<br />

T ‡ 1<br />

@‰'s4<br />

'K 4<br />

4 È 1¨È 4 T<br />

È ¼ 1MÈ 4 <br />

1s4<br />

1s4 v v<br />

1}T 4 v<br />

1K 4<br />

T 4<br />

4 T 1<br />

4 È<br />

1¨È 4 T <br />

where the last equality (3.64) is satisfied identically by virtue of (3.60). From (3.61)<br />

<strong>and</strong> (3.59) if follows that ¤ç T . Similarly, from (3.62) it follows that á¿ =U, where<br />

is an integer number. The only way to reconcile boundary condition<br />

1MÈ 4 = (3.63):<br />

, with boundary conditions (3.61) <strong>and</strong> (3.62) <strong>and</strong> the analytical form of<br />

given by (3.59), is to express as a Fourier series:<br />

4 1MÈ<br />

1MÈ 4<br />

1¨È 4 <br />

È <br />

=UÈ4<br />

Coefficients ¡ B<br />

can be found as:<br />

ÈdefÆ 1<br />

Computing the integral:x\"defÆ 1<br />

"a/cyd1<br />

, we obtain:<br />

=UÈ4<br />

<br />

deKÆ 1<br />

4í`<br />

R <br />

4í`<br />

"a/ced1<br />

l 1 `<br />

ødeKÆ 1<br />

=U<br />

Thus for<br />

we have:<br />

=U4<br />

1MÈ 4


wBYV<br />

) R<br />

"<br />

R<br />

(3.66) º<br />

º<br />

¦ì¦G 8 )<br />

U<br />

Ò<br />

(3.67) º<br />

p<br />

¦<br />

æ<br />

j<br />

wBYV<br />

@ h`j|k1<br />

=<br />

9 )<br />

4<br />

B<br />

¦G 8 )<br />

<br />

@<br />

`<br />

T<br />

)<br />

4<br />

¦<br />

æ<br />

;<br />

=<br />

=<br />

rUrs4defÆ 1<br />

T<br />

3.4. CREEPING FLOWS 71<br />

` l<br />

1 `<br />

1<br />

=UÈ4<br />

B{deKÆ<br />

1MÈ 4 <br />

And the final function becomes:<br />

(3.65)<br />

U<br />

Ò<br />

1 `<br />

1MÈ<br />

` l<br />

=UÈ4<br />

's4 <br />

3.4 Creeping flows<br />

If we combine equations (3.2) <strong>and</strong> (3.5), <strong>and</strong> use the expression for substantial<br />

derivative (1.7), we obtain yet another form of Navier-Stokes equation (2.24):<br />

¦<br />

¦f} ` º<br />

The LHS of this equation represents the inertial forces. The assumption of creeping<br />

flow or Stokes flow states that the inertial forces are negligible. With this<br />

assumption the last equation becomes:<br />

¦f}<br />

Differentiating over # , we get:<br />

from which we obtain a Laplace equation for pressure:<br />

¦f ¦/}“ 8 1<br />

¦f ¦ 4 }<br />

¦ì¦G<br />

Forming a product with p5¦/ j <strong>and</strong> using the symmetric identity (A.27), we obtain:<br />

T p<br />

¦f }<br />

j<br />

¦<br />

j¡º<br />

jè)


(3.68) n<br />

(3.69) ë<br />

<br />

n<br />

æ<br />

<br />

ë<br />

`<br />

ë<br />

<br />

x<br />

T<br />

T<br />

Î<br />

Î<br />

72 CHAPTER 3. LAMINAR FLOWS<br />

Swapping indexes # <strong>and</strong> ° in this equation <strong>and</strong> subtracting one from another we<br />

can express it in terms of vorticity vector (1.12):<br />

}<br />

Thus both the vorticity <strong>and</strong> pressure satisfy Laplace equation in a creeping flow.<br />

Important cases of creeping flow include:<br />

1. Fully developed duct flow. Re-number independent.<br />

2. Flow about immersed bodies (Stokes solution or the sphere).<br />

3. Flow in narrow but variable passages. (Lubrication theory).<br />

4. Flow through porous media.<br />

3.4.1 Stokes flow around a sphere<br />

Consider a laminar viscous flow around a sphere of radius A , with the velocity at<br />

infinity. The solution to this problem will be two-dimensional, since by symmetry<br />

nothing should depend on the azimuthal direction. It was shown in Sec.2.2.5 that<br />

in a 2D limit the vorticity vector has only one component <strong>and</strong> it is related to the<br />

Laplacian of the stream function (2.46):<br />

¦ì¦<br />

With these assumptions (3.68) becomes:<br />

¦ì¦/}<br />

In spherical coordinates the relation between the velocity <strong>and</strong> stream-function<br />

(2.6) will become:<br />

(3.70)<br />

(3.71)<br />

)œx<br />

`<br />

)GF<br />

¢ rdefÆ<br />

F<br />

¢defÆ<br />

ë


(3.73) ë<br />

(3.76) X<br />

Y<br />

ë<br />

r<br />

<br />

ë<br />

@ r Y<br />

¢<br />

Î Y<br />

<br />

1<br />

A<br />

Î<br />

X<br />

@<br />

rdeKÆ<br />

—<br />

r<br />

ë<br />

rdefÆ<br />

¥<br />

A<br />

r<br />

` À 8 Ò<br />

l ¢ r A}<br />

Î<br />

A<br />

A<br />

<br />

Î<br />

<br />

4<br />

`<br />

`<br />

À<br />

A<br />

¢<br />

r<br />

<br />

><br />

<br />

ë<br />

<br />

A<br />

r<br />

r<br />

T<br />

3.4. CREEPING FLOWS 73<br />

<strong>and</strong> the Laplacian operator in spherical coordinates is:<br />

(3.72)<br />

" 1<br />

—\Y<br />

¢ r ><br />

r ` –<br />

¢ r Y<br />

where we neglected the azimuthal direction angle because of the symmetry of<br />

the problem. The boundary conditions for this problem are:<br />

Y Îò<br />

<br />

x<br />

4 <br />

F1<br />

4 T<br />

<br />

l<br />

¢<br />

ÎŒ>\Ä<br />

Ä where is any constant.<br />

4<br />

The solution satisfying (3.72) is:<br />

1K<br />

l ¢<br />

1 ¢<br />

',Î<br />

@ 4<br />

Ñ£<br />

A —<br />

¢<br />

<br />

which can be checked by direct substitution 3 into (3.72). The velocity components<br />

can be found from (3.70) <strong>and</strong> (3.71) to be:<br />

(3.74)<br />

À‰A<br />

¢ l<br />

)œx<br />

ba/cedåÎ<br />

@“><br />

l ¢ <br />

(3.75)<br />

qdefÆ<br />

À‰A<br />

¢ Ñ<br />

)GF<br />

— `<br />

The important quantity is the fluid drag on the sphere. It consists of the<br />

contribution of the shear stress (tangential friction at the surface), <strong>and</strong> the pressure<br />

forces normal to the surface (Fig.3.2). Pressure distribution on the surface<br />

of the sphere is computed from the momentum equation (3.66), <strong>and</strong> results in the<br />

following expression:<br />

@“><br />

Ñ ¢ ><br />

arcedWÎ<br />

3 See Problem 3.7.8


½ ¡<br />

=<br />

1~<br />

(3.77)<br />

6 x„F<br />

¤<br />

=<br />

<br />

¤<br />

¤<br />

õ<br />

¤<br />

Î<br />

T<br />

4<br />

õ<br />

<br />

¤<br />

=<br />

<br />

<br />

Ö<br />

x<br />

`<br />

¡<br />

=<br />

><br />

<br />

Î<br />

<br />

4<br />

Ö<br />

R<br />

¡<br />

><br />

¡<br />

Î<br />

1~<br />

Î<br />

4<br />

R<br />

¡<br />

¤<br />

=<br />

<br />

4 <br />

a/ced1<br />

Î<br />

4<br />

74 CHAPTER 3. LAMINAR FLOWS<br />

žI~<br />

=^)GF<br />

L)œx<br />

Â<br />

The total force on the sphere can be obtained by integrating both shear (surface<br />

friction) <strong>and</strong> the pressure components over the surface of the sphere:<br />

F<br />

¢ M )GF<br />

l ¢<br />

(3.78)<br />

~<br />

¤<br />

=


¤<br />

¤<br />

¤<br />

r<br />

Ö 1<br />

Vz<br />

(3.81) Ĥ<br />

Ö<br />

Ö<br />

z<br />

¤<br />

r<br />

4<br />

<br />

4<br />

<br />

V<br />

z<br />

1<br />

<br />

4<br />

<br />

><br />

l<br />

A"<br />

¤<br />

l<br />

@<br />

a/ced `<br />

r<br />

Ö<br />

Ö `<br />

Vz<br />

À<br />

V<br />

z<br />

<br />

<br />

4<br />

4<br />

6 x„FGdeKÆ<br />

OzV<br />

<br />

¡ l§UA<br />

<br />

À<br />

l<br />

Ñ<br />

À<br />

r<br />

l<br />

<br />

4<br />

1<br />

Î<br />

Ò<br />

4<br />

3.4. CREEPING FLOWS 75<br />

<strong>and</strong> substituting the expression for the area R element:<br />

have:<br />

A‹deKÆ<br />

R Î , we<br />

(3.79)<br />

¤xñw‡<br />

X‰defÆ<br />

l§UA<br />

R Î ÎŒa/cydWÎ<br />

Î R Î<br />

where the negative sign in the second term is due to the fact the direction of<br />

increase Î of is opposite to the selected of~<br />

= direction . Substituting x„Ffrom<br />

6<br />

(3.77) <strong>and</strong> X from (3.76), <strong>and</strong> setting ¢ A , we obtain:<br />

—WÖ<br />

ÎŒarcedr<br />

where we omitted the term involving the constant pressure X<br />

component , since<br />

its contribution will become zero after the integration. Computing the integrals:<br />

ÀU8CA<br />

ΠR Ω><br />

Î R Î<br />

¤xðwZ‡<br />

deKÆ <br />

defÆ<br />

a/ced1<br />

À<<br />

Ô{a/cyd1<br />

4a`<br />

R <br />

OzV<br />

V<br />

defÆ 1<br />

4arcedr<br />

we finally obtain:<br />

1<br />

deKÆ<br />

<br />

Vz<br />

R <br />

1<br />

(3.80)<br />

1 Ñ<br />

l 4 <br />

¤xñw‡<br />

U8tA<br />

[U8<br />

where the plus sign signifies that the force is directed toward the flow velocity.<br />

It is interesting to note, that the contribution of the viscous wall friction due to the<br />

A}<br />

shear stress 6 xFis twice as big as that one from the normal pressure term.<br />

A useful engineering formula for a drag coefficient is obtained by relating the<br />

drag net pressure, to flow inertia (kinetic ;r<br />

energy), :<br />

: 1UA<br />

: l<br />

Ĥ<br />

which in terms of non-dimensional Reynolds number, A\<br />

;rUA<br />

A}:J9<br />

becomes:<br />

l Ñ<br />

<br />

which is valid for A#<br />

. Ž@


(3.82)<br />

;t)<br />

(3.83) Ä<br />

which is valid for T Ž<br />

<br />

@<br />

A!<br />

—<br />

`<br />

@•> mA! [<br />

X<br />

À<br />

T<br />

Ñ<br />

I<br />

<br />

76 CHAPTER 3. LAMINAR FLOWS<br />

3.4.2 2D Creeping flows<br />

Let’s consider the limit of 2D creeping flows over plane surfaces. These flows can<br />

be described by equation (3.69). It can be rigorously shown that this equation<br />

can not have a non-zero steady-state solution with the steady-state boundary<br />

conditions at infinity [12]. Intuitively, it is clear that if an infinitely large plate starts<br />

moving with a constant velocity it will tend to impose this velocity on the rest of<br />

the flow-field, but for an infinite flow field it will take infinite time to accomplish.<br />

Therefore, there will be no steady state solution to this problem. This situation<br />

became known as the Stokes paradox. To remove this paradox it was proposed<br />

to add a convective derivative to a momentum equation [12, 2]:<br />

¦G<br />

¦<br />

¦f€<br />

is the known free-stream velocity.<br />

coefficient becomes:<br />

where<br />

With these assumptions the drag<br />

> 8 )<br />

l Ñ<br />

x which is called the Oseen approximation, <strong>and</strong> is valid I for .<br />

In addition to this there are various engineering approximations obtained for<br />

T Ž‡À<br />

the drag coefficients of a sphere <strong>and</strong> a cylinder for various Reynolds numbers [2].<br />

A reasonably good curve-fit approximation for the sphere is:<br />

@“><br />

@[A" (>¼I¡IèI<br />

(3.84) Ä‘<br />

l Ñ<br />

ÏŽ A\<br />

A"<br />

><br />

><br />

<br />

Í<br />

l<br />

3.4.3 Lubrication theory<br />

’Ž A!<br />

T”“<br />

Lubrication theory focuses on the study of 2D creeping flows between the contracting<br />

or exp<strong>and</strong>ing surfaces, when the surfaces are also moving with respect<br />

to each other.


T<br />

<br />

Ç<br />

<br />

T<br />

<br />

Ç<br />

<br />

Ç<br />

<br />

Ç<br />

<br />

T<br />

<br />

T<br />

<br />

T<br />

4<br />

3.4. CREEPING FLOWS 77<br />

Figure 3.3: Flow between contracting plates<br />

Pressure inside a non-uniform gap<br />

Let’s consider a flow between two plates one of which is at an angle to the other<br />

(Fig.3.3). We shall use the coordinates in horizontal, <strong>and</strong> ® in the vertical directions,<br />

<strong>and</strong> denote the corresponding velocity components ) as <strong>and</strong>•. Without<br />

loss of generality we may presume the lower plate ® at to be horizontal <strong>and</strong><br />

the upper plate is given by a known profile of its height at each -position: .<br />

Also without loss of generality we can assume that the lower plate is moving at<br />

a constant velocity. Let’s consider a flow in a limited section stretching from<br />

à1<br />

coordinate to .<br />

From Sec.3.2.1 we know that the steady flow between parallel plates has<br />

a linear profile given by (3.9). In the case of tilted plates we can not expect this<br />

profile to hold, because it would to the loss of mass-flow conservation, i.e. more<br />

flow will enter the section at than exit at . Indeed, since the flow<br />

velocity at the lower moving plate should be always <strong>and</strong> at the upper fixed<br />

plate it should always be zero, there will always be more flow entering at ,<br />

where the plates separation is wider than at where it is narrower. This is<br />

because a linear velocity profile takes shape of a triangle, <strong>and</strong> the mass flow rate<br />

is proportional to the area of this triangle. The area of the triangle at will<br />

always be larger than the one at . From this we conclude that the linear<br />

velocity profile can not be a solution to our problem.<br />

In this situation a pressure distribution arises in the flow that leads to nonlinear<br />

flow profiles at the inlet <strong>and</strong> the outlet, such that the mass conservation is<br />

satisfied. Thus, the nature of this flow will be that of combined Couette-Poiseuille<br />

flow discussed in Sec.3.2.5.<br />

If we assume that the solution is of combined Couette-Poiseuille type, we


1<br />

)<br />

<br />

<br />

>•<br />

¯<br />

ã )<br />

(3.87)<br />

4<br />

)<br />

)<br />

<br />

<br />

)<br />

Ä<br />

l<br />

<br />

F<br />

Ä<br />

<br />

F<br />

r<br />

R<br />

r<br />

) ã<br />

®<br />

Ä<br />

r<br />

1<br />

@<br />

r<br />

F<br />

Ä<br />

¯<br />

•<br />

V<br />

r<br />

<br />

`<br />

r<br />

º<br />

R<br />

F<br />

@<br />

)<br />

<br />

<br />

º<br />

R<br />

@<br />

@<br />

<br />

)<br />

<br />

<br />

T<br />

<br />

)<br />

<br />

<br />

<br />

T<br />

78 CHAPTER 3. LAMINAR FLOWS<br />

can find the equation for the pressure distribution in the section between the converging<br />

plates, that will lead to mass conservation. For this purpose, let’s use the<br />

incompressibility condition (2.4), <strong>and</strong> apply it to our case:<br />

¦f ¦G<br />

rq r“<br />

If we integrate the last equality over the cros-section, we obtain:<br />

>–•<br />

¯<br />

>])<br />

(3.85)<br />

V<br />

Ö˜—<br />

V<br />

ÖJ—<br />

V<br />

Öš—<br />

V<br />

¥Öš—<br />

R ®»>–•1U 4­`<br />

•1}T 4<br />

R ®<br />

R ®»><br />

R ®<br />

4 4 T<br />

where we walls:•1UT<br />

•1}<br />

used the no-slip condition at the . The solution<br />

for the combined Couette-Poiseuille flow should be obtained from equation (3.34):<br />

V<br />

ÖJ—<br />

R ®<br />

(3.86)<br />

r <br />

R ã<br />

<strong>and</strong> boundary conditions:<br />

8<br />

R<br />

1}T 4 <br />

) ã<br />

4 T<br />

) ã<br />

Looking for a solution in the same form as (3.35):<br />

® ã<br />

>\Ä<br />

®ƒ>\Ä ã<br />

V<br />

we can obtain the following values for the constants:<br />

>\Ä<br />

r“<br />

8<br />

R


Ä<br />

(3.88) Ä<br />

ã )<br />

(3.89)<br />

<br />

<br />

Ö<br />

V—<br />

Ç<br />

<br />

)<br />

<br />

<br />

T<br />

ø<br />

ø<br />

Ä<br />

@<br />

À<br />

º<br />

V<br />

<br />

V<br />

) R ã ® ù ã<br />

<br />

<br />

4<br />

<br />

<br />

ø<br />

Ç<br />

º<br />

r<br />

¢<br />

1<br />

<br />

<br />

l 8<br />

<br />

r<br />

Ä<br />

º<br />

R<br />

<br />

<br />

Ö<br />

V<br />

T<br />

4<br />

F<br />

r<br />

4<br />

º<br />

º<br />

4<br />

V<br />

V<br />

<br />

<br />

<br />

ù<br />

<br />

T<br />

<br />

T<br />

3.4. CREEPING FLOWS 79<br />

where the first two equations follow from the boundary conditions, <strong>and</strong> the last<br />

equality follows from the substitution of (3.87) into (3.86). If we define a constant<br />

as:<br />

we finally obtain:<br />

R<br />

` 1<br />

ã ® Ä<br />

@•>^Ä<br />

Since is now a function of , is also a function of , given by (3.88). Transferring<br />

(3.85) to dimensionless variables, <strong>and</strong> substituting (3.89), we obtain:<br />

F<br />

—<br />

` 1<br />

4 Ö F<br />

R ®<br />

® ã<br />

@“>^Ä<br />

®»>¼@ ã<br />

(3.90)<br />

` o1<br />

Ö<br />

@•>^Ä<br />

[<br />

><br />

@ 4<br />

>¼@ l<br />

Substituting Ä<br />

from (3.88), we obtain:<br />

(3.91) R<br />

—<br />

` 1<br />

—<br />

À©>\Ä<br />

<br />

<br />

T<br />

à1<br />

<br />

(<br />

(<br />

º<br />

This is the equation for pressure distribution inside the section between the two<br />

converging plates. It should be solved with a known profile of the gap width<br />

as a function of :<br />

, <strong>and</strong> with the given pressure distribution at the<br />

inlet ) <strong>and</strong> the outlet ) of the section. Constant pressure boundary<br />

conditions can be used in a simplified case: .<br />

ù<br />

<br />

<br />

<br />

[8<br />

1}T 4 <br />

1 Ç 4 <br />

<br />

<br />

Remark 3.4.1 Contracting vs exp<strong>and</strong>ing gap<br />

When we solve equation (3.91) with the constant pressure<br />

4<br />

boundary<br />

4 <br />

conditions:<br />

, for a linear profile of we should obtain a parabolic<br />

ºg1}T<br />

ºg1Ç<br />

solution with an à1 extremal point, or <br />

maximum)<br />

4<br />

between<br />

<br />

<strong>and</strong><br />

. The condition for this º point is . Opening the parentheses on the<br />

LHS of (3.91), we can obtain for the point following relation:<br />

I G I<br />

,(minimum<br />

the


À<br />

r<br />

<br />

<br />

º<br />

<br />

<br />

><br />

<br />

<br />

<br />

)<br />

º<br />

<br />

)<br />

<br />

,<br />

<br />

<br />

T<br />

><br />

)<br />

<br />

,<br />

1•<br />

,<br />

<br />

º<br />

<br />

,<br />

)<br />

<br />

¯m¯<br />

<br />

4<br />

4<br />

<br />

<br />

[8<br />

<br />

Y<br />

80 CHAPTER 3. LAMINAR FLOWS<br />

T<br />

from which we see that if the slope of the upper wall is negative:<br />

T<br />

(contracting<br />

gap), º<br />

,<br />

then , <strong>and</strong> the extremal point is a maximum. This means that<br />

the pressure will be everywhere higher than the ambient inside the contracting<br />

gap. The opposite conclusion follows for the exp<strong>and</strong>ing gap.<br />

Ž Ž<br />

In reality the case of the exp<strong>and</strong>ing gap may lead to higher flow instabilities<br />

<strong>and</strong> cavitation. This is the result of the counter-flow pressure gradients arising<br />

in the case of exp<strong>and</strong>ing gap that may lead to the possibility of flow reversal <strong>and</strong><br />

separation as determined by relation (3.38).<br />

Validity of the pressure equation<br />

In arriving at solution (3.91) we assumed that we can use the solution of combined<br />

Couette-Poiseuille flow given by (3.36). But that solution was obtained under the<br />

assumption that flow velocity has ) only component which depends only on one<br />

¦f<br />

® coordinate - . This assumption made the )Gj) j convective term in the Navier-<br />

Stokes equation (3.2) equal to zero. However, this assumption is not strictly valid<br />

in our case since we have a non-zero vertical velocity component,•, <strong>and</strong> both<br />

horizontal <strong>and</strong> vertical velocity components are functions ® of <strong>and</strong> . In this situation<br />

we can still justify dropping the convective term if we use the assumption of<br />

Stokes flow, that is, consider the inertial forces to be negligible as compared to<br />

viscous forces. Mathematically this means that:<br />

¦f}<br />

or expressed in terms of , ® , ) ,•<br />

have two relations:<br />

¦f› 9 )<br />

1<br />

we<br />

)W)<br />

>•§)<br />

>^)<br />

From the form of these equations we can deduce more specific relations between<br />

the parameters of our problem (, , , ). In particular, if we impose<br />

Ç<br />

condition<br />

of a narrow gap: , then it would lead Ç <strong>and</strong><br />

› 9 : Y : Y Y ® . With<br />

these conditions we can neglect the equation for•<strong>and</strong> simplify the ) -equation by<br />

neglecting a smaller order term on the RHS:<br />

› ›<br />

to•<br />

)G•<br />

<br />

>•I• 9<br />

¯›<br />

>•<br />

¯5¯<br />

¯› 9


<br />

<br />

<br />

@<br />

)<br />

<br />

¯m¯<br />

3.5. BOUNDARY LAYERS 81<br />

Figure 3.4: Integral analysis of a boundary layer<br />

9<br />

Both terms on the LHS are of the same order <strong>and</strong> can be approximated by<br />

¯›<br />

the order of the first one, which isr<br />

, we finally have:<br />

: Ç<br />

. Approximating the term on the RHS as<br />

)W)<br />

>•§)<br />

9:<br />

(3.92)<br />

<br />

r<br />

Ç 9<br />

›<br />

This condition together with justify the Stokes flow assumption <strong>and</strong> constitute<br />

the validity limits for equation (3.91).<br />

› Ç<br />

3.5 Boundary layers<br />

3.5.1 Flat plate integral analysis<br />

We shall consider a two-dimensional flow above a semi-infinite plate (Fig.3.4).<br />

Our objective is to introduce a quantitative measure of the thickness of the<br />

boundary layer, <strong>and</strong> to estimate it’s growth with the distance from the edge of the<br />

plate, .


¤<br />

<br />

`<br />

Ö<br />

¤<br />

V—<br />

¤<br />

V<br />

Öšœ<br />

1 ;4R ®<br />

¤<br />

`<br />

<br />

Ö `<br />

Vœ<br />

V<br />

Ö˜œ<br />

V<br />

<br />

1 ; )<br />

Ö<br />

V<br />

Ò<br />

œ<br />

[<br />

1<br />

@<br />

4<br />

)<br />

1<br />

@<br />

)<br />

1 4<br />

®<br />

<br />

<br />

MR<br />

><br />

V<br />

֜<br />

` Ö —œ<br />

Vœ<br />

)<br />

)<br />

r<br />

4<br />

<br />

82 CHAPTER 3. LAMINAR FLOWS<br />

The displacement thickness<br />

The balance of mass dictates:<br />

(3.93) <br />

R ®<br />

from which we get<br />

<br />

1U ` þ 4<br />

1<br />

`<br />

1\>])<br />

4<br />

R ®<br />

R ®<br />

` )<br />

Let’s define the displacement thickness as<br />

R ®<br />

(3.94)<br />

<br />

4<br />

` Ö þ<br />

Vœ<br />

<br />

4<br />

ß”¢×Ö<br />

` )<br />

R ®<br />

Momentum thickness<br />

General conservation of momentum dictates: ¤<br />

1 4<br />

Applied to the case of the boundary layer (Fig.3.4) it will provide the expression<br />

for a drag force, ¤, on the plate:<br />

) R ®<br />

;r<br />

expressing<br />

from (3.93), <strong>and</strong> substituting into the above, we get:<br />

r R ®<br />

` )<br />

¤ ;r<br />

®<br />

<br />

L@


(3.95) Î<br />

(3.96) Äkõ<br />

(3.97) Ĥ¢<br />

Considering that ¤<br />

(3.98) ÄŒõ<br />

<br />

¤<br />

l<br />

¤ <br />

¤<br />

<br />

@<br />

Ç Ä¤<br />

(3.99)<br />

V<br />

Ò<br />

l<br />

¤<br />

l<br />

)<br />

¢<br />

¤<br />

¤<br />

V<br />

l<br />

;r<br />

6Q<br />

1<br />

<br />

4<br />

<br />

<br />

<br />

l<br />

<br />

<br />

åĤ4<br />

<br />

l<br />

Î<br />

Ç 4 1<br />

Ç<br />

<br />

R Î l<br />

R<br />

3.5. BOUNDARY LAYERS 83<br />

We shall define the momentum thickness as:<br />

¢ÚÖ<br />

` )<br />

Using this definition we can also define the skin-friction coefficient as:<br />

<br />

L@<br />

® MƒR<br />

<strong>and</strong> the drag coefficient over the length Ç of the plate as:<br />

Î : Ç<br />

;r Ç<br />

¤¥Ö]<br />

6QR <br />

where is the distance from the edge of the plate in the downstream direction,<br />

<strong>and</strong> consequently, , we can rewrite (3.96) as:<br />

6Q<br />

<br />

¤<br />

¤<br />

1<br />

;r —<br />

;r<br />

<br />

And the inverse relation:<br />

Äkõ<br />

R <br />

Remark 3.5.1 Displacement vs Momentum thickness Displacement thickness is<br />

a more universal notion than a momentum thickness, since the former is obtained<br />

from the mass conservation, which is a always true in incompressible flows,<br />

whereas the momentum conservation may be violated due to the dissipative effects.<br />

Relation (3.97) will only hold for the flat-plate boundary layers.<br />

V<br />

Ö˜ž


(3.100) )<br />

(3.103) Äkõ<br />

)<br />

)<br />

1<br />

®<br />

1<br />

®<br />

¦<br />

Î<br />

Ä<br />

<br />

r<br />

®<br />

r<br />

@<br />

À<br />

Ï<br />

8 R ) 6<br />

® R<br />

<br />

<br />

ß<br />

8 Ñ<br />

ß ;<br />

F<br />

4<br />

84 CHAPTER 3. LAMINAR FLOWS<br />

Guessed solution<br />

Let’s use a general parabolic velocity profile:<br />

4 <br />

>^Ä<br />

®»>^Ä<br />

V<br />

Imposing the boundary conditions appropriate for the boundary layer flow:<br />

1UT 4 T<br />

1 ßJ4 <br />

<br />

1 ßJ4 T<br />

we can determine the constants Ä<br />

)<br />

, <strong>and</strong> obtain:<br />

)<br />

<br />

¯<br />

®<br />

ß<br />

l ` ® 1<br />

ß<br />

4 <br />

<br />

Substituting this solution into (3.94) <strong>and</strong> (3.95), we obtain:<br />

(3.101)<br />

ߧ<br />

(3.102)<br />

l ß<br />

Using the definition of 6 (1.28):<br />

we can rewrite (3.96) as:<br />

Äkõ<br />

8 l<br />

R )<br />

R ® ;r<br />

<strong>and</strong> using the parabolic velocity profile (3.100), we obtain:<br />

On the other h<strong>and</strong>, according to (3.98), <strong>and</strong> (3.102), we can write:


–<br />

P<br />

¦<br />

s<br />

¦<br />

)<br />

<br />

ß<br />

ß<br />

ß<br />

R<br />

<br />

ß<br />

r<br />

)<br />

¦<br />

<br />

<br />

À<br />

@<br />

Ñ<br />

Ï<br />

@<br />

Ï 8<br />

¦f“ 9 )<br />

T<br />

<br />

R<br />

ß<br />

r<br />

)<br />

) ¦<br />

X<br />

T<br />

¦<br />

3.5. BOUNDARY LAYERS 85<br />

Äkõ<br />

R <br />

Equating this to (3.103), we get:<br />

ß <br />

<br />

Integrating, we obtain:<br />

;<br />

R<br />

(3.104)<br />

8 T<br />

<br />

;<br />

Introducing the Reynolds number as: A#<br />

:])<br />

; ><br />

¦ 4 <br />

¦ì¦<br />

m ¦ 1<br />

¦f<br />

m ¦ 4<br />

1<br />

sà>])<br />

°<br />

s ;<br />

> 9 )<br />

>^)


)<br />

)<br />

F<br />

(3.107) )<br />

<br />

F<br />

ã<br />

<br />

<br />

<br />

F<br />

)<br />

F<br />

ã<br />

r<br />

<br />

<br />

F<br />

r<br />

)<br />

F<br />

F<br />

<br />

F<br />

ã s<br />

F<br />

<br />

F<br />

)<br />

r<br />

F<br />

<br />

)<br />

)<br />

ß<br />

Y<br />

Y F<br />

<br />

r<br />

F<br />

F<br />

ã r<br />

s<br />

¦<br />

<br />

V<br />

¾<br />

r<br />

r<br />

Y<br />

Y <br />

F›<br />

F<br />

)<br />

)<br />

s ã<br />

r<br />

><br />

) ã<br />

<br />

F<br />

X ã<br />

1<br />

r<br />

><br />

<br />

F<br />

¤<br />

ã<br />

<br />

X<br />

V<br />

`<br />

V r X<br />

) ;<br />

F<br />

T x<br />

r<br />

T<br />

V<br />

<br />

V<br />

¦ V<br />

V<br />

"<br />

<br />

F<br />

86 CHAPTER 3. LAMINAR FLOWS<br />

¦( 1 r 4<br />

where we shall assume a 2D 'q<br />

approximation<br />

¦G 1 r 4<br />

for the boundary layer: ,<br />

, that is, we consider that there is no variation in the span-wise direction.<br />

Consider also, that the layer is thin:<br />

'*)<br />

rŸ›<br />

)<br />

which in our particular case is reflected in relation (3.105):<br />

r ¦<br />

@ r<br />

F}É A<br />

Therefore,<br />

Using this assumption we can introduce the following scales for dimensionless<br />

variables.<br />

@ r<br />

F}É A<br />

(3.108)<br />

V ã ¾<br />

r“<br />

" )<br />

A F}É<br />

ã<br />

r ¾ r*<br />

F ã ¾ r<br />

ã <br />

V<br />

(3.109)<br />

<br />

ã<br />


¤<br />

(3.114) „©z<br />

If we apply the momentum equation (3.111) at the wall (®<br />

)<br />

)<br />

¯<br />

<br />

'<br />

¢<br />

F<br />

'<br />

)<br />

F<br />

r<br />

T<br />

¢<br />

)<br />

F<br />

<br />

)<br />

F<br />

<br />

¯<br />

><br />

¢<br />

–<br />

P<br />

¯<br />

<br />

V<br />

s<br />

r<br />

<br />

F<br />

<br />

F<br />

T<br />

¢<br />

F<br />

<br />

r<br />

¢<br />

<br />

)<br />

T<br />

F<br />

), where )<br />

¢<br />

)<br />

r<br />

F<br />

<br />

1<br />

<br />

¢<br />

F<br />

'<br />

)<br />

<br />

F T F<br />

F<br />

3.5. BOUNDARY LAYERS 87<br />

x where is the Froude number (2.111), <strong>and</strong> we introduced the dimensionless<br />

Eckert number:<br />

¢ )<br />

[2].<br />

V<br />

Equations (3.110) - (3.113) represent the Pr<strong>and</strong>tl boundary layer equations<br />

Remark 3.5.2 Parabolic character Note, that all the second derivatives over<br />

have disappeared from the momentum equations. This means the that equations<br />

of motion are parabolic with respect to direction . This in turn enables to use<br />

simpler solution procedures.<br />

Flow separation<br />

, <strong>and</strong> consider the gravity forces acting normal to the wall ( J¦ <br />

r 1<br />

T 4 <br />

- ), we obtain:<br />

T 4 <br />

% T<br />

)<br />

¯5¯


(3.117)<br />

¢<br />

<br />

4<br />

)<br />

1<br />

)<br />

¡<br />

1<br />

Ø<br />

1<br />

1<br />

1<br />

)<br />

1<br />

ß<br />

)<br />

<br />

1<br />

1<br />

¦<br />

o<br />

9<br />

l 9 <br />

®o<br />

Y Ø 4<br />

® Y<br />

` Y Ø 4<br />

Y<br />

4<br />

¯<br />

šg<br />

<br />

<br />

4<br />

š<br />

1<br />

T<br />

<br />

T<br />

<br />

4<br />

4<br />

l 9 <br />

4o šg1K<br />

1<br />

)<br />

1<br />

1<br />

4<br />

88 CHAPTER 3. LAMINAR FLOWS<br />

which should be solved with boundary conditions:<br />

(3.116)<br />

T 4 1<br />

T 4 <br />

('<br />

('<br />

Relation (3.104) indicates that the boundary layer growth with as:<br />

('K 4 <br />

('K 4 <br />

Considering this, let’s look for a similarity solution )<br />

4 <br />

® : ß


`m<br />

@ 91<br />

lm<br />

(3.123) š<br />

)<br />

<br />

<br />

`<br />

<br />

`<br />

Ø<br />

š<br />

<br />

`<br />

Cšgg<br />

<br />

<br />

l<br />

Ø<br />

`ml 91 @ <br />

lm<br />

@<br />

lm<br />

<br />

)<br />

¯<br />

1<br />

Cšgg—<br />

<br />

¯<br />

Cšgg<br />

)<br />

¯<br />

`<br />

l<br />

<br />

)<br />

l<br />

<br />

l<br />

<br />

9 o<br />

@<br />

<br />

<br />

<br />

Ø<br />

¯<br />

4<br />

<br />

mGšg š2><br />

<br />

) 9<br />

4<br />

<br />

l<br />

`"<br />

9 l<br />

<br />

r<br />

`<br />

l<br />

<br />

<br />

<br />

š<br />

<br />

4<br />

4<br />

š<br />

šgg<br />

šgg<br />

šggg<br />

4<br />

3.5. BOUNDARY LAYERS 89<br />

Since<br />

set šg1K šg1K<br />

4<br />

<br />

is a constant, we can incorporate it into ¡ 1 <br />

, which is to say, that we<br />

:<br />

@ without loss of generality, <strong>and</strong> obtain for ¡ 1 <br />

¡ 1<br />

4 m<br />

9©<br />

Now the expressions of other terms in (3.115):<br />

4<br />

1 4<br />

(3.119)<br />

(3.120)<br />

9§$š<br />

('*®<br />

4 m<br />

Cšg<br />

<br />

1<br />

kšg`<br />

<br />

<br />

l<br />

šg<br />

4<br />

o <br />

)W)<br />

1<br />

kšg`<br />

r<br />

l<br />

Œšgg šgg šg<br />

(3.121)<br />

r<br />

l<br />

¯m¯<br />

And substituting them into (3.115), we obtain:<br />

ššgg<br />

<strong>and</strong> finally:<br />

(3.122) šggg>Þš$šgg T<br />

»>Þšggšg<br />

`<br />

šggšg<br />

which should be solved with the boundary conditions:<br />

šggg<br />

1}T 4 <br />

T šg1K 4<br />

<br />

@<br />

šg1}T 4


V<br />

Ò<br />

o<br />

l<br />

1<br />

@<br />

9 <br />

(3.124) Î<br />

Ö<br />

Ö<br />

V<br />

V<br />

Ò<br />

Ò<br />

@<br />

@<br />

`<br />

Î<br />

`<br />

<br />

o<br />

l<br />

9 <br />

Y ) 8<br />

® Y<br />

<br />

<br />

<br />

V<br />

Ò<br />

V<br />

o<br />

l<br />

1<br />

@<br />

Ò<br />

š<br />

`<br />

9 <br />

@<br />

9<br />

`<br />

V<br />

Ò<br />

V<br />

l<br />

l<br />

<br />

9 o<br />

Ò<br />

`<br />

T<br />

R<br />

Ö<br />

V<br />

Ò<br />

@<br />

`<br />

9<br />

T<br />

I<br />

R<br />

90 CHAPTER 3. LAMINAR FLOWS<br />

<br />

where the last condition also implies that . This is a Blasius equation.<br />

It has no analytical solutions in a general case, but can be solved numerically.<br />

šgg1K<br />

Knowing š function<br />

the boundary layer.<br />

1 4<br />

Displacement thickness (3.94):<br />

, we can obtain important integral characteristics of<br />

4<br />

ß ¢ Ö<br />

` )<br />

R ®<br />

<br />

4<br />

<br />

Ö<br />

šg1<br />

šg4<br />

ߧ<br />

4 <br />

@‰I<br />

Momentum thickness (3.95):<br />

¡i Ò 1 `<br />

@[£¢Wo<br />

l<br />

<br />

dfeKg<br />

where we can estimate the integral as:<br />

<br />

Ö<br />

šg1<br />

šg4<br />

±Áš<br />

ššggR<br />

šg1<br />

<br />

R šg4<br />

šg4 ²<br />

Using equation (3.122) <strong>and</strong> the boundary conditions (3.123), we obtain:<br />

šgggR<br />

<br />

šg1<br />

¥Ö<br />

R šg4<br />

šgg1UT 4 <br />

Thus:<br />

Ñ[‰Ôy[<br />

šgg1}T 4o<br />

For the wall shear stress we have:<br />

(3.125)<br />

6Q<br />

; šgg1UT 4<br />

l<br />

The friction coefficient:


l<br />

;r 6Q<br />

(3.126) Äkõ<br />

@<br />

Ç Ä¤<br />

(3.127)<br />

<br />

4<br />

<br />

)<br />

<br />

V<br />

1<br />

Ö<br />

V<br />

ž<br />

®o 1<br />

<br />

<br />

1<br />

<br />

<br />

V<br />

)<br />

4<br />

Î<br />

<br />

¯<br />

<br />

<br />

4<br />

<br />

<br />

l<br />

<br />

<br />

<br />

9 l<br />

<br />

4<br />

9<br />

3.5. BOUNDARY LAYERS 91<br />

šgg1}T 4o<br />

<strong>and</strong> comparing the latter to (3.124), we obtain:<br />

Äkõ<br />

©<br />

l<br />

And the total drag on the plate:<br />

1 Ç 4<br />

Äkõ<br />

R <br />

Äkõ<br />

Wedge flows<br />

A similarity solution for the wedge flows was found as an extension of the Blasius<br />

solution. This solution, called after the authors . This solution is obtained by<br />

eliminating -velocity from the continuity equation:<br />

` —åÖ ¯<br />

) R ®<br />

<strong>and</strong> substituting it into the momentum equation:<br />

` —åÖ ¯<br />

)W)<br />

) R ®<br />

> 9 ) ¤<br />

¯m¯<br />

is a free-stream velocity profile. Then introducing the dimension-<br />

<br />

less variable, :<br />

where<br />

1<br />

<strong>and</strong> looking for the solution in the form:<br />

>¼@<br />

41<br />

('*®<br />

šg1 4<br />

4 <br />

1


š<br />

4<br />

Ø<br />

1<br />

4<br />

¦<br />

º<br />

<br />

`<br />

1<br />

@<br />

š<br />

`<br />

<br />

1<br />

±<br />

P<br />

|<br />

šgr<br />

4 ²<br />

Î<br />

92 CHAPTER 3. LAMINAR FLOWS<br />

4<br />

1 4<br />

where is the first derivative of the so-called Blasius š<br />

stream function, šg1<br />

.<br />

With these assumptions one can obtain the Falkner-Skan equation in terms of<br />

:<br />

1 4<br />

l ¹: 1<br />

@‰><br />

where , <strong>and</strong> the boundary conditions are the same as (3.123). The<br />

Ë<br />

Falkner-Skan formulation is consistent with the free-stream velocity distribution of<br />

the type:<br />

šggg>Þš$šgg>^Ë<br />

4 T<br />

4 £¸<br />

I happens so that this equation provides similarity solutions that represent<br />

wedge flows, with the stream function of the type:<br />

1<br />

',Î<br />

>¼@<br />

1 ¢<br />

¢ |<br />

DGFdefÆ<br />

Wall suction or blowing<br />

The case of wall suction or blowing can be modeled in the Blasius solution by a<br />

non-zero vertical velocity at the plate surface. It can be seen from (3.120) that this<br />

be accomplished by specifying a non-zero value of the Blasius stream function at<br />

the wall:<br />

T 4 <br />

1}T 4o<br />

('<br />

l<br />

<br />

An interesting feature of the solution is the effect of blow-off of the boundary<br />

9<br />

layer, where ) the velocity becomes identically zero. This occurs at the blow-off<br />

limit š of :<br />

1}T 4 ` T<br />

I4[Ã@¡Ôm l<br />

3.5.4 Reynolds analogy<br />

An important empirical relation for the flat plate flows with heat transfer is called<br />

the Reynolds analogy. To introduce it, let’s recall that the Pr<strong>and</strong>tl number was<br />

defined as a relation between momentum transport to the heat transport (1.30):<br />

¢C¢ 8 –<br />

°


Ä<br />

~<br />

P<br />

(3.128) Ä<br />

Ä<br />

u<br />

6<br />

<br />

…<br />

~<br />

ô<br />

~<br />

¢<br />

<br />

8 R ) <br />

® R<br />

<br />

…<br />

°<br />

6 l<br />

;r<br />

s<br />

ô<br />

Ç<br />

<br />

…<br />

<br />

r<br />

3.5. BOUNDARY LAYERS 93<br />

Ä where is the specific heat at constant pressure, (1.24), is the coefficient<br />

of viscosity, defined as a proportionality constant between the shear stress <strong>and</strong><br />

velocity gradient (1.28):<br />

8<br />

<strong>and</strong> ° is the heat conduction coefficient defined through the relation between the<br />

heat flux <strong>and</strong> temperature gradient (1.29):<br />

,¯<br />

¦$<br />

`t°<br />

¦<br />

The Nusselt number was introduced to quantify the heat conduction at the<br />

walls with the temperature difference [^s (2.114):<br />

where …<br />

<br />

is the absolute value of the heat flux:<br />

Reynolds analogy is the statement of proportionality between the heat flux<br />

<strong>and</strong> shear stress at the wall:<br />

… O<br />

¦O(±/…<br />

²GÚ´ : 1<br />

. 4 ° :‰. <br />

).<br />

[^s<br />

<br />

…Ú¦<br />

which is usually formulated as relation between dimensionless Stanton number,<br />

6Q<br />

:<br />

¢ u<br />

<strong>and</strong> friction coefficient, Äkõ (3.11):<br />

A! 5º_x<br />

such that<br />

Äkõ<br />

£¡<br />

x ÄkõEº9¥<br />

where <strong>and</strong> are empirical constants depending on geometry. The Reynolds<br />

analogy is approximately valid for shear layers, boundary layers, <strong>and</strong> pipe flows.


X<br />

<br />

(3.130)<br />

¦$<br />

)<br />

<br />

)<br />

)<br />

9 ¦K<br />

)<br />

<br />

l<br />

)<br />

<br />

F<br />

><br />

¯<br />

¯<br />

<br />

4<br />

T<br />

l<br />

<br />

l<br />

4<br />

T<br />

94 CHAPTER 3. LAMINAR FLOWS<br />

3.5.5 Free shear flows<br />

In this section we shall be looking for similarity solutions of free shear flows. Such<br />

solutions can only be accurate in the regions of the flow far from the disturbances<br />

that generate the characteristic flow patterns.<br />

In the case of free shear flows we can still assume that the boundary layer<br />

approximations (3.107) are valid when the Reynolds number is<br />

sufficiently r“<br />

large.<br />

Consequently, we could neglect the X<br />

vertical pressure gradient, , (3.112). In<br />

addition to this we shall assume that the axial pressure gradient is small<br />

T<br />

as well:<br />

, which reflects the fact that free-flows by their nature are not pressuredriven<br />

flows. Thus, the equations of motion are:<br />

F§¦<br />

¦f ¦G T #­<br />

@‰'<br />

'„¨<br />

rðr #<br />

¦í 1<br />

which in terms <br />

of explicit )<br />

variables , <strong>and</strong><br />

to the Blasius equations for the boundary layer (3.115):<br />

('q®<br />

1 ¦a<br />

)_'<br />

becomes identical<br />

@‰'<br />

(3.129)<br />

9 )<br />

)W)<br />

> )<br />

¯m¯<br />

Shear layer<br />

Consider a shear layer created by two streams with velocitiesF<br />

<strong>and</strong>r<br />

uniform<br />

initially separated by a flat plate. After passing the plate the streams mix forming a<br />

shear layer. The treatment of this case is done similarly to the Blasius approach,<br />

by selecting the non-dimensional variables as in (3.117), <strong>and</strong> looking for a solution<br />

in the form (3.119), but applied separately for each of the streams:<br />

®o F<br />

)<br />

which leads to the Blasius equation for each layer:<br />

šI©<br />

9 ¦<br />

F<br />

<br />

š©©©<br />

0=2<br />

T<br />

>\š0=2š©©


F<br />

8<br />

F<br />

4<br />

F<br />

š<br />

F<br />

F<br />

š<br />

é<br />

Ò<br />

r<br />

r<br />

<br />

–<br />

¥<br />

)<br />

F<br />

<br />

F<br />

T<br />

T<br />

<br />

)<br />

F<br />

)<br />

T<br />

3.5. BOUNDARY LAYERS 95<br />

Boundary conditions at far ends:<br />

1K 4<br />

r 1 `K<br />

4<br />

F<br />

r<br />

F<br />

)<br />

r 1 `K šI© @<br />

)<br />

F<br />

šI©<br />

, which is the imaginary<br />

continuation of the plate. For this purpose we need to consider boundary conditions<br />

on that plane 5 r 1}T 1}T 4 4<br />

)<br />

. One condition is derived from continuity: ,<br />

:<br />

4<br />

<br />

We can sew both solutions at a horizontal plane at<br />

1K<br />

<br />

<br />

4<br />

1UT 4 r 1}T 4<br />

1}T 4 <br />

r 1}T 4 T<br />

1UT 4 <br />

<strong>and</strong> another from the equality of shear<br />

8<br />

stresses:<br />

translates into:<br />

F<br />

š©<br />

š©<br />

<br />

¯<br />

1UT 4 <br />

8 r<br />

rq<br />

¯<br />

1}T 4<br />

, which<br />

r}ª<br />

1}T 4 °<br />

r 1UT 4<br />

šI©©<br />

; 8 r : 1 ; r<br />

where . It is an interesting fact, that in the case of the same<br />

<br />

fluid<br />

in both jets ° ( ) <strong>and</strong> zero free stream velocity in the one of the (r<br />

jets @ ),<br />

the vertical component of the second jet at infinity is a finite constant, ° r 1 `<br />

determined<br />

T<br />

I,[Ã@EÔml<br />

by , which is the same as the blow-off limit in the case of š<br />

the<br />

plate with suction or blowing considered in Sec.3.5.3.<br />

`K<br />

F<br />

šI©©<br />

F}É<br />

4<br />

Jet<br />

A free jet is characterized by the conservation of momentum in each plane:<br />

(3.131)<br />

´Ÿ<br />

; Ö<br />

= . "<br />

Ò )<br />

R ®<br />

The similarity solution for this case as obtained by Schlichting [2] is:<br />

5 This plane does not represent any physical interface between the two streams. The latter should be<br />

defined as a surface formed by flow streamlines.


Ø<br />

<br />

9 `<br />

r<br />

r F}É<br />

r<br />

F}É 9<br />

4<br />

with the equation for the Blasius stream function:<br />

šg<br />

1<br />

T<br />

(3.132) š<br />

<br />

r<br />

š<br />

1<br />

<br />

<br />

É<br />

¯<br />

<br />

É<br />

)<br />

1<br />

š<br />

`<br />

r<br />

<br />

<br />

šggg>\ššgg>\šgr<br />

1<br />

1 4 l £%«B¬Æ®­1<br />

)<br />

1<br />

´<br />

l<br />

š<br />

šg1K 4<br />

r<br />

<br />

É<br />

<br />

r<br />

<br />

T<br />

T<br />

T<br />

<br />

1<br />

96 CHAPTER 3. LAMINAR FLOWS<br />

®<br />

À 9<br />

F}É<br />

1 4<br />

F}É<br />

4<br />

šg1<br />

<br />

F}É À<<br />

À<<br />

<strong>and</strong> the boundary ) conditions (axial symmetry), ) <strong>and</strong><br />

(quiescent ambient fluid), which translate into:<br />

('K 4 <br />

('<br />

('<br />

('K 4 <br />

T 4 1<br />

T 4 <br />

<br />

The solution is<br />

1UT 4 <br />

šgg1}T 4 <br />

£<br />

4<br />

with the constant £ determined from the conservation of the total jet’s momentum.<br />

Thus, substituting (3.132) into (3.131) we obtain:<br />

£â — Ô<br />

F}É<br />

The maximum centraline velocity drops off as<br />

@[m;Ž8 <br />

l £<br />

4 <br />

T 4 <br />

)œ|aw<br />

('<br />

¦£ é F}É<br />

<strong>and</strong> the jet spreads as ¦<br />

.<br />

À< F}É<br />

Wake<br />

Wakes are flow patterns generated by bodies moving in a stagnant fluid. We need<br />

to consider these flow patterns far enough from the body for similarity solutions


(3.134) ‹<br />

<br />

T<br />

`<br />

r<br />

@<br />

é <br />

É l ¡1<br />

`<br />

r<br />

@<br />

é <br />

É l 1<br />

<br />

‹<br />

1<br />

r<br />

É<br />

<br />

1<br />

‹<br />

1<br />

ø<br />

@<br />

<br />

r<br />

<br />

r<br />

É<br />

<br />

¡<br />

®<br />

<br />

r<br />

r<br />

ø<br />

‹<br />

®<br />

<br />

<br />

we<br />

r<br />

r 4h`j|k<br />

r<br />

¯<br />

1<br />

)<br />

<br />

1<br />

É<br />

<br />

4<br />

r<br />

T<br />

@<br />

r<br />

r<br />

ù<br />

r ¯<br />

r<br />

3.5. BOUNDARY LAYERS 97<br />

to be valid. Since it is more convenient to select a coordinate system moving<br />

with the body, we will be considering a non-moving body immersed in a fluid with<br />

the free stream velocity. It is also convenient to describe the wake in terms of<br />

deviation of velocity from the free-stream velocity:<br />

‹ since usually constitutes a small value compared ) to. ‹ Replacing with in<br />

(3.129), ‹ <strong>and</strong> considering obtain:<br />

› ›<br />

<strong>and</strong><br />

('*®<br />

('*®<br />

<br />

‹<br />

4 <br />

<br />

`<br />

(3.133) ‹<br />

9<br />

¯m¯<br />

with the boundary conditions:<br />

4 <br />

('`¯K<br />

T 4 <br />

('<br />

Equation (3.133) is of a heat-conduction type, with the solution:<br />

4 <br />

` ®<br />

('*®<br />

—<br />

Substituting it into (3.133) we can find :<br />

rhjlk<br />

F}É<br />

<br />

` ®<br />

>à é F}É<br />

<br />

—<br />

£¡<br />

` ®<br />

` l 9<br />

¿<br />

` l 9<br />

®h`j|k —<br />

¡<br />

` l ®<br />

ùhjlkû— `<br />

®<br />

<br />

` l 9<br />

` l ®<br />

>à é F}É<br />

<strong>and</strong> at ®<br />

we obtain:<br />

r 4 <br />

r —<br />

<br />

(3.135)<br />

Ó<br />

Ñ 9


where ¤<br />

é<br />

Ò<br />

Ò<br />

T<br />

¤<br />

@<br />

l Ĥ;r<br />

¤<br />

¦<br />

R<br />

ß<br />

é<br />

Ò<br />

Ò<br />

T<br />

Ç<br />

ß<br />

<br />

ß<br />

U9<br />

<br />

T<br />

98 CHAPTER 3. LAMINAR FLOWS<br />

The constant is evaluated from the second Newton law, that the drag force<br />

should equal to the momentum deficit in the wake:<br />

¡<br />

¤<br />

<strong>and</strong><br />

[J°<br />

[J°<br />

¥Ö<br />

;<br />

Ö<br />

l ;¡ o<br />

with the final result:<br />

(3.136)<br />

; )W‹ R ®¦<br />

¡¼<br />

ĤÇ<br />

Ñ<br />

U9<br />

o<br />

; ‹ R ®<br />

When substituted into (3.134) it shows that the wake defect is proportional to body<br />

drag coefficient.<br />

3.6 Integral methods<br />

To derive the integral formulations for the boundary layer equations, discussed in<br />

Sec.3.5.2, let’s consider the conservation of mass <strong>and</strong> momentum in a controlvolume<br />

formed by four ((' points: ), as depicted<br />

in Figure 3.5.<br />

), (('<br />

), (t> R ('<br />

From the conservation of mass we have:<br />

> R<br />

), (t> R ('<br />

£‰¦$<br />

?<br />

)<br />

£‰¦<br />

where is the vector area element. Summing up over all the four faces, <strong>and</strong><br />

R<br />

considering that the flow is uniform in the transverse direction (2D flow), we can<br />

exp<strong>and</strong> the equation above as:


Q is<br />

`<br />

V<br />

Ö˜±<br />

; ) R ®<br />

`<br />

;R<br />

; ) R ®<br />

w¦<br />

ß<br />

¤<br />

<br />

<br />

ß<br />

><br />

R<br />

R<br />

; ) R ®C> R<br />

¦$ w¦ š$›M)œ<br />

; ) R ®»> ; QR <br />

)<br />

4<br />

; ) R ®<br />

<br />

T<br />

3.6. INTEGRAL METHODS 99<br />

Figure 3.5: Control volume for an integral formulation<br />

where the normal velocity at the wall, which is non-zero in the case of wall<br />

suction or effusion. Rewriting the second integral as:<br />

V<br />

Öš±D¤±<br />

<strong>and</strong> considering constant density, we obtain:<br />

V<br />

Öš±D¤±<br />

V<br />

Öš±<br />

V<br />

Ö±<br />

(3.137) <br />

) R ®t> Q<br />

R <br />

In a similar manner, considering the momentum conservation, we obtain:<br />

V<br />

Ö˜±<br />

R<br />

¦ 1<br />

where the summation is done over all faces, <strong>and</strong> we are considering projection<br />

of forces <strong>and</strong> fluxes on the horizontal direction. Writing the terms explicitly, we<br />

obtain:


Now we eliminate R<br />

X<br />

ß<br />

><br />

R X `tß<br />

R<br />

)<br />

r<br />

1<br />

<br />

R X `§ß<br />

R<br />

> ;<br />

R<br />

R<br />

X R<br />

R<br />

<br />

R<br />

` ;<br />

><br />

Ö<br />

V±<br />

`<br />

V r<br />

) ;<br />

r<br />

R<br />

ß<br />

Î<br />

` ;<br />

<br />

Ö<br />

V±<br />

`<br />

<br />

; )<br />

R<br />

R<br />

r<br />

r<br />

R<br />

<br />

`<br />

Ö<br />

V±<br />

; )<br />

; )<br />

R<br />

R<br />

r<br />

r<br />

R<br />

Î<br />

ß<br />

; )<br />

4<br />

r<br />

100 CHAPTER 3. LAMINAR FLOWS<br />

ߧ` 1<br />

4 1 ß<br />

ß R X : l 4<br />

6QR <br />

X><br />

R X<br />

> R<br />

;r<br />

R ®<br />

` Ö ±<br />

R ®C><br />

R ®<br />

where X is the pressure. Canceling terms <strong>and</strong> neglecting higher order terms, we<br />

obtain:<br />

(3.138)<br />

6Q<br />

> ;r<br />

R ®<br />

R <br />

<br />

: R <br />

ß<br />

by mens of (3.137), so that (3.138) becomes:<br />

V<br />

Ö˜±<br />

(3.139)<br />

6Q<br />

R ®<br />

) R ®<br />

One can relate the pressure gradient to the free-stream velocity gradient:<br />

<br />

V<br />

Öš±<br />

Q<br />

`<br />

V<br />

Ö˜±<br />

` ;9<br />

It is also convenient to express (3.139) in terms of displacement thickness, (3.94)<br />

<strong>and</strong> momentum thickness (3.95):<br />

R <br />

R<br />

1 ßt`]ߧ,4<br />

) R ®<br />

1 ßt`]ß `<br />

R ®<br />

) R ®<br />

with these transformations (3.139) can be written as:<br />

V<br />

Ö˜±<br />

Ö˜±<br />

V


À<br />

<br />

l ß<br />

1 4<br />

®<br />

ß<br />

<br />

¥<br />

<br />

<br />

s<br />

V<br />

@<br />

<br />

`<br />

<br />

Î<br />

<br />

l<br />

1<br />

<br />

Î<br />

><br />

<br />

@<br />

><br />

s<br />

F<br />

<br />

Î<br />

<br />

`<br />

`<br />

<br />

1<br />

<br />

@ R<br />

R r<br />

ߧ><br />

Q<br />

<br />

<br />

À<br />

l<br />

Î<br />

V<br />

4<br />

)<br />

Î<br />

r<br />

Î<br />

<br />

<br />

`<br />

`<br />

`<br />

Q<br />

QQ<br />

<br />

Q<br />

`<br />

3.7. PROBLEMS 101<br />

<br />

><br />

) R ®<br />

R ®<br />

;r<br />

R <br />

R 1 ßC`]ß”*4a`<br />

@ R<br />

R r<br />

r<br />

ß<br />

<br />

R<br />

6Q<br />

Ö±<br />

1 ߧ`àß”c`<br />

*4a`<br />

V<br />

Ö˜±<br />

<br />

><br />

1 ßt`]ߧ,4<br />

ßt`àߧê`<br />

,4a`<br />

1 ßt`]ߧÛ`<br />

,4<br />

ß<br />

<br />

R<br />

<br />

`<br />

<br />

><br />

>^Î<br />

which finally becomes:<br />

1 l<br />

Î ><br />

For steady flow with an impermeable wall we obtain:<br />

<br />

;r 6Q<br />

ߧ,4<br />

(3.140)<br />

Äkõ<br />

l<br />

ߧ,4<br />

1 l<br />

which is called a von Karman integral momentum relation.<br />

<br />

Î ><br />

3.7 Problems<br />

Problem 3.7.1 Couette flow equations<br />

Show how to obtain equation (3.7) <strong>and</strong> (3.8).<br />

Problem 3.7.2 Couette plates solutions<br />

Solve equations (3.7) <strong>and</strong> (3.8) with the boundary conditions u(0) = 0 <strong>and</strong><br />

u(H) = U, s <strong>and</strong> s <strong>and</strong> .<br />

1}T 4 <br />

1} 4 <br />

Problem 3.7.3 Flow of a liquid film<br />

Ï ¿ , flowing steadily<br />

Consider a wide fluid film of ’ @‰I<br />

constant<br />

T<br />

thickness,<br />

due to the gravity down the inclined plate Î<br />

{<br />

at angle . Find an analytical<br />

expression of a fluid velocity distribution as a function of a distance from the plate<br />

<br />

surface: . Assuming the viscosity <strong>and</strong> the density of the fluid are @‰I4[Jˆ<br />

°<br />

) 1 ¹. 4 WT‰T ° :< <br />

respectively, find the volumetric flow rate, 8 , per … :<br />

1m of the plate. Atmospheric pressure can be considered constant.<br />

, <strong>and</strong> ¢ ’


I<br />

º<br />

¾<br />

r<br />

ë<br />

X<br />

<br />

R<br />

R<br />

r<br />

ë<br />

R<br />

R<br />

<br />

r<br />

Ñ<br />

ˆ<br />

l<br />

T<br />

@<br />

Ï<br />

F<br />

<br />

T<br />

<br />

102 CHAPTER 3. LAMINAR FLOWS<br />

Problem 3.7.4 Couette solution for non-Newtonian fluids<br />

How will the solution (3.9) change for a non-Newtonian fluid?<br />

Problem 3.7.5 Momentum equation for Couette flow between concentric cylinders<br />

Using the assumptions on the Couette velocity profile between the rotating<br />

concentric cylinders (Sec.3.2.3) <strong>and</strong> the expression for the momentum equation<br />

<strong>and</strong> the Laplacian operator in cylindrical coordinates:<br />

<br />

)IF­>])Wx5)GF<br />

x_><br />

¢ >^)GH,)GFH­><br />

¢<br />

)œx5)IF<br />

¼<br />

F`<br />

)GFœ)GFF<br />

)œx<br />

; ¢ > 9 1 ¾<br />

)GF­><br />

r `<br />

¢ ¢ r 4 )GF<br />

@<br />

¢<br />

<br />

x<br />

4<br />

x_><br />

@ r ë ¢<br />

F„F­>Þë<br />

F<br />

F<br />

HH<br />

1 ¢<br />

Derive equation (3.18):<br />

¢ r ><br />

<br />

Problem 3.7.6 Rotation torque <strong>and</strong> power<br />

¢NM ¢EL)IF )GF<br />

In the system of two rotating cylinders (Sec.3.2.3) consider the torque applied<br />

to the inner rotating cylinder when the outer cylinder is (n fixed ). What<br />

is the power required to rotate the inner cylinder?<br />

Problem 3.7.7 Flow between parallel plates under pressure<br />

A viscous fluid with viscosity (<br />

parallel plates<br />

8<br />

` l<br />

ωÏ<br />

) is driven between two<br />

`ÞÑ•° : 1 ¹. 4<br />

Ï @‰I<br />

¿ apart by an imposed pressure gradient of º : R R<br />

volume flow rate per 1m of the plates’ width. What pressure gradient will cause<br />

the flow to reverse?<br />

£ :< . The upper plate is moving with velocity<br />

<br />

– ¹:‰.<br />

. Find the<br />

Problem 3.7.8 Verifying the Stokes solution<br />

Verify that the solution (3.73) satisfies the equation (3.72).<br />

Problem 3.7.9 Stokes velocity <strong>and</strong> shear stress<br />

Using Stokes solution for the stream function (3.73), obtain velocity components<br />

(3.74), (3.75) <strong>and</strong> the shear stress component 6 x„F(3.77).


Hint: Assume )<br />

@<br />

<br />

T<br />

Ï<br />

( 1 ß : <br />

º<br />

<br />

) <br />

<br />

( 1 ß : <br />

<br />

@<br />

(Äkõm A"<br />

@<br />

<br />

@<br />

<br />

T<br />

3.7. PROBLEMS 103<br />

Problem 3.7.10 Falling sphere in oil<br />

Ý<br />

A sphere of density dropped into oil of density {<br />

<strong>and</strong> viscosity 8 T £ .<br />

. Estimate the terminal velocity of the sphere if its<br />

; ;<br />

IK@<br />

diameter is (a) R<br />

<strong>and</strong> (c) R<br />

<br />

T ¿ .<br />

¢IW;—rZ²<br />

is<br />

IWeW;—rZ²<br />

@ ¿<br />

Problem 3.7.11 Boundary layer analysis for cubic velocity profile<br />

Repeat the boundary layer analysis of Sec.3.5.1 with assumed velocity profile:<br />

IO@ ¿ , (b) R<br />

<br />

À<br />

L®<br />

l<br />

@<br />

L®<br />

ßM<br />

l<br />

ßM<br />

`<br />

4m A! ),<br />

4m A! ),<br />

4m A" ),<br />

Compute ( Î : <br />

1 :‰9<br />

. ©<br />

(Ĥm<br />

where A!<br />

<br />

.<br />

1}T 4 T<br />

¯m¯<br />

),<br />

A" ),<br />

Problem 3.7.12 Drag force on a triangle<br />

A thin equilateral triangle plate is immersed parallel to a stream of air with<br />

velocity<br />

<br />

l<br />

the<br />

£ "<br />

. Assuming<br />

{<br />

laminar flow, estimate the drag on this plate.<br />

l ¹:‰.<br />

at temperature s<br />

Ä <strong>and</strong> pressure º<br />

Problem 3.7.13 Boundary layer equations<br />

Derive Pr<strong>and</strong>tl boundary layer equations (3.110) - (3.113):


104 CHAPTER 3. LAMINAR FLOWS


A<br />

V<br />

1<br />

A<br />

F<br />

`<br />

9 ¦<br />

A<br />

V<br />

1<br />

n<br />

F<br />

`<br />

9 ¦<br />

n<br />

V<br />

¦<br />

l<br />

Chapter 4<br />

Turbulent flows<br />

4.1 Transition to turbulence<br />

In the case of free flows transition to turbulence occurs much earlier than in confined<br />

flows. In terms of Reynolds number, it is a matter of several hundred for<br />

the unbounded flows around objects, <strong>and</strong> a matter of several thous<strong>and</strong> for the<br />

confined flows.<br />

Thus, the transition to turbulence for the case of parallel plates usually occurs<br />

at:<br />

Ï TJT<br />

<br />

The transition to turbulence for the case of rotating cylinders is usually measured<br />

in terms of Taylor number <strong>and</strong> occurs at the critical value of:<br />

A!<br />

@<br />

4 <br />

s$w<br />

4 <br />

@³¢TJT<br />

For the flow in ducts the transition to turbulence occurs at A#<br />

.<br />

T‰T‰T<br />

4.2 Turbulence Modeling<br />

Let’s consider the incompressible forms of the mass <strong>and</strong> momentum equations,<br />

(2.4), (2.24):<br />

105


¦<br />

> )<br />

1 <br />

) )<br />

(4.3)<br />

9 9<br />

9 Ê 1<br />

Ä<br />

(4.4)<br />

Ý<br />

[<br />

V<br />

r<br />

4<br />

)<br />

r<br />

T<br />

<br />

)<br />

X<br />

Ý<br />

¢<br />

)<br />

<br />

)<br />

106 CHAPTER 4. TURBULENT FLOWS<br />

(4.1)<br />

(4.2)<br />

¦f ¦G<br />

where we introduced an abbreviation for pressure-to-density ã ratio:<br />

also use the conservative expression for the pressure ) term:<br />

which is true due to continuity (4.1).<br />

: ;<br />

. We<br />

¦K] 1 ¦ 4 X<br />

The solution to this equation system for high Reynolds numbers will result in<br />

a turbulent flow field. This field is highly unsteady with a broad spectrum of eddies,<br />

which makes it difficult, if not impossible to resolve it on even the most powerful<br />

computers. However, for some limited range of Reynolds numbers the solution<br />

can be obtained numerically. The technique that uses this direct approach of<br />

computing turbulence is called direct numerical simulation (DNS). The difficulty<br />

with this approach is that in order to reproduce all the turbulent eddies from the<br />

largest to the smallest, the simulation has to resolve the smallest space <strong>and</strong> time<br />

scales of turbulence. This in turn may require a very fine grid <strong>and</strong> time-resolution.<br />

,<br />

¦ 4 k `<br />

¦<br />

¦f}<br />

X ã<br />

> 9 )<br />

4.2.1 LES models<br />

In order to go beyond the Reynolds number limits of DNS another technique<br />

is employed. In this approach the grid cell sizes used are usually greater than<br />

the smallest turbulent eddies. This amounts to a space-averaging of the Navier-<br />

Stokes equation. In this case only the largest turbulent eddies are resolved, <strong>and</strong><br />

the unresolved eddies are modeled as an addition to viscosity:<br />

> 9 Ê<br />

where 9 V<br />

is the molecular viscosity, <strong>and</strong> 9 Ê is the eddy viscosity, associated with<br />

the cumulative action of unresolved eddies on the resolved large eddies. This<br />

approach to turbulence modeling is called Large eddy simulation. In the first<br />

proposed LES model, Smagorinsky model, the turbulent viscosity 9 Ê is set proportional<br />

to the strain-rate tensor (1.9), <strong>and</strong> the computational grid size::<br />

1 . ¦7 . ¦/ 4<br />

[ where is the grid cell size, <strong>and</strong> the coefficient of Ä<br />

proportionality,<br />

Smagorinsky constant.<br />

, called the<br />

F}É


(4.5) Š<br />

<br />

)<br />

1<br />

<br />

xÊ "54<br />

V<br />

1 It is also common to denote the fluctuating velocity asńµ<br />

)<br />

1<br />

Ý<br />

1<br />

)<br />

R<br />

"<br />

Ö<br />

V<br />

<br />

Ê<br />

…<br />

)<br />

Ý<br />

1<br />

<br />

><br />

4<br />

¦<br />

R<br />

4.2. TURBULENCE MODELING 107<br />

In the Smagorinsky model Ä the constant is considered fixed, <strong>and</strong> is selected<br />

by matching the experimental data or those produced by DNS. In more<br />

sophisticated LES Ä models is no longer considered a constant, <strong>and</strong> its value<br />

is determined in a more complex way, for example by comparing some integral<br />

measures produced from solutions obtained from space-averaged equations using<br />

different averaging filters [13]. The common feature of LES models is that they<br />

are based on solving for a time-dependent flow field using space-averaged form<br />

of the Navier-Stokes equation.<br />

4.2.2 RANS models<br />

Historically, another turbulence modeling approach was used first. This approach<br />

is based on time averaging, rather than space averaging of the NS equation.<br />

Let’s use the Reynolds assumption that any turbulent quantity, … , can be<br />

decomposed it into the time average <strong>and</strong> fluctuating components:<br />

"54 <br />

"54<br />

('<br />

('<br />

In particular, for the velocity components, we have:<br />

>^Šg1<br />

(4.6)<br />

(4.7)<br />

¦$<br />

¦<br />

X ã<br />

º >”X<br />

, <strong>and</strong> time averaging of the fluctuating compo-<br />

where1<br />

nent, , is zero 1 :<br />

4 ¢<br />

('<br />

('<br />

"54<br />

(4.8)<br />

¦$<br />

¦ 1<br />

"54<br />

" T<br />

('<br />

The averaging time is usually much greater than the time scale of turbulent fluctuations.<br />

Since the time averaging interval does not stretch to infinity, the resultant<br />

average quantities can still be slowly varying functions of time. But the time scale<br />

of these variations will be larger than largest turbulence time scales. Under this<br />

assumptions decomposition (4.8) is called the Reynolds decomposition.


><br />

><br />

><br />

><br />

À<br />

l<br />

º<br />

º<br />

¦<br />

)<br />

¦<br />

T<br />

T<br />

><br />

108 CHAPTER 4. TURBULENT FLOWS<br />

Applying (4.6) to (4.1), (4.2), we have:<br />

(4.9)<br />

<br />

¦<br />

¦<br />

1<br />

4 1¦<br />

¦ 4 ²¨k ` 1<br />

¦<br />

¦ 4<br />

¦ 4 €<br />

<strong>and</strong> after <strong>and</strong> time averaging the latter <strong>and</strong> the continuity relation (4.1), <strong>and</strong> using<br />

(4.8), we obtain:<br />

>±<br />

>ûX<br />

> 9 1¦<br />

¦f ¦G<br />

¦f ¦G<br />

(4.10)<br />

<br />

¦<br />

1¦ 4 <br />

1 ¦ 4 `<br />

¦<br />

As can be seen, both the mean <strong>and</strong> fluctuating fields satisfy the continuity. The<br />

last equation is called the Reynolds averaged Navier-Stokes equation (RANS),<br />

thus the name of the approach. As can be seen the equation contains <strong>and</strong> extra<br />

unknown term composed of derivatives of the so-called the Reynolds<br />

(¢<br />

stress<br />

¦<br />

tensor: .<br />

6<br />

To close the new equation system one needs to formulate extra equations<br />

for the components of Reynolds stress tensor (closure). One of the simplest<br />

<strong>and</strong> first closure was suggested by Boussinesq, <strong>and</strong> is called the Boussinesq<br />

approximation. In this approximation the Reynolds stress tensor is considered<br />

proportional to the mean velocity gradient:<br />

> 9¦f€<br />

(4.11)<br />

¦ k<br />

ß ¦7· `<br />

l 9 Ê ¨ ¦/<br />

(4.12)<br />

@ ä¢<br />

) l<br />

¨ ¦/k¢<br />

1¦f<br />

@<br />

l<br />

>¸m ¦ 4<br />

(4.13)<br />

where we introduced the mean strain rate tensor, analogous to (1.9). Quantity<br />

defined by (4.12) is called the Turbulent kinetic energy, <strong>and</strong> ¨ ¦7 is called Ê<br />

the eddy Both<br />

viscosity,. <strong>and</strong> 9 represent two new unknown variables in the<br />

Ê<br />

model. Empirical algebraic relations can be devised for 9 <strong>and</strong><br />

connecting<br />

Ê<br />

to¦f<br />

them <strong>and</strong> length-scales of the problem, as is done in the mixing-length<br />

9<br />

theory [2], or in a more sophisticated RANS models discussed below.<br />

On the other end of the turbulence closure spectrum is the Reynolds stress<br />

model (RSM). In this model each component of the Reynolds stress tensor is


)<br />

¦<br />

><br />

j<br />

><br />

<br />

`<br />

><br />

><br />

¦<br />

j<br />

)<br />

><br />

¦<br />

X<br />

<br />

><br />

j<br />

<br />

j<br />

`<br />

j<br />

X<br />

><br />

¦ ­j<br />

j<br />

¦<br />

¦<br />

X<br />

j<br />

`<br />

)<br />

j<br />

¦<br />

X<br />

¦<br />

j<br />

4.2. TURBULENCE MODELING 109<br />

obtained from a separate equation. Since there are six independent components,<br />

there should be six equations. In the full Reynolds stress model these equations<br />

are represented by the PDEs of the transport type. In an approximate version of<br />

RSM - algebraic Reynolds stress model these equations are given by algebraic<br />

relations.<br />

To obtain the equations for Reynolds stress tensor, let’s first subtract (4.10)<br />

from (4.9):<br />

¦<br />

1 ¦ 4 <br />

1 ¦ 4 <br />

1 ¦ 4 ` 1 ¦ 4 <br />

¦<br />

> 9 ¦f€<br />

which, using continuity (4.1), can be rewritten as:<br />

(4.14)<br />

¦<br />

¦f<br />

¦f<br />

1 ¦ ` ¦ 4 k `<br />

¦<br />

Let’s multiply this equation by )Gj :<br />

> 9 ¦K€<br />

¦<br />

¦K<br />

¦f<br />

1 ¦ ` ¦ 4 k `<br />

¦<br />

¦K€<br />

)Wj<br />

>^)Wj<br />

>])åj<br />

>^)Wj<br />

)WjmX<br />

> 9 )Wj<br />

If we swap the indexes # <strong>and</strong> ° of this equation, we get:<br />

¦ <br />

<br />

<br />

¦ 1<br />

`<br />

4 k<br />

€<br />

¦ aj<br />

9 ) jc><br />

jê>])<br />

>^)<br />

>^)<br />

Now add the two last equations:<br />

¦<br />

¦ <br />

)åj<br />

>])<br />

¦f<br />

<br />

¦f<br />

<br />

>C)Wj<br />

>^)<br />

>^)Wj<br />

>])<br />

1 ¦ ` ¦ 4 <br />

¦ 1<br />

`<br />

4 <br />

>C)Wj<br />

>^)<br />

¦ `<br />

¦f€<br />

} 4<br />

Now apply Reynolds decomposition ) (4.6):<br />

equation:<br />

¦G<br />

¦<br />

, <strong>and</strong> time average the last<br />

)Wj5X<br />

)Wj<br />

>^)<br />

jc> 9 1


1ajê><br />

j<br />

` 1aj•> j<br />

` 1aj•> j<br />

¦ <br />

)<br />

<br />

`<br />

4<br />

X<br />

<br />

><br />

4<br />

X<br />

¦ <br />

¦ <br />

)<br />

><br />

`<br />

><br />

1­jê><br />

¦ <br />

X<br />

><br />

<br />

><br />

j<br />

j<br />

j<br />

X<br />

X<br />

j<br />

<br />

<br />

X<br />

<br />

><br />

j<br />

j<br />

1­jê><br />

j<br />

j<br />

><br />

j<br />

><br />

j<br />

><br />

><br />

><br />

j<br />

1ajê><br />

j<br />

¦<br />

j<br />

><br />

><br />

><br />

j<br />

><br />

j<br />

><br />

¦ <br />

)<br />

><br />

¦<br />

1<br />

j<br />

><br />

j<br />

j<br />

><br />

><br />

><br />

¦ <br />

)<br />

><br />

j<br />

j ><br />

4 ¦ ­j<br />

j ><br />

¦ ­j<br />

j<br />

j<br />

j<br />

j<br />

4<br />

110 CHAPTER 4. TURBULENT FLOWS<br />

4 ¦<br />

1¦<br />

¦ 4 <br />

4 ¦f<br />

1¦<br />

¦ 4<br />

<br />

4 ¦f<br />

1¦<br />

<br />

4 1 ¦ ` ¦ 4 <br />

1¦<br />

¦ 4 1<br />

`<br />

4 <br />

¦ ` 1¦<br />

¦ 4<br />

4 ¦f}<br />

1¦<br />

¦ 4<br />

} 4<br />

jÛ> 9 151­j•> j<br />

which after taking into account (4.8) simplifies to:<br />

¦<br />

¦ <br />

¦<br />

4 <br />

¦f<br />

<br />

>} 1<br />

> j<br />

1 ¦ ` ¦ 4 <br />

1 ¦ 4 `<br />

1 ¦ 4 <br />

>}­j<br />

> j<br />

>}¦ 1<br />

`<br />

4 <br />

¦ 1<br />

4 ` ¦ 1<br />

4 <br />

¦ ` 1¦<br />

¦ 4<br />

4 ¦f}<br />

1¦<br />

¦ 4<br />

} 4<br />

jÛ> 9 151­jê> j<br />

changing the order of time differentiation <strong>and</strong> time averaging, <strong>and</strong> canceling terms<br />

we obtain:<br />

4 <br />

¦f<br />

<br />

jê>¸ 1 ¦<br />

¦ ­j<br />

> j<br />

(4.15)<br />

jÛ> 9 1<br />

1 ¦ 4 <br />

¦ 1<br />

4 <br />

> j<br />

¦ ` ¦<br />

¦f}<br />

} 4<br />

jqX<br />

<strong>and</strong> finally:<br />

4 <br />

¦f<br />

<br />

(4.16)<br />

jê>¸ 1 ¦<br />

j ><br />

9 1<br />

><br />

¦ ­j<br />

¦ ` ¦<br />

1 ¦ <br />

4 <br />

¦f}<br />

} 4<br />

j5X<br />

jê><br />

This is the equation for the components of the Reynolds stress tensor. Since the<br />

Reynolds stress tensor is symmetric, it has only six independent components,<br />

<strong>and</strong> the number of equations is 6. As can be seen the equation contains 3-rd<br />

order ) )Wj correlations: . One can write an equation ) )Wj for as well, but it<br />

will depend on 4-th order correlations, <strong>and</strong> so on. In practice, an empirical relationship<br />

is proposed, that links the third order ) )Wj tensor to the second order<br />

) tensor . This relationship is called closure. In fact all the terms on the RHS


is<br />

r<br />

r<br />

¹<br />

r<br />

r<br />

4.2. TURBULENCE MODELING 111<br />

of (4.16) require some kind of closure to make the problem complete. If such<br />

closures are established, the obtained equation system will constitute a turbulence<br />

model. In this particular case, when the equation system provides PDEs<br />

for Reynolds stress tensor components, the corresponding turbulence model is<br />

called Reynolds stress model (RSM). Considering that the Reynolds stress tensor<br />

is symmetric, this model may include as many as 12 PDEs: 3 - velocity, 1<br />

- pressure, 6 - Reynolds stress tensor. The 12-th equation is the one for the<br />

turbulent dissipation rate, which will be discussed in the next section.<br />

Two equation turbulence models<br />

The RSM model described above may be prohibitively expensive in terms of the<br />

number of equations <strong>and</strong> the complexity of implementation. In this case simpler<br />

models can be used, which are based on smaller number of equations. The<br />

first step to reduce the number of equations is to relate the components of the<br />

Reynolds tensor to the mean velocity gradients following the Boussinesq approximation<br />

(4.11). Then one can formulate a separate transport equation for<br />

<strong>and</strong><br />

an algebraic relation for 9 Ê as a function of the latter <strong>and</strong> the mean velocity gradients.<br />

This approach will constitute a one equation turbulence model.<br />

A more popular approach is to formulate two transport equations: one for<br />

turbulent kinetic energy,<br />

<strong>and</strong> another for it’s dissipation rate¹. The eddy viscosity,<br />

9 Ê is then related to<br />

<strong>and</strong>¹as<br />

(4.17)<br />

9 Ê <br />

Ä ©r<br />

±º¹²»<br />

±ƒ² :‰. :J. <br />

±ƒ²? :‰. ± 9 Ê ²? :‰.<br />

the `<br />

which can be shown from dimensional reasoning. Indeed, if¹represents the rate<br />

of change of: then its physical dimensions should be ,<br />

Considering that , <strong>and</strong> , we can obtain (4.17). This<br />

approach is called ¹turbulence model (KE) [14, 15], which is the most<br />

popular turbulence model for engineering computations today. The equation for<br />

formulated as a transport equation of the form:<br />

(4.18)<br />

1 9§»<br />

¦ 4 ¦ 2<br />

R<br />

"<br />

R<br />

where the effective eddy viscosity of. This equation is solved with the<br />

boundary conditions of zero<br />

at the walls. Usually a non-zero<br />

is set at the<br />

9§»<br />

> 9 ʦK 1¦K<br />

>¸m ¦ 4a`<br />

¹<br />

is


T<br />

T<br />

I<br />

F<br />

<br />

F<br />

¦f<br />

F<br />

r<br />

Ä<br />

<br />

112 CHAPTER 4. TURBULENT FLOWS<br />

open boundaries, where its value is related to level of turbulent fluctuations expected<br />

in each particular flow case. The equation for the turbulence dissipation<br />

rate,¹, is written in analogy to the one for. Namely, we multiply the-equation<br />

by¹:, <strong>and</strong> introduce the effective eddy viscosity of¹: 9”¼to obtain:<br />

(4.19)<br />

r¹r<br />

where we introduced two new Ä ',Ä<br />

constants . Boundary conditions on¹can be<br />

obtained from those on, relating¹to<br />

by through equation (4.18) where the<br />

.<br />

R<br />

"<br />

R¹<br />

non-steady term should be set to zero at the R: R<br />

" T<br />

boundary:<br />

Now equations (4.18) <strong>and</strong> (4.19) have three effective viscosities: Ê (diffusiv-<br />

9<br />

ity of momentum), (diffusivity of turbulent kinetic energy,), <strong>and</strong> 9§¼(diffusivity<br />

of turbulence dissipation rate,¹). It should be noted that unlike the molecular<br />

viscosity,<br />

V<br />

, which is a property of the fluid <strong>and</strong> is usually independent on coordinates<br />

or time 2 , all turbulent viscosities are functions of space, <strong>and</strong> are therefore<br />

represent the new dependent variables of the problem along 9 <strong>and</strong>¹.<br />

with¦<br />

' 9”»<br />

Another assumption of the model is that <strong>and</strong> 9”»<br />

9 Ê with different proportionality constants:<br />

9¼are both proportional to<br />

1 9¼¹<br />

4 ¦ ¦<br />

>\Ä<br />

9 Ê ¹<br />

1¦f<br />

>Jm ¦ 4a`<br />

(4.20)<br />

9 Ê<br />

(4.21)<br />

9¼<br />

9 Ê<br />

where j , ö¼<br />

model.<br />

ö<br />

the effective ”Pr<strong>and</strong>tl numbers”, which are constants of the<br />

ö¼<br />

are<br />

The system of equations (4.18), (4.19), (4.20), (4.21), (4.17) is closed <strong>and</strong> constitutes<br />

¹turbulence model. It has 5 empirical the `<br />

constants:<br />

9”» <br />

ö»<br />

r<br />

l ö» <br />

T ö¼<br />

@‰IÕÀ<br />

the values of which are determined by comparison of computations with experimental<br />

data.<br />

KE model belongs to the class of two-equation turbulence models. Another<br />

important two equation turbulence model is the °Z` n model [16], where instead<br />

of the turbulence dissipation rate¹a turbulence frequency scale, n , is used as an<br />

extra variable.<br />

2 In flows with heat transport this may not be the case<br />

Ä ©<br />

Ô<br />

Ä<br />

щÑ<br />

Ä @‰I<br />

@‰IÕÔ<br />

@‰I


Bibliography<br />

[1] Y.A. Cengel <strong>and</strong> M.A. Boles. Thermodynamics. McGraw-Hill, Inc., 2002.<br />

[2] Frank White. Viscous <strong>Fluid</strong> Flow. Second Edition‘. WCB/McGraw-Hill, 1991.<br />

[3] K.Jr. Wark. Advanced Thermodynamics for Engineers. McGraw-Hill, Inc.,<br />

1995.<br />

[4] C.A.J Fletcher. Computational techniques for fluid dynamics. Springer-<br />

Verlag, 1991.<br />

[5] C.R. Chester. Techniques in partial differential equations. McGraw-Hill, NY,<br />

1971.<br />

[6] J.H. Ferziger <strong>and</strong> M. Peric. Computational Methods for <strong>Fluid</strong> Dynamics.<br />

Springer Verlag, 1997.<br />

[7] L.D. L<strong>and</strong>au <strong>and</strong> E.M. Lifshitz. <strong>Fluid</strong> <strong>Mechanics</strong>. Course of Theoretical<br />

Physics, volume 6. Butterworth-Heinemann; 2nd edition, 1987.<br />

[8] O. Reynolds. An experimental investigation of the circumstances which determine<br />

whether the motion of water shall be direct or sinuous, <strong>and</strong> of the<br />

law of resistance in parallel channels. Royal Society, Phil. Trans., 1883.<br />

[9] E. Buckingham. Model experiments <strong>and</strong> the form of empirical equations.<br />

Trans. ASME, 37, 1915.<br />

[10] F. White. <strong>Fluid</strong> <strong>Mechanics</strong>. Fifth Edition‘. WCB/McGraw-Hill, 2002.<br />

[11] E. Buckingham. On physically similar systems: Illustrations of the use of<br />

dimensional equations. Phys. Rev., 4(4):345–376, 1914.<br />

[12] C.W. Pseen. Ueber die stokes’sche formel und ueber eine verw<strong>and</strong>te aufgabe<br />

in der hydrodynamik. Ark. f. Math. Astron. och Fys., 6(29), 1910.<br />

113


114 BIBLIOGRAPHY<br />

[13] M. Germano. Turbulence: the filtering approach. Journal of <strong>Fluid</strong> <strong>Mechanics</strong>,<br />

238:325–336, 1992.<br />

[14] B.E. Launder <strong>and</strong> D.B. Spalding. The numerical computation of turbulent<br />

flows. Computer Methods in Applied Mech. <strong>and</strong> Eng., 3:269–289, 1974.<br />

[15] W.P. Jones <strong>and</strong> B.E. Launder. The prediction of laminarization with a twoequation<br />

model of turbulence. Int. J. Heat Mass Transfer, 15:301–314, 1972.<br />

[16] D.C. Wilcox. Turbulence modeling for CFD. DCW Industries, Inc., 1993.<br />

[17] Barry Spain. Tensor Calculus. Oliver <strong>and</strong> Boyd, 1965.<br />

[18] J.L. Synge <strong>and</strong> A. Schild. Tensor Calculus. Dover Publications, 1969.<br />

[19] P. Morse <strong>and</strong> H. Feshbach. Methods of Theoretical Physics. McGraw-Hill,<br />

New York, 1953.


Appendix A<br />

Introduction to Tensor Calculus<br />

115


"<br />

¦<br />

-<br />

¦<br />

¦<br />

¡<br />

¦<br />

ã<br />

<br />

<br />

¦<br />

<br />

<br />

T<br />

1<br />

¦<br />

¦<br />

<br />

<br />

¦<br />

¦<br />

<br />

<br />

4<br />

<br />

¦<br />

<br />

<br />

l<br />

116 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

There are two aspects of tensors that are of practical <strong>and</strong> fundamental importance:<br />

tensor notation <strong>and</strong> tensor invariance. Tensor notation is of great practical<br />

importance, since it simplifies h<strong>and</strong>ling of complex equation systems. The<br />

idea of tensor invariance is of both practical <strong>and</strong> fundamental importance, since it<br />

provides a powerful apparatus to describe non-Euclidean spaces in general <strong>and</strong><br />

curvilinear coordinate systems in particular.<br />

A definition of a tensor is given in Section A.1. Section A.2 deals with an<br />

important class of Cartesian tensors, <strong>and</strong> describes the rules of tensor notation.<br />

Section A.3 provides a brief introduction to general curvilinear coordinates, invariant<br />

forms <strong>and</strong> the rules of covariant differentiation.<br />

A.1 Coordinates <strong>and</strong> Tensors<br />

Consider a space of real numbers of dimension = , A B , <strong>and</strong> a single real time,<br />

. Continuum properties in this space can be described by arrays of different<br />

dimensions, , such as scalars @ ( ), vectors ( ), matrices ( ), <strong>and</strong><br />

general multi-dimensional arrays. In this space we shall introduce a coordinate<br />

system, , as a way of assigning real numbers 1 for every point of space<br />

¦YF„½º½B<br />

<br />

=<br />

There can be a variety of possible coordinate systems. A general transformation<br />

rule between the coordinate systems is<br />

%<br />

(A.1)<br />

Consider a small R <br />

displacement . Then it can be transformed from coordinate<br />

system to a new coordinate ã system using the partial differentiation<br />

rules applied to (A.1):<br />

ã<br />

ã<br />

F IOI/ B<br />

(A.2)<br />

ã R<br />

R <br />

Y ã<br />

This transformation rule 2 can be generalized to a set of vectors that we shall call<br />

contravariant vectors:<br />

Y <br />

(A.3)<br />

ã Y ¡<br />

1<br />

Y <br />

Super-indexes denote components of a vector (¾†¿¸ÀÁ4ÁÂ) <strong>and</strong><br />

not the power exponent, for the reason<br />

explained later (Definition A.1.1)<br />

2 The repeated indexes imply summation (See. Proposition A.21)


<br />

Y<br />

ã Y<br />

Y ¦<br />

ã Y<br />

¦<br />

¡<br />

£<br />

|<br />

¦<br />

<br />

<br />

<br />

<br />

¦<br />

<br />

<br />

<br />

<br />

Y ¦<br />

Y<br />

<br />

A.1. COORDINATES AND TENSORS 117<br />

That is, a contravariant vector is defined as a vector which transforms to a new<br />

coordinate system according to (A.3). We can also introduce the transformation<br />

matrix as:<br />

(A.4)<br />

¦<br />

¢ <br />

Y ã<br />

With which (A.3) can be rewritten as:<br />

Y <br />

(A.5)<br />

¦<br />

¡ <br />

£<br />

Transformation rule (A.3) will not apply to all the vectors in our space. For<br />

will transform as:<br />

example, a partial derivative Y : Y <br />

(A.6)<br />

¦ Y <br />

<br />

Y <br />

Y <br />

that is, the transformation coefficients are the other way up compared to (A.2).<br />

Now we can generalize this transformation rule, so that each vector that transforms<br />

according to (A.6) will be called a Covariant vector:<br />

Y ã<br />

(A.7)<br />

Y <br />

¦ ¡<br />

ã ¡©¦<br />

This provides the reason for using lower <strong>and</strong> upper indexes in a general<br />

tensor notation.<br />

Y ã<br />

Definition A.1.1 Tensor<br />

Tensor of order is a set of =<br />

numbers identified by<br />

integer indexes.<br />

For example, a 3rd order tensor ¡ can be denoted as ¡§¦/ j <strong>and</strong> an -order tensor<br />

can be denoted as ¡§¦6½º½¦ÄÃ.Each index of a tensor changes between 1 <strong>and</strong><br />

n. For example, in a 3-dimensional space (n=3) a second order tensor will be<br />

represented À Ô by components.<br />

Each index of a tensor should comply to one of the two transformation rules:<br />

(A.3) or (A.7). An index that complies to the rule (A.7) is called a covariant index<br />

<strong>and</strong> is denoted as a sub-index, <strong>and</strong> an index complying to the transformation rule<br />

(A.3) is called a contravariant index <strong>and</strong> is denoted as a super-index.


¡©¦6½º½¦Ã ¡§¦6½º½¦Ã“1<br />

<br />

¨<br />

¦<br />

¦<br />

@<br />

¦<br />

¦<br />

B<br />

118 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

Each index of a tensor can be covariant or a contravariant, thus tensor<br />

j ¦/<br />

is a 2-covariant, 1-contravariant tensor of third order.<br />

¡<br />

Tensors are usually functions of space <strong>and</strong> time:<br />

which defines a tensor field, i.e. for every point <br />

<strong>and</strong> time " there are a set of<br />

F IOI/ B '<br />

"54<br />

nubers ¡©¦6½º½¦Ã.<br />

Remark A.1.2 Tensor character of coordinate vectors<br />

Note, that the <br />

coordinates are not tensors, since generally, they are not<br />

transformed as (A.5). Transformation law for the coordinates is actually given<br />

by (A.1). Nevertheless, we shall use the upper (contravariant) indexes for the<br />

coordinates.<br />

Definition A.1.3 Kronecker delta tensor<br />

Second order delta tensor, ß ¦/<br />

is defined as<br />

(A.8)<br />

#­<br />

¦7Œ<br />

¦7Œ T<br />

ß<br />

From this definition <strong>and</strong> since coordinates <br />

v ¨<br />

it follows that:<br />

are independent of each other<br />

#ª<br />

v ß<br />

(A.9)<br />

<br />

Y <br />

ß ¦/<br />

Y <br />

Corollary A.1.4 Delta product<br />

that<br />

From the definition (A.1.3) <strong>and</strong> the summation convention (A.21), follows<br />

(A.10)<br />

ß ¦/ ¡k¡§¦


£<br />

<br />

¦<br />

A<br />

j<br />

¦<br />

<br />

¦<br />

¦<br />

<br />

<br />

<br />

j<br />

<br />

<br />

<br />

¦<br />

<br />

<br />

¦<br />

<br />

j<br />

<br />

¦<br />

j<br />

¦<br />

<br />

j<br />

j<br />

<br />

A.2. CARTESIAN TENSORS 119<br />

<br />

A¦<br />

:<br />

Assume that there exists the transformation inverse to (A.5), which we call<br />

(A.11)<br />

<br />

A¦<br />

Then by analogy (A.4)A¦<br />

<br />

to<br />

can be defined as:<br />

ã R<br />

R <br />

A¦<br />

<br />

(A.12)<br />

Y <br />

£<br />

A¦<br />

<br />

j<br />

j , namely:<br />

Y ã<br />

¦<br />

From this relation <strong>and</strong> the independence of coordinates (A.9) it follows<br />

A<br />

that £<br />

ß ¦<br />

Y <br />

(A.13)<br />

Y ã<br />

Y <br />

ã j Y<br />

Y <br />

ß ¦<br />

Y ã<br />

Y ã<br />

Y <br />

Y ã<br />

j<br />

j<br />

Y ã<br />

A.2 Cartesian Tensors<br />

Cartesian tensors are a sub-set of general tensors for which the transformation<br />

matrix (A.4) satisfies the following relation:<br />

(A.14)<br />

¦ Y ã<br />

<br />

j ¦ £ j £<br />

Y ã<br />

ß ¦/<br />

For Cartesian tensors we have<br />

Y <br />

Y <br />

(A.15)<br />

ïY <br />

Y ã<br />

Y ã<br />

(see Problem A.4.3), which means that both (A.5) <strong>and</strong> (A.6) are transformed ¦<br />

with<br />

the same j matrix . This in turn means that the difference between the covariant<br />

<strong>and</strong> contravariant indexes vanishes for the Cartesian tensors. Considering this<br />

we shall only use the sub-indexes whenever we deal with Cartesian tensors.<br />

£<br />

Y j


j<br />

j<br />

„<br />

æ<br />

¤<br />

„<br />

P<br />

æ<br />

¤<br />

120 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

A.2.1<br />

Tensor Notation<br />

Tensor notation simplifies writing complex equations involving multi-dimensional<br />

objects. This notation is based on a set of tensor rules. The rules introduced<br />

in this section represent a complete set of rules for Cartesian tensors <strong>and</strong> will<br />

be extended in the case of general tensors (Sec.A.3). The importance of tensor<br />

rules is given by the following general remark:<br />

Remark A.2.1 Tensor rules Tensor rules guarantee that if an expression follows<br />

these rules it represents a tensor according to Definition A.1.1.<br />

Thus, following tensor rules, one can build tensor expressions that will preserve<br />

tensor properties of coordinate transformations (Definition A.1.1) <strong>and</strong> coordinate<br />

invariance (Section A.3).<br />

Tensor rules are based on the following definitions <strong>and</strong> propositions.<br />

Definition A.2.2 Tensor terms<br />

A tensor term is a product of tensors.<br />

For example:<br />

(A.16)<br />

¡§¦/<br />

¤“<br />

jÙÄ<br />

Pæ<br />

Definition A.2.3 Tensor expression<br />

Tensor expression is a sum of tensor terms. For example:<br />

(A.17)<br />

P<br />

Generally the terms in the expression may come with plus or minus sign.<br />

¡§¦/<br />

¤“<br />

¦ ½<br />

jê>\Ä<br />

Pæ<br />

Proposition A.2.4 Allowed operations<br />

The only allowed algebraic operations in tensor expressions are the addition,<br />

subtraction <strong>and</strong> multiplication. Divisions are only allowed for constants, like<br />

: Ä . If a tensor index appears in a denominator, such term should be redefined,<br />

@<br />

so as not to have tensor indexes in a denominator. For @ : ¡t¦<br />

example, should be<br />

redefined as: @ : ¡©¦<br />

.<br />

¤§¦¢


Ä<br />

¦ ½<br />

j<br />

P<br />

Ä<br />

P<br />

<br />

Ä<br />

<br />

j<br />

j<br />

A.2. CARTESIAN TENSORS 121<br />

Definition A.2.5 Tensor equality<br />

Tensor equality is an equality of two tensor expressions.<br />

For example:<br />

(A.18)<br />

¡§¦/Ù¤“k<br />

¦f¤<br />

jÙ„<br />

>^„<br />

Definition A.2.6 Free indexes<br />

A free index is any index that occurs only once in a tensor term. For example,<br />

index # is a free index in the term (A.16).<br />

Proposition A.2.7 Free index restriction<br />

Every term in a tensor equality should have the same set of free indexes.<br />

For example, if index # is a free index in any term of tensor equality, such as<br />

(A.18), it should be the free index in all other terms. For example<br />

¡©¦/¤k<br />

½ <br />

is not a valid tensor equality since index # is a free index in the term on the<br />

RHS but not in the LHS.<br />

Definition A.2.8 Rank of a term<br />

A rank of a tensor term is equal to the number of its free indexes.<br />

¤“<br />

For example, the rank of the j Äj term is equal to 1.<br />

It follows from (A.2.7) that ranks of all the terms in a valid tensor ¡§¦/ expression<br />

should be the same. Note, that the difference between the order <strong>and</strong> the rank is<br />

that the order is equal to the number of indexes of a tensor, <strong>and</strong> the rank is equal<br />

to the number of free indexes in a tensor term.<br />

Proposition A.2.9 Renaming of free indexes<br />

Any free index in a tensor expression can be named by any symbol as long<br />

as this symbol does not already occur in the tensor expression.


°<br />

¡<br />

j<br />

Ä<br />

„<br />

„<br />

<br />

<br />

122 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

For example, the equality<br />

(A.19)<br />

is equivalent to<br />

¡§¦7Ù¤<br />

¦ ½ <br />

(A.20)<br />

Ù¤ ½ <br />

Äj<br />

Here we replaced the free index # with ° .<br />

Definition A.2.10 Dummy indexes<br />

A dummy index is any index that occurs twice in a tensor term.<br />

For example, indexes¨‰'<br />

'€Xò'qŠ in (A.16) are dummy indexes.<br />

Proposition A.2.11 Summation rule<br />

Any dummy index implies summation, i.e.<br />

(A.21)<br />

B w¦<br />

¡§¦M¤§¦$<br />

¡§¦³¤§¦<br />

Proposition A.2.12 Summation rule exception If there should be no summation<br />

over the repeated indices, it can be indicated by enclosing such indices in parentheses.<br />

For example, expression:<br />

½ ¦/<br />

does not imply summation over # .<br />

Ä0¦32¡0¦32¤“k<br />

Corollary A.2.13 Scalar product<br />

A scalar product notation from vector algebra: 1 ¡ûž¤ 4<br />

notation as ¡§¦M¤§¦ .<br />

is expressed in tensor


¡§¦<br />

æ P<br />

j<br />

¤<br />

P<br />

Ä<br />

æ<br />

¡<br />

¤<br />

j<br />

P<br />

P<br />

<br />

Ä<br />

„<br />

„<br />

<br />

<br />

A.2. CARTESIAN TENSORS 123<br />

The scalar product operation is also called a contraction of indexes.<br />

Proposition A.2.14 Dummy index restriction<br />

No index can occur more than twice in any tensor term.<br />

Remark A.2.15 Repeated indexes<br />

In case if an index occurs more than twice in a term this term should be<br />

redefined so as not to contain more than two occurrences of the same index. For<br />

is defined as<br />

j<br />

¤<br />

½ <br />

example, term should be rewritten j j as , where jèÄj j<br />

½ ¡t¦ ¡§¦<br />

½ <br />

¢¤ 0j2Ä0j2with no summation over ° in the last term.<br />

Proposition A.2.16 Renaming of dummy indexes<br />

Any dummy index in a tensor term can be renamed to any symbol as long<br />

as this symbol does not already occur in this term.<br />

For example, term is equivalent to , <strong>and</strong> so are terms ¡§¦/ j<br />

.<br />

¡Ù¤“ ¡§¦M¤§¦<br />

¤<br />

Äj <strong>and</strong><br />

Remark A.2.17 Renaming rules<br />

Note that while the dummy index renaming rule (A.2.16) is applied to each<br />

tensor term separately, the free index naming rule (A.2.9) should apply to the<br />

whole tensor expression. For example, the equality (A.19) above<br />

¡©¦/¤k<br />

¦ ½ <br />

can also be rewritten as<br />

(A.22)<br />

½ <br />

without changing its meaning.<br />

Äj<br />

(See Problem A.4.1).<br />

Definition A.2.18 Permutation tensor<br />

The components of a third order permutation tensor pq¦/ j are defined to be<br />

equal to 0 when any index is equal to any other index; equal to 1 when the set of


% £A<br />

£<br />

£A<br />

4<br />

(A.25)<br />

~<br />

j<br />

l<br />

À @<br />

l 4Pv<br />

<br />

j<br />

¡¼ ¤ Í ~ ~<br />

j<br />

j<br />

j<br />

j<br />

<br />

<br />

<br />

`<br />

T<br />

@<br />

@<br />

£A<br />

124 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

indexes can be obtained by cyclic permutation of 123; <strong>and</strong> -1 when the indexes<br />

can be obtained by cyclic permutation from 132. In a mathematical language it<br />

can be expressed as:<br />

#_<br />

¨#Å<br />

°Å%¨<br />

°Æv p5¦7<br />

(A.23)<br />

#¨°ËÇ º9É<br />

1<br />

p5¦/<br />

#­<br />

#¨°ÈÇ º9É<br />

1<br />

4Êv p5¦7<br />

where º9É<br />

–<br />

is a permutation group of a triple of indexes abc, i.e. º9É<br />

£AE- . For example, the permutation group of 123 will consist of three<br />

–<br />

4 <br />

@EÀ<br />

combinations: 123, 231 <strong>and</strong> 312, <strong>and</strong> the permutation group of 123 consists of<br />

132, 321 <strong>and</strong> 213.<br />

–§'BA<br />

–<br />

',–<br />

1<br />

1<br />

Corollary A.2.19 Permutation of the permutation tensor indexes<br />

From the definition of the permutation tensor it follows that the permutation<br />

of any of its two indexes changes its sign:<br />

(A.24)<br />

p5¦/<br />

` p5¦<br />

A tensor with this property is called skew-symmetric.<br />

Corollary A.2.20 Vector product<br />

A vector product (cross-product) of two vectors in vector notation is expressed<br />

as<br />

Ä<br />

which in tensor notation can be expressed as<br />

(A.26)<br />

¡§¦$£p5¦/<br />

¤<br />

Äj<br />

Remark A.2.21 Cross product<br />

Tensor expression (A.26) is more accurate than its vector counterpart (A.25),<br />

since it explicitly shows how to compute each component of a vector product.


j<br />

j<br />

p5¦<br />

æ P<br />

j<br />

j<br />

j<br />

¡<br />

<br />

j<br />

j<br />

j<br />

j<br />

j<br />

<br />

<br />

P<br />

j<br />

ß<br />

j<br />

<br />

æ<br />

j<br />

j<br />

T<br />

j<br />

j<br />

¡<br />

<br />

j<br />

j<br />

æ<br />

j<br />

ß<br />

j<br />

P<br />

A.2. CARTESIAN TENSORS 125<br />

Theorem A.2.22 Symmetric identity<br />

If ¡§¦/ is a symmetric tensor, then the following identity is true:<br />

(A.27)<br />

p5¦/<br />

¡“<br />

Proof:<br />

From the symmetry of ¡§¦/ we have:<br />

(A.28)<br />

p5¦/<br />

¡“<br />

£pm¦/<br />

Let’s rename index¨into ° <strong>and</strong> ° into¨in the RHS of this expression, according<br />

to rule (A.2.16):<br />

p5¦/<br />

k£pm¦<br />

¡<br />

Using (A.24) we finally obtain:<br />

p5¦<br />

,¡<br />

` p5¦/<br />

¡<br />

j<br />

Comparing the RHS of this expression to the LHS of (A.28) we have:<br />

p5¦/<br />

¡<br />

` p5¦/<br />

¡<br />

from which we conclude that (A.27) is true.<br />

Theorem A.2.23 Tensor identity<br />

The following tensor identity is true:<br />

(A.29)<br />

pq¦7<br />

ß <br />

`]ß <br />

Proof<br />

This identity can be proved by examining the components of equality (A.29)<br />

component-by-component.


(A.30)<br />

~<br />

Í ¤ Í ¡<br />

Ä~<br />

1~<br />

¤ ¡‡ž 4<br />

Ä~<br />

1~ ~<br />

<br />

Y<br />

¡<br />

"<br />

¡<br />

¦<br />

4Û`<br />

Ä~<br />

¡‡ž ¤ 4 ~ 1~<br />

126 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

Corollary A.2.24 Vector identity<br />

Using the tensor identity (A.29) it is possible to prove the following important<br />

vector identity:<br />

See Problem A.4.4.<br />

A.2.2<br />

Tensor Derivatives<br />

For Cartesian tensors derivatives introduce the following notation.<br />

Definition A.2.25 Time derivative of a tensor<br />

A partial derivative of a tensor over time is designated as<br />

¡£¢ Y<br />

Definition A.2.26 Spatial derivative of a tensor<br />

A partial derivative of a tensor ¡<br />

over one or its spacial components is denoted<br />

as ¡ ¦ :<br />

(A.31)<br />

¡ ¦¢Y<br />

that is, the index of the spatial component that the derivation is done over is<br />

delimited by a comma (’,’) from other indexes. For example, ¡t¦/y j is a derivative of<br />

a second order tensor ¡§¦7 .<br />

Y <br />

Definition A.2.27 Nabla<br />

Nabla operator acting on a tensor ¡<br />

is defined as<br />

(A.32)<br />

¾ ¦f¡¢¡ƒ ¦


F<br />

<br />

F<br />

<br />

<br />

<br />

À<br />

A.2. CARTESIAN TENSORS 127<br />

Even though the notation in (A.31) is sufficient to define the derivative, In<br />

some instances it is convenient to introduce the nabla operator as defined above.<br />

Remark A.2.28 Tensor derivative<br />

In a more general context of non-Cartesian tensors the coordinate independent<br />

derivative will have a different form from (A.31). See the chapter on covariant<br />

differentiation in [17].<br />

Remark A.2.29 Rank of a tensor derivative<br />

The derivative of a zero order tensor (scalar) as given by (A.31) forms a<br />

first order tensor (vector). Generally, the derivative of an -order tensor forms an<br />

order tensor. However, if the derivation index is a dummy index, then the<br />

>×@<br />

rank of the derivative will be lower than that of the original tensor. For example,<br />

the rank of the derivative is one, since there is only one free index in this<br />

term.<br />

¡§¦/m<br />

Remark A.2.30 Gradient<br />

Expression (A.31) represents a gradient, which in a vector notation is ¾ ¡<br />

:<br />

¾ ¡ `åb ¡ƒ ¦<br />

Corollary A.2.31 Derivative of a coordinate<br />

From (A.9) it follows that:<br />

(A.33)<br />

¦Kk<br />

ß ¦/<br />

In particular, the following identity is true:<br />

(A.34)<br />

<br />

<br />

¦f ¦$<br />

rq r<br />

>]<br />

>] <br />

@•>¼@•>@<br />

Remark A.2.32 Divergence operator<br />

A divergence operator in a vector notation is represented in a tensor notation<br />

as ¡©¦f ¦ :<br />

¾ 1 ¡ 4]`$b ¡©¦f ¦ ž~


<br />

r<br />

› R (A.35)<br />

¦<br />

¦<br />

[<br />

¦<br />

<br />

128 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

Remark A.2.33 Laplace operator<br />

¡ƒ ¦ì¦<br />

The Laplace operator in vector notation is represented in tensor notation as<br />

:<br />

¡ `åb ¡ƒ ¦ì¦<br />

Remark A.2.34 Tensor notation<br />

Examples (A.2.30), (A.2.32) <strong>and</strong> (A.2.33) clearly show that tensor notation<br />

is more concise <strong>and</strong> accurate than vector notation, since it explicitly shows how<br />

each component should be computed. It is also more general since it covers<br />

.<br />

cases that don’t have representation in vector notation, for example: ¡C¦ j<br />

<br />

j<br />

A.3 Curvilinear coordinates<br />

In this section 3 we introduce the idea of tensor invariance <strong>and</strong> introduce the rules<br />

for constructing invariant forms.<br />

A.3.1<br />

Tensor invariance<br />

The distance between the material points in a Cartesian coordinate system is<br />

. The metric tensor, is introduced to generalize the<br />

notion of distance (A.39) to curvilinear coordinates.<br />


› R (A.36)<br />

r<br />

› R (A.39)<br />

¦<br />

<br />

¦<br />

¦<br />

<br />

¦<br />

¦<br />

<br />

<br />

<br />

ã<br />

¦<br />

¦<br />

¦<br />

j<br />

R ã <br />

¦<br />

¦<br />

<br />

<br />

j<br />

<br />

<br />

<br />

j<br />

¦<br />

<br />

<br />

¦<br />

ã<br />

<br />

<br />

j<br />

¦<br />

<br />

<br />

¦<br />

<br />

¦<br />

¦<br />

<br />

¦<br />

<br />

<br />

¦<br />

<br />

<br />

<br />

<br />

T<br />

j<br />

A.3. CURVILINEAR COORDINATES 129<br />

systems, that is, the distance should be independent of the coordinate system,<br />

thus:<br />

¦/<br />

¦/<br />

The metric tensor is symmetric, which can be shown by rewriting (A.35) as<br />

follows:<br />

R <br />

R <br />

ã R<br />

ã R<br />

¦/<br />

¼


¡<br />

ã<br />

¡ X<br />

¦<br />

<br />

<br />

¦<br />

<br />

¦<br />

<br />

<br />

130 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

Using (A.38) we can also find its inverse as:<br />

(A.41)<br />

¦7<br />

Y ã<br />

Y ã<br />

in various curvilinear coordi-<br />

Using these expression one can compute <strong>and</strong><br />

nate systems (see Problem A.4.6).<br />

J¦7<br />

¦7<br />

Y j<br />

Y j<br />

Definition A.3.2 Conjugate tensors<br />

For each index of a tensor we introduce the conjugate tensor where this<br />

index is transfered to its counterpart (covariant/contravariant) using the relations:<br />

(A.42)<br />

(A.43)<br />

¼<br />

¡“<br />

¦/<br />

¡©¦‡


j<br />

¦<br />

)<br />

¦<br />

)<br />

j<br />

)<br />

¤<br />

<br />

¦ <br />

> 9 6 ¦<br />

j ;<br />

¦ <br />

> 9 6 j ¦f<br />

j ;<br />

<br />

)<br />

<br />

A.3. CURVILINEAR COORDINATES 131<br />

Definition A.3.4 Invariant Scalar Product<br />

The invariant form of the scalar product between ¦7<br />

two<br />

¡©¦M¤<br />

covariant vectors<br />

<strong>and</strong> is . Similarly, the invariant form of a scalar product between ¡t¦<br />

¦ ¦ ¦<br />

two<br />

contravariant vectors <strong>and</strong> is , where is the metric tensor (A.40)<br />

<strong>and</strong> is its conjugate (A.38).<br />

\)<br />

where the rising of indexes was done using relation ) (A.42):<br />

.<br />

6 ¦/<br />

, <strong>and</strong> j ¦ 6<br />

j ¥ j


¡<br />

<br />

¦<br />

Y<br />

¡<br />

¦<br />

Y<br />

<br />

¡<br />

¦<br />

><br />

¡<br />

y͡ j<br />

Y<br />

|<br />

¨<br />

¦<br />

j<br />

|<br />

132 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

A.3.2<br />

Covariant differentiation<br />

A simple scalar value, , is invariant under coordinate transformations. A partial<br />

derivative of an invariant is a first order covariant tensor (vector):<br />

¨<br />

¨ ¦G˜Y<br />

However, a partial derivative of a tensor of the order one <strong>and</strong> greater is not<br />

generally an invariant under coordinate transformations of type (A.7) <strong>and</strong> (A.3).<br />

In curvilinear coordinate system we should use more complex differentiation<br />

rules to preserve the invariance of the derivative. These rules are called the rules<br />

of covariant differentiation <strong>and</strong> they guarantee that the derivative itself is a tensor.<br />

According to these rules the derivatives for covariant <strong>and</strong> contravariant indices<br />

will be slightly different. They are expressed as follows:<br />

Y <br />

(A.46)<br />

¡§¦<br />

(A.47)<br />

><br />

̦<br />

j<br />

<br />

¡§¦f¢<br />

Y <br />

`Ìj ¦/yÍ¡ <br />

¢SY<br />

Y <br />

where contstructÌj<br />

the<br />

¦/£Íis defined as<br />


><br />

¡<br />

%<br />

%<br />

æ<br />

¦6P<br />

¡<br />

æ<br />

¡<br />

-<br />

-<br />

<br />

¦<br />

)<br />

¦<br />

¡<br />

j<br />

)<br />

¡<br />

¦<br />

R<br />

%<br />

¦6½º½¦Î§<br />

P<br />

<br />

`<br />

><br />

><br />

<br />

%<br />

%<br />

¦<br />

º<br />

j<br />

æ<br />

<br />

Y<br />

P Y<br />

-<br />

-<br />

><br />

¡<br />

¡<br />

%<br />

¦6½º½¦Î ¡<br />

6½º½ÃˆÏ6æ<br />

j<br />

¦<br />

j<br />

j<br />

¡<br />

¦<br />

|<br />

A.3. CURVILINEAR COORDINATES 133<br />

the contravariant second order tensor ¡ ¦/<br />

we have:<br />

(A.48)<br />

¦/<br />

<br />

j<br />

¦<br />

- |aj<br />

<br />

- |­j<br />

¦/<br />

üY<br />

¡ |<br />

Y j ><br />

And for a general = -covariant,<br />

-contravariant tensor we have:<br />

žEžEž<br />

(A.49)<br />

¦6½º½¦Î<br />

><br />

Despite their seeming complexity, the relations of covariant differentiation<br />

can be easily implemented algorithmically <strong>and</strong> used in numerical solutions on<br />

arbitrary curved computational grids (Problem A.4.8).<br />

æ<br />

6½º½Ã<br />

¦½º½¦ÄÎ><br />

¦ÎP<br />

¦6½º½¦ÎÏ6æ<br />

6<br />

½º½Ã<br />

Ã<br />

æfP<br />

žEžEž<br />

6½º½Ã<br />

æfP<br />

6½º½Ã<br />

6½º½Ã ¦6½º½¦Î<br />

Remark A.3.7 Rules of invariant expressions<br />

As was pointed out in Corollary A.3.6, the rules to build invariant expressions<br />

involve raising or lowering indexes (A.42), (A.43). However, since we did not<br />

introduce the notation for contravariant derivative, the only way to raise the index<br />

.<br />

of a covariant derivative, say ¡ ¦ , it to use the relation (A.42) directly, that is:<br />

¦7<br />

¡ƒ<br />

For example, we can re-formulate the momentum equation (A.45) in terms<br />

of contravariant free index # as:<br />

(A.50)<br />

¦<br />

<br />

j<br />

<br />

j<br />

>^)<br />

; > 9 6<br />

where the index of the pressure term was raised by means of (A.42).<br />

Using the invariance of the scalar product one can<br />

#<br />

construct<br />

¡¼¢¥¡<br />

two ¦ important<br />

differential operators in curvilinear coordinates: R<br />

¦<br />

divergence of a vector<br />

(A.51) [<br />

<strong>and</strong> Laplacian, (A.55).<br />

¡£¢£<br />

¡<br />

j<br />

j<br />

Definition A.3.8 Divergence<br />

¦<br />

Divergence of a vector is defined<br />

¦<br />

as :<br />

¡<br />

(A.51)<br />

¦<br />

¦ <br />

# ¡£¢¡


(A.55) [<br />

¡<br />

Y<br />

¡<br />

¡<br />

¦<br />

> ¦<br />

¡<br />

<br />

Y<br />

¡<br />

@<br />

m<br />

m<br />

¦<br />

> ¦<br />

Y<br />

Y<br />

¦<br />

%<br />

-<br />

¦<br />

<br />

¡<br />

¦<br />

134 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

From this definition <strong>and</strong> the rule of covariant differentiation (A.47) we have:<br />

(A.52)<br />

¦<br />

¦ <br />

¦<br />

j<br />

¡ j<br />

this can be shown [18] to be equal to:<br />

Y <br />

¦<br />

¦ <br />

— @<br />

Y <br />

(A.53)<br />

Q¡<br />

where is the determinant of the metric tensor .<br />


(A.61) ’<br />

F<br />

r<br />

F<br />

4<br />

><br />

£<br />

–<br />

F<br />

¦<br />

¦<br />

£ T<br />

> F<br />

£<br />

><br />

rAF<br />

F<br />

F<br />

<br />

F<br />

¦<br />

T<br />

¦<br />

T T<br />

-<br />

'BAr '<br />

T<br />

'<br />

T<br />

'<br />

-<br />

<br />

-<br />

<br />

><br />

<br />

F<br />

<br />

<br />

<br />

T<br />

4<br />

A.3. CURVILINEAR COORDINATES 135<br />

¦<br />

Consider three unit vectors, , each directed along one of the coordinate<br />

axis (tangential unit vectors), that is:<br />

'BA¦<br />

'*–<br />

£<br />

(A.56)<br />

(A.57) A¦<br />

(A.58)<br />

T ×%<br />

T Ú%<br />

ä% £<br />

F '<br />

The condition of orthogonality means that the scalar product between any<br />

two of these unit vectors should be zero. According to the definition of a scalar<br />

product (Definition A.3.4) it should be written in form (A.44), that is, a scalar product<br />

between vectors <strong>and</strong>A¦<br />

can be written as: £ ¦A¦<br />

or £‰¦A¦<br />

. Let’s use the first<br />

form for definiteness. Then, applying the operation of rising indexes (A.42), we<br />

can express the scalar product in contravariant components only:<br />

£Ž¦<br />

',–<br />

¦A<br />

T ¼£<br />

r £<br />

T‰T<br />

FðF<br />


› R (A.66)<br />

¦<br />

¦<br />

£<br />

£<br />

¦<br />

¦<br />

£ £<br />

F F<br />

£<br />

–<br />

¦<br />

¦<br />

£<br />

@ Ú%<br />

’<br />

¦<br />

F<br />

¦<br />

<br />

F<br />

'<br />

'<br />

’<br />

@<br />

’<br />

T<br />

'<br />

F<br />

r<br />

4<br />

@<br />

’<br />

¯<br />

'<br />

T<br />

<br />

'<br />

@<br />

’<br />

<br />

–<br />

T<br />

T<br />

¦<br />

–<br />

-<br />

-<br />

-<br />

’<br />

F<br />

r<br />

4<br />

–<br />

¦ <br />

–<br />

r<br />

1<br />

R ã <br />

¦<br />

<br />

<br />

@<br />

¦ r<br />

4<br />

@<br />

136 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

£‰¦$<br />

¦<br />

@<br />

Or, expressed in contravariant components only the condition of unity is:<br />

A¦A¦$<br />


F<br />

¦<br />

¡<br />

F<br />

’<br />

¡<br />

F<br />

ˆ<br />

F<br />

'<br />

ˆ<br />

@<br />

<br />

ˆ<br />

¦<br />

r<br />

r<br />

¦<br />

F<br />

'<br />

F<br />

¡<br />

ˆ<br />

<br />

’0¦32’0¦32<br />

¦<br />

¦<br />

ˆ<br />

¦<br />

¦<br />

T<br />

'<br />

T<br />

A.3. CURVILINEAR COORDINATES 137<br />

Combining the latter with (A.38), we obtain: ß ¦/<br />

that<br />

ß ¦/<br />

, from which it follows<br />

(A.68)<br />

’0¦32<br />

@ : ’0¦32<br />

Physical components of tensors<br />

Consider a direction in space determined by a unit ˆ vector . Then the physical<br />

is given by a scalar product between<br />

ˆ <strong>and</strong> (Definition A.3.4), namely:<br />

component of a vector ¡§¦ in the direction ˆ<br />

¡§¦<br />

¦/<br />

¡ 1<br />

4 ¼<br />

¡©¦<br />

According to Corollary A.3.5 the above can also be rewritten as:<br />

(A.69)<br />

¡ 1<br />

4 ¼¡§¦<br />

¡<br />

Suppose the unit vector is directed along one of the ˆ axis:<br />

From (A.63) it follows that:<br />

%<br />

ˆ


F<br />

r<br />

[<br />

¡<br />

@ Y ¡£<br />

Y<br />

F<br />

<br />

<br />

¦<br />

—<br />

¦<br />

@<br />

<br />

—<br />

¦<br />

’0¦32¡§¦<br />

¡<br />

Y ’0¦32<br />

Î<br />

›<br />

B ¢<br />

ó¦YF<br />

¦<br />

<br />

’<br />

¦<br />

138 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

General rules of covariant differentiation introduced in (Sec.A.3.2) simplify<br />

considerably in orthogonal coordinate systems. In particular, we can define the<br />

nabla operator by the physical components of a covariant vector composed of<br />

partial differentials:<br />

(A.71)<br />

’0¦32<br />

Y<br />

¾ ¦$<br />

Y <br />

where the parentheses indicate that there’s no summation with respect to index # .<br />

In orthogonal coordinate system the general expressions for divergence<br />

(A.53) <strong>and</strong> Laplacian (A.55)) operators can be expressed in terms of stretching<br />

factors only [19]:<br />

(A.72)<br />

¦<br />

¦ @ Y <br />

Y <br />

Important examples of orthogonal coordinate systems are spherical <strong>and</strong> cylindrical<br />

coordinate systems. Consider the example of a cylindrical coordinate system:<br />

',Î'W›ñ- :<br />

¦$ä%<br />

¦$ä%§¢<br />

'5<br />

- <strong>and</strong> ã 'q<br />

¼¢arcedWÎ<br />

rk¢deKÆ<br />

<br />

According to (A.40) only few components of the metric tensor will survive<br />

(Problem A.4.5). Then we can compute nabla, divergence <strong>and</strong> Laplacian operators<br />

according to (A.71), (A.52) <strong>and</strong> (A.55), or using simplified relations (A.72)-<br />

(A.73):


[<br />

R<br />

1<br />

Y<br />

r<br />

<br />

<br />

¡<br />

F<br />

<br />

¡<br />

Y<br />

Y<br />

r > 4<br />

¡<br />

r<br />

Y<br />

1<br />

Y<br />

> F<br />

¡<br />

@<br />

ã<br />

F<br />

> r<br />

x<br />

r<br />

— Y ¾<br />

Y<br />

@<br />

F ã<br />

1<br />

Y<br />

<br />

@<br />

r<br />

Y<br />

<br />

Y<br />

¡<br />

1<br />

r ><br />

> r<br />

¡<br />

r<br />

r > 4<br />

><br />

1<br />

Y<br />

<br />

Y<br />

Y<br />

¡<br />

r<br />

Y<br />

1<br />

Y<br />

Î Y<br />

¡<br />

<br />

r > 4<br />

¡<br />

r<br />

><br />

'<br />

Y<br />

+ Y<br />

@<br />

F ã<br />

r > 4<br />

@ ¡<br />

¢<br />

@<br />

ã<br />

¡<br />

F<br />

@<br />

¢<br />

x<br />

F<br />

Y<br />

¡<br />

F<br />

<br />

¡<br />

Y<br />

Y<br />

¢<br />

A.3. CURVILINEAR COORDINATES 139<br />

@<br />

¢<br />

¢ '<br />

¡tr<br />

# ¡£ Y<br />

Y ã<br />

><br />

F<br />

Y ã<br />

Y ã<br />

@<br />

¢<br />

¢ ><br />

Note, that instead of using the contravariant components as implied by the general<br />

definition of the divergence operator (A.51) we are using the covariant components<br />

as dictated by relation (A.70). The expression of the Laplacian becomes:<br />

Y Î<br />

¡F<br />

Y +<br />

¡H<br />

¡<br />

Y ã<br />

r 4<br />

Y ã<br />

Y ã<br />

ã Y<br />

¢ 4<br />

¢ r Y<br />

Y Î<br />

Y +<br />

(see Problems A.4.9,A.4.10).<br />

The advantages of the tensor approach are that it can be used for any type<br />

of curvilinear coordinate transformations, not necessarily analytically defined, like<br />

cylindrical (C.64) or spherical. Another advantage is that the equations above can<br />

be easily produced automatically using symbolic manipulation packages, such<br />

as Mathematica (wolfram.com) (Problems A.4.6,A.4.7,A.4.9). For further reading<br />

see [17, 18].


¤<br />

~<br />

„<br />

„<br />

j<br />

¦<br />

¤<br />

<br />

Ä j<br />

j<br />

j<br />

p5¦<br />

Pæ<br />

Ä<br />

P<br />

<br />

¤<br />

¤<br />

P<br />

P<br />

ß<br />

½<br />

j<br />

Ä<br />

æ<br />

æ<br />

½<br />

Ä<br />

j<br />

¤<br />

<br />

æ<br />

ß<br />

<br />

T<br />

j<br />

P<br />

4Û`<br />

Ä~<br />

¤<br />

j<br />

¦<br />

'<br />

j<br />

P<br />

<br />

j<br />

½<br />

'<br />

¡<br />

140 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

A.4 Problems<br />

Problem A.4.1 Check tensor expressions for consistency<br />

Check if the following Cartesian tensor expressions violate tensor rules:<br />

¡©¦/ ¤<br />

jê><br />

P æ<br />

½ <br />

P æ<br />

æÉæ<br />

jÛ><br />

¦/<br />

¡“Ù¤<br />

¦/¡©¦M¤Œ ¦/É<br />

`à½<br />

Problem A.4.2 Construct tensor expression<br />

¡t¦/ ¤§¦/ ½ ¦7<br />

j


^<br />

<br />

®<br />

<br />

®<br />

+<br />

+<br />

<br />

Î<br />

›<br />

ë<br />

A.4. PROBLEMS 141<br />

£¢a/cedWÎ<br />

£¢defÆ<br />

Obtain the components of the metric tensor (A.40) <strong>and</strong> its inverse<br />

(A.38) in cylindrical coordinates.<br />

J¦7<br />

¦/<br />

Problem A.4.6 Metric tensor in curvilinear coordinates<br />

Using Mathematica Compute the metric tensor,<br />

, (A.38) in spherical coordinate system ( ¢ ' ë_',Î ):<br />

, (A.40) <strong>and</strong> its conjugate,<br />

¼¢defÆ<br />

ÎŒarcedåë<br />

(A.73)<br />

ÎŒdefÆ<br />

¼¢a/cedåÎ<br />

¼¢defÆ<br />

Problem A.4.7 Christoffel’s symbols with Mathematica<br />

Using the Mathematica package, write the routines for computing Christoffel’s<br />

symbols.<br />

Problem A.4.8 Covariant differentiation with Mathematica<br />

Using the Mathematica package, <strong>and</strong> the routines developed in Problem A.4.7<br />

write the routines for covariant differentiation of tensors up to second order.<br />

Problem A.4.9 Divergence of a vector in curvilinear coordinates<br />

Using the Mathematica package <strong>and</strong> the solution of Problem A.4.8, write the<br />

routines for computing divergence of a vector in curvilinear coordinates.<br />

Problem A.4.10 Laplacian in curvilinear coordinates<br />

Using the Mathematica package <strong>and</strong> the solution of Problem A.4.8, write the<br />

routines for computing the Laplacian in curvilinear coordinates.<br />

Problem A.4.11 Invariant expressions


¡<br />

¤§¦<br />

æ P<br />

(A.76) „<br />

Ä<br />

j<br />

¤<br />

¦<br />

æ<br />

¦<br />

j<br />

`<br />

P<br />

><br />

¤<br />

j<br />

½<br />

Éj<br />

P<br />

Pj<br />

æ<br />

Ä<br />

<br />

æ<br />

<br />

<br />

½<br />

j<br />

<br />

<br />

Ä j æ<br />

¦ÉP<br />

¤<br />

P<br />

¦ì³<br />

æ<br />

142 APPENDIX A. INTRODUCTION TO TENSOR CALCULUS<br />

not:<br />

Check if any of these tensor expressions are invariant, <strong>and</strong> correct them if<br />

(A.74)<br />

¦<br />

<br />

Ä j<br />

<br />

³<br />

j<br />

½ <br />

¡§¦³¤<br />

(A.75)<br />

¦/<br />

<br />

j<br />

î¦<br />

j ¡<br />

<br />

<br />

¦<br />

Problem A.4.12 Contraction invariance<br />

¦<br />

¤©¦<br />

Prove that ¡<br />

is an invariant <strong>and</strong> ¡§¦M¤©¦ is not.


¾<br />

r<br />

)<br />

¾<br />

r<br />

r<br />

)<br />

Y<br />

<br />

<br />

®<br />

Y<br />

r<br />

> r<br />

4<br />

Y ) ¢<br />

¢<br />

Y<br />

Y<br />

r<br />

> r<br />

Y<br />

4<br />

Y<br />

r<br />

r<br />

r<br />

)<br />

r<br />

4<br />

Appendix B<br />

Curvilinear coordinate systems<br />

Here we will learn how to express the Laplacian<br />

in polar, cylindrical or spherical coordinates 1<br />

Y <br />

Y ®<br />

Y +<br />

Laplace equation in different coordinate systems. When solving boundary value<br />

problems in more than one dimension it is often necessary to use other coordinate<br />

systems than the Cartesian. It is then important to be able to express the<br />

Laplacian operator in these coordinate systems. We first consider<br />

POLAR COORDINATES<br />

¼¢a/cydWÎ<br />

¼¢defÆ<br />

Î<br />

Laplace’s equation in this coordinate system can be shown to be:<br />

1 ¢<br />

1 ¢<br />

1 ¢<br />

1 ¢<br />

4 Y<br />

',Î<br />

Y )<br />

',Î<br />

',Î<br />

@<br />

¢<br />

@ r Y<br />

¢<br />

',Î<br />

¢ r ><br />

¢ ><br />

Y Î<br />

PROOF; Let us for now apply the convention that subscript implies partial differentiation<br />

e.g.<br />

)œx<br />

1 This material is boroowed from http://www.physics.ubc.ca/ birger/n312l8/<br />

143


)<br />

4<br />

4 ¢ <br />

1<br />

1<br />

)GF4<br />

<br />

<strong>and</strong><br />

¢<br />

¾<br />

r<br />

)<br />

)<br />

)<br />

<br />

)<br />

<br />

<br />

r<br />

r<br />

Î<br />

Î<br />

<br />

<br />

<br />

<br />

¢<br />

<br />

)<br />

¢<br />

<br />

Î<br />

<br />

<br />

<br />

<br />

r<br />

<br />

<br />

<br />

<br />

¢ÑÐ<br />

Ò«B¬Æ é Î<br />

`<br />

<br />

¢<br />

¢<br />

<br />

> r<br />

4<br />

¢ <br />

Î<br />

<br />

r <br />

<br />

`<br />

@<br />

Ð<br />

1<br />

¯<br />

4<br />

<br />

®<br />

¢ )IF„F`<br />

Â<br />

r<br />

<br />

¢ )IF„F­><br />

Â<br />

Y<br />

r<br />

)<br />

Y<br />

<br />

®<br />

<br />

l<br />

¢<br />

4<br />

l<br />

l<br />

r<br />

<br />

F<br />

`<br />

1<br />

<br />

¢<br />

<br />

>^)GF„F1<br />

r <br />

<br />

r<br />

®<br />

<br />

®<br />

<br />

<br />

r<br />

Y<br />

r<br />

l<br />

Î<br />

`<br />

<br />

®<br />

Î<br />

r<br />

r<br />

<br />

®<br />

4<br />

4<br />

`<br />

l<br />

l<br />

W®<br />

¢<br />

Â<br />

W®<br />

¢<br />

Â<br />

r<br />

)<br />

)GF<br />

r<br />

4<br />

144 APPENDIX B. CURVILINEAR COORDINATE SYSTEMS<br />

Applying the chain rule we find<br />

)Wx<br />

>^)GF*Î<br />

1<br />

)GF4<br />

)œx<br />

>])œx<br />

>])GF,Î<br />

,<br />

,<br />

,<br />

1 ¢<br />

)œx<br />

)œxñx<br />

>^)œx„F*Î<br />

We have<br />

)IFñx<br />

>]®<br />

r <br />

<br />

¢<br />

>^®<br />

@<br />

¢<br />

¢ <br />

,<br />

¢ r ®<br />

r 1<br />

r 4 <br />

¢ r<br />

@•><br />

1 `<br />

Collecting terms<br />

W®<br />

¢<br />

Â<br />

4 ¢<br />

,<br />

¢ )Wx­><br />

W®<br />

)œx„F­> ¢<br />

Similarly we can show that<br />

,<br />

¢ r )œxñx_><br />

®<br />

W®<br />

)WxFa> ¢<br />

Again, collecting terms<br />

¢ )œx<br />

¯m¯<br />

¢ r )œxñx_><br />

1 ¢<br />

)GF<br />

1 ¢<br />

1 ¢<br />

Y )<br />

@<br />

¢<br />

@ r Y<br />

¢<br />

',Î<br />

',Î<br />

>])<br />

¢ r ><br />

which is the desired result!<br />

'*Î<br />

> ¢<br />

,<br />

¯m¯<br />

Y Î<br />

CYLINDRICAL COORDINATES<br />

It is easy to generalize the result for polar coordinates to cylindrical coordinates<br />

;arcedåë


¾<br />

r<br />

Î<br />

)<br />

¾<br />

1 ; ' ë_'*+<br />

><br />

r<br />

)<br />

@<br />

;<br />

Î<br />

r<br />

)<br />

®<br />

<br />

<br />

+ +<br />

1<br />

r<br />

4<br />

' ë(',+ ; ) r ><br />

; 1 4<br />

r ',ë_',+<br />

><br />

Y<br />

¢<br />

Y<br />

4<br />

4<br />

><br />

Y<br />

ë<br />

r<br />

r<br />

)<br />

; 4 1<br />

r ë_'*+ '<br />

r @<br />

Î<br />

Y<br />

¢<br />

Y<br />

r<br />

1 ; ',ë_',+<br />

)<br />

4<br />

4<br />

4<br />

4<br />

145<br />

;deKÆ<br />

@<br />

;<br />

Y )<br />

4 üY<br />

Y ;<br />

Y ;<br />

r Y<br />

Y ë<br />

Y +<br />

SPHERICAL COORDINATES<br />

Finally we give without proof the result for Laplace’s equation in spherical coordinates:<br />

1 ¢<br />

1 ¢<br />

Y )<br />

Î',ë<br />

1 ¢<br />

'*Î',+<br />

@<br />

¢ r 4<br />

±<br />

1 ¢<br />

1 ¢<br />

@<br />

defÆ<br />

Y<br />

Î Y<br />

r ²<br />

Y ë<br />

Y Î deKÆ<br />

For proof see e.g. Chapter 8 of Riley et al. or Chapter 2 of Arfken <strong>and</strong> Weber.<br />

1deKÆ<br />

Y )<br />

',Î',ë<br />

',Î',ë


146 APPENDIX B. CURVILINEAR COORDINATE SYSTEMS


ˆ<br />

`<br />

<br />

À<br />

@<br />

<br />

¦ 1 ; ¾<br />

Ä<br />

j Ä<br />

Ä<br />

Ä<br />

j<br />

j<br />

À<br />

<br />

T<br />

Appendix C<br />

Solutions to problems<br />

Chapter 1<br />

Problem 1.4.1: Mass diffusivity in terms of concentration<br />

Show how to obtain (1.31) from (1.32).<br />

Solution:<br />

Consider the product ¾ ¦ ; j<br />

<strong>and</strong> the definition of Ä<br />

j ¢ ; j :


as:<br />

! ` ¢<br />

©<br />

<br />

|‚‡~nÔÖÕº×F<br />

Define<br />

R<br />

R<br />

!<br />

"<br />

R<br />

R<br />

R<br />

R<br />

!<br />

"<br />

!<br />

"<br />

<br />

R<br />

¡<br />

’<br />

Î<br />

Î<br />

8<br />

Î<br />

’<br />

R<br />

’<br />

Î<br />

’<br />

!<br />

’<br />

Î<br />

<br />

148 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

Figure C.1: Sliding plate<br />

!WÒ<br />

defÆ<br />

from which<br />

!åÒ ’‰defÆ<br />

8 ¡<br />

To find the time required to accelerate to the velocity<br />

equation of motion of the accelerated plate:<br />

!<br />

¼ !GÒ , consider the<br />

` 8 ¡<br />

dividing by<br />

, we get:<br />

<br />

defÆ<br />

` ! 8 ¡<br />

after rearranging:<br />

¼<br />

defÆ<br />

— ! ` ’‰defÆ<br />

` 8 ¡<br />

. Then:<br />

8 ¡<br />

` 8 ¡<br />

"


whereV<br />

<br />

corresponds to the initial velocity ! V<br />

<br />

Dividing the above by !åÒ 1 ’‘deKÆ<br />

’<br />

which can be solved for time as a function of<br />

: !GÒ !<br />

!<br />

!åÒ<br />

`<br />

Î<br />

Vhjlkû—<br />

@<br />

` hjlk<br />

8 ¡<br />

hjlk<br />

Î<br />

’<br />

<br />

Î<br />

@<br />

"<br />

<br />

`<br />

T<br />

"<br />

<br />

!<br />

!WÒ<br />

<br />

’<br />

"<br />

<br />

149<br />

With the solution:<br />

8 ¡<br />

`<br />

. Thus<br />

` 8 ¡<br />

! ` ’‰deKÆ<br />

` ’‘deKÆ<br />

8 ¡<br />

, we obtain:<br />

—<br />

4 1 8 ¡ 4<br />

` 8 ¡<br />

:<br />

—<br />

1 ! : !åÒ 4 ` ’<br />

"<br />

¡ 8<br />

dKÆ —<br />

Matlab solution:<br />

% Lubrication<br />

g=9.8<br />

% m/sˆ2<br />

m=2 % kg<br />

tet=10/180*pi % RAD: angle<br />

mu=5e-3 % kg/(m s)<br />

h=3e-4 % m<br />

alp=0.99<br />

A=0.3*0.4 % mˆ2: area of the plate<br />

Vinf=m*g*sin(tet)*h/(mu*A) % =1.702 m/s<br />

t=-m*h/(mu*A)*log(1-alp) % =4.605 s


j<br />

<br />

<br />

)<br />

<br />

n<br />

<br />

Ý<br />

Ä<br />

<br />

)<br />

ß<br />

j<br />

Ý<br />

<br />

Ý <br />

ß<br />

n<br />

<br />

j<br />

j<br />

1<br />

)<br />

n<br />

j<br />

`<br />

<br />

Ý<br />

<br />

Ý<br />

`<br />

j<br />

j<br />

Ý<br />

Ý 4<br />

;Ãé F<br />

n<br />

Ý<br />

y<br />

PmP<br />

j<br />

j<br />

150 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

Chapter 2<br />

Problem 2.7.2: 2D vorticity limit<br />

Perform the missing steps in (2.42).<br />

Solution:<br />

Using the tensor identity (A.29), we have:<br />

p5¦<br />

m$<br />

9 p5¦/<br />

naj<br />

9 1 ß <br />

`]ß m ß<br />

y<br />

n­j<br />

y `]ß m ß<br />

m 4<br />

9 1 ß <br />

(C.1)<br />

n­j<br />

n­j<br />

9 1<br />

³<br />

`<br />

î 4<br />

Problem 2.7.3: Incompressible viscous limit<br />

Derive (2.50):<br />

(C.2)<br />

k<br />

9 n<br />

m<br />

j*j<br />

from (2.49)<br />

(C.3)<br />

¦ 9 )<br />

¦f<br />

j*j<br />

` ;Ãé F º<br />

¦<br />

Solution:<br />

Using the expression of vorticity vector (1.12):<br />

(C.4)<br />

¦$<br />

pm¦/<br />

4<br />

@<br />

l<br />

m<br />

j<br />

we can form the following equation from (C.3):<br />

)åj<br />

(C.5)<br />

m<br />

j<br />

æ€æ<br />

pm¦/<br />

<br />

p5¦7<br />

m<br />

j<br />

9 p5¦7<br />

jè)<br />

jèº<br />

9 pm¦/<br />

jÙ)


¦<br />

R<br />

j<br />

R<br />

¦<br />

"<br />

`<br />

<br />

j<br />

?<br />

R<br />

<br />

n<br />

¦<br />

?<br />

<br />

R<br />

¦<br />

¦<br />

><br />

?<br />

T<br />

R<br />

¦<br />

j<br />

¦<br />

1<br />

)<br />

¦<br />

<br />

æñæ<br />

¦<br />

R<br />

¦<br />

–<br />

`<br />

¥<br />

R<br />

¦<br />

"<br />

l<br />

4 <br />

æñæ<br />

<br />

<br />

T<br />

where the pressure term on the RHS became zero by symmetric identity (A.27).<br />

Now we can form another equation similar to (C.5):<br />

151<br />

(C.6)<br />

9 p5¦/<br />

<br />

)Wj<br />

p5¦/<br />

<br />

jè)Wj<br />

Subtracting (C.6) from (C.5), we have:<br />

p5¦/<br />

4 <br />

9 p5¦7<br />

which after comparison with (C.4) is reduced to:<br />

1<br />

)<br />

m<br />

j<br />

<br />

)Wj<br />

m<br />

j<br />

)Wj<br />

Problem 2.7.4: Conservation of circulation<br />

The velocity circulation is defined as<br />

9 n<br />

<br />

æñæ<br />

(C.7) ><br />

)<br />

where the integration is over any closed loop inside the fluid.<br />

Ø?<br />

Show that for irrotational (n flow<br />

time, we have:<br />

¦­<br />

R <br />

):><br />

<br />

= . "<br />

. Differentiating (C.7) over<br />

)<br />

— R <br />

" R <br />

)<br />

<br />

Using (1.2), we can rewrite the last term on the RHS as:<br />

R<br />

"<br />

R><br />

¦<br />

r<br />

Ù? R<br />

— R <br />

¦<br />

— )<br />

?<br />

)<br />

)<br />

R )<br />

?<br />

R<br />

the last equality stems from the fact that the integral of a total differential over a<br />

closed loop is zero. Thus we have for the rate of change of circulation:<br />

Substituting the velocity derivative from the Euler equation in form (2.53), we<br />

have:<br />

)<br />

R<br />

"<br />

R><br />

" R


; R ’<br />

(C.8) "<br />

R<br />

<br />

Y<br />

"<br />

Y<br />

Ö<br />

j<br />

<br />

<br />

º<br />

X R<br />

" ><br />

R<br />

¦<br />

)<br />

¦<br />

<br />

s<br />

<br />

T<br />

`<br />

<br />

X<br />

–<br />

¥<br />

Ö<br />

¦<br />

<br />

j<br />

¦<br />

R<br />

T<br />

152 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

` @<br />

¦<br />

;?<br />

º<br />

R <br />

According to the Stokes theorem the contour integral on the RHS can be<br />

rewritten as the integral over the surface encircled by the contour:<br />

R<br />

"<br />

R><br />

<br />

R j<br />

<br />

j<br />

p5¦/<br />

¾ <br />

£‰¦$<br />

pm¦/<br />

£‰¦$<br />

R<br />

"<br />

R><br />

jÙº<br />

@ `<br />

;<br />

@ `<br />

;<br />

where the last equality is due to the symmetry º of <strong>and</strong> the symmetric identity<br />

(A.27). Thus, the time derivative of circulation is zero:<br />

This is the law of circulation or Kelvin’s theorem, which states that in an ideal<br />

fluid the velocity circulation round a closed contour is constant in time.<br />

Problem 2.7.5: Bernoulli’s equation.<br />

R<br />

"<br />

R><br />

Using the energy equation (2.77):<br />

ã 6 ¦/<br />

1 °<br />

¦ 4 ¦<br />

m ¦<br />

>^)<br />

<strong>and</strong> momentum equation (2.21):<br />

(C.9)<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

6 ¦<br />

j >äã<br />

<br />

j<br />

¦ 4<br />

¦<br />

>])Wj<br />

>Þš<br />

derive the strong formulation of the Bernoulli’s equation:<br />

’¬><br />

> +<br />

@<br />

) l<br />

= . "<br />

<strong>and</strong> formulate it’s applicability limits.<br />

Solution:<br />

Method 1<br />

Assuming steady state inviscid fluid at constant temperature, we have according<br />

to (C.8):


(C.12)<br />

; )<br />

<strong>and</strong> the fact that X<br />

<br />

<br />

T<br />

¦åR<br />

R<br />

`<br />

X<br />

R ;<br />

R<br />

¦<br />

)<br />

X R<br />

"<br />

R<br />

¦<br />

R ’ ;<br />

"<br />

R<br />

’ R<br />

"<br />

R<br />

<br />

<br />

`<br />

<br />

)<br />

X<br />

X R<br />

"<br />

R<br />

X R<br />

"<br />

R<br />

¦<br />

X<br />

R X `<br />

" > ; R<br />

R<br />

"<br />

R<br />

for a steady-state solution. Solving for¤P ¤<br />

(C.12)<br />

X R<br />

"<br />

R<br />

’ R<br />

"<br />

R<br />

R<br />

R<br />

"<br />

<br />

—<br />

<br />

R ;<br />

R<br />

R<br />

R<br />

"<br />

"<br />

<br />

<br />

¦<br />

)<br />

R<br />

R<br />

<br />

R<br />

R<br />

"<br />

¦<br />

<br />

"<br />

1<br />

)<br />

1<br />

)<br />

¦<br />

)<br />

<br />

¦<br />

)<br />

<br />

153<br />

which can be rewritten as<br />

(C.10)<br />

@<br />

;<br />

On the other h<strong>and</strong> with the same assumptions (C.9) can be rewritten as<br />

(C.11)<br />

¦$<br />

¦<br />

" )<br />

> ; ;


`<br />

1<br />

)<br />

¦<br />

¦<br />

¦<br />

¦<br />

R<br />

R<br />

X<br />

¦<br />

)<br />

)<br />

<br />

`<br />

¦<br />

’<br />

¦<br />

’<br />

T<br />

’<br />

<br />

)<br />

)<br />

–<br />

¥<br />

T<br />

T<br />

`<br />

<br />

154 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

which means:<br />

(C.13)<br />

¦ ` ¦<br />

¦$<br />

@<br />

) l<br />

= . "<br />

This is the strong form of a Bernoulli’s equation valid for steady-state inviscid flow.<br />

When gravity is directed opposite to z-axis the last term on the RHS will be equal<br />

+ to .<br />

¦ <br />

Method 2<br />

¦$<br />

From (C.10) above it follows that<br />

’><br />

Ä where is the integration constant. Substituting it into (C.11) above <strong>and</strong> dividing<br />

by we obtain:<br />

;<br />

; ’>\Ä<br />

¦G<br />

¦<br />

Expansing the expression for substantial derivative (1.7), rearranging terms, <strong>and</strong><br />

using the steady state assumption , we have:<br />

<br />

¦G<br />

" )<br />

> ; Þ’<br />

¦ `<br />

¦K ^)<br />

)Wj)<br />

¦<br />

¦ `<br />

¦K ^)<br />

¦<br />

¦ ` ¦<br />

The first term can be rewritten as 1 )åj)Wj<br />

, <strong>and</strong> the last as 1 j Wj<br />

:<br />

)Wjè)Wj<br />

>Þ’<br />

4 ¦ : l<br />

4 ¦<br />

4 ¦<br />

¦ ` 1<br />

T<br />

)Wj)åj<br />

>\’<br />

jWj<br />

>Þ’<br />

j WjM ¦<br />

@<br />

l<br />

¦G<br />

L)WjÙ)åj 4<br />

l


¦<br />

R<br />

R<br />

!<br />

"<br />

in (2.94) in terms of n©'q®<br />

<br />

@<br />

À<br />

R<br />

R<br />

3 Ð R©<br />

R<br />

<br />

F<br />

@ !<br />

À<br />

"<br />

r<br />

Ö<br />

<br />

@ !<br />

À<br />

¦<br />

R<br />

F<br />

Ö<br />

'<br />

r<br />

<br />

¦<br />

R<br />

@<br />

À<br />

Ö<br />

<br />

@ !<br />

À<br />

l<br />

Ö<br />

n<br />

R<br />

Í ¢<br />

!<br />

R<br />

R<br />

><br />

!<br />

¦<br />

Ö<br />

<br />

¦<br />

R<br />

n<br />

@<br />

À<br />

Ö<br />

n<br />

)<br />

¦<br />

R<br />

'<br />

'<br />

£<br />

'BAF<br />

'BAr<br />

'BA<br />

155<br />

From which we obtain (C.13).<br />

Problem 2.7.6: Volume change inside a moving boundary<br />

Suppose that a region of space is enclosed by a moving boundary. The<br />

velocity of motion of the ) boundary, , is given at each point on the boundary.<br />

Show that the rate of change of the volume, , of that region will be equal to:<br />

!<br />

Solution<br />

Consider any volume of a moving fluid<br />

! Ö<br />

where the integration is done over an arbitrary control volume inside the fluid.<br />

Using identity (A.34), we can rewrite the latter as:<br />

¦f ¦<br />

Then by Gauss theorem:<br />

¦f ¦<br />

£‰¦<br />

£Ž¦<br />

where is a coordinate vector at the boundary R <strong>and</strong> is an element of the<br />

<br />

boundary surface area (2.1). Using the definition of velocity (1.1), we can then<br />

obtain the relation for a volume change:<br />

(C.14)<br />

£‰¦$<br />

£‰¦$<br />

£‰¦<br />

R <br />

" R<br />

Problem 2.7.7: Rotating coordinates<br />

Solution<br />

Writing out (2.93):<br />

£Žr<br />

Obtain explicit relations for the components of acceleration vectors<br />

F<br />

£ . '<br />

'*®<br />

'*® '<br />

Í]1<br />

Í ¡4<br />

" ><br />

by components yields the needed terms in (2.94):<br />

R<br />

" ><br />


R<br />

R<br />

r<br />

ÿ<br />

r<br />

"<br />

3 R¢<br />

R<br />

" > <br />

; –<br />

(C.18)<br />

P<br />

s R<br />

"<br />

R<br />

3 £¢<br />

<br />

l<br />

n<br />

£<br />

F<br />

AF<br />

R<br />

R<br />

X üR<br />

" ><br />

R<br />

n<br />

Í ¢<br />

<br />

<br />

ÿ<br />

"<br />

><br />

°<br />

£<br />

<br />

`<br />

R<br />

s<br />

<br />

`<br />

ÿ<br />

"<br />

l<br />

n<br />

n<br />

n<br />

n<br />

n<br />

<br />

r<br />

n<br />

<br />

TF<br />

F r ®<br />

r<br />

®<br />

<br />

T<br />

r<br />

n<br />

'<br />

n<br />

T<br />

)<br />

n<br />

156 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

` l<br />

£Žr <br />

(C.15)<br />

Ar <br />

Problem 2.7.8: Rotation with separated coordinate origins<br />

Consider a simple rotation with n<br />

A<br />

T<br />

˜%<br />

n<br />

ý<br />

origin of the rotating coordinate system rotate with the same around the<br />

origin of :<br />

'5nŒ- as in (2.94), but now let the<br />

Derive the expression for<br />

3<br />

Solution<br />

in this case.<br />

With these assumptions (2.89) becomes<br />

Í ÿ<br />

(C.16)<br />

> ¡4<br />

Í]1 ÿ<br />

After second differentiation<br />

>Ìn<br />

͇R<br />

Í]1<br />

Í ÿ 4<br />

the realtion (2.93) becomes<br />

(C.17)<br />

Í]1<br />

Í]1 ÿ<br />

> ¡4q4<br />

Problem 2.7.9: Nondimesionalizing energy equation<br />

Write a non-dimensional form of the heat convection equation (2.79):<br />

¦ì¦<br />

m ¦ 1<br />

¦K<br />

y ¦ 4<br />

> 8 )<br />

>])


V<br />

V<br />

)<br />

r<br />

V<br />

<br />

<br />

Since the density is constant we can replace it with ; ;<br />

)<br />

V<br />

<br />

<br />

V<br />

V<br />

s<br />

"<br />

;<br />

– s<br />

" P V<br />

s<br />

"<br />

V<br />

–<br />

P<br />

V<br />

–<br />

P<br />

s<br />

"<br />

r<br />

V<br />

V<br />

s<br />

r<br />

V<br />

V<br />

s<br />

X<br />

X<br />

¦<br />

r<br />

" V<br />

V<br />

)<br />

;<br />

<br />

V<br />

<br />

V<br />

r<br />

V<br />

X<br />

; ã s V<br />

–<br />

P V<br />

°<br />

;<br />

V<br />

V<br />

–<br />

P<br />

°<br />

)<br />

V<br />

s r<br />

<br />

s ã<br />

¦<br />

V<br />

–<br />

P<br />

><br />

–<br />

P<br />

V<br />

r<br />

V<br />

V<br />

s<br />

s ã<br />

)<br />

><br />

r<br />

V<br />

V<br />

s<br />

V<br />

<br />

<br />

><br />

) r<br />

8<br />

;<br />

<br />

V<br />

V<br />

V<br />

<br />

s<br />

s<br />

) 8<br />

V<br />

r<br />

<br />

"<br />

r<br />

V<br />

V<br />

s –<br />

P V<br />

;<br />

V<br />

V<br />

– s<br />

P<br />

V<br />

r<br />

V<br />

V<br />

V<br />

V<br />

V<br />

V<br />

<br />

;<br />

V<br />

X<br />

)<br />

V<br />

P<br />

r<br />

V<br />

'<br />

°<br />

157<br />

;<br />

Select for the pressure scale. Determine the minimum number of dimensionless<br />

parameters. Write the equation using the Eckert number as one of<br />

X<br />

the parameters:<br />

¢ )<br />

„Œz<br />

Solution<br />

Introducing the non-dimensional variables:<br />

(C.19)<br />

"<br />

"<br />

¦$ )<br />

ã " <br />

¦$<br />

ã X<br />

V<br />

ã s<br />

V<br />

ã <br />

V<br />

ã )<br />

V<br />

, <strong>and</strong> obtain:<br />

¦ì¦<br />

m ¦ 4<br />

m ¦ 1<br />

R ã<br />

)<br />

¦f<br />

R ã<br />

R ã<br />

ã " ><br />

R<br />

) ã<br />

) >«ã<br />

) ã<br />

Rearranging:<br />

¦ì¦<br />

y ¦ 4<br />

m ¦ 1<br />

R ã<br />

)<br />

¦f<br />

R ã<br />

°•"<br />

" ><br />

) ã<br />

) ã<br />

) >çã<br />

considering that the velocity scale should relate to length <strong>and</strong> time scales as<br />

, we have:<br />

R ã<br />

R ã<br />

: "<br />

(C.20)<br />

R ã<br />

8 )<br />

)<br />

R ã<br />

¦ì¦<br />

m ¦ 1<br />

¦f<br />

m ¦ 4<br />

" ><br />

) ã<br />

) ã<br />

) >çã<br />

R ã<br />

R ã<br />

This equation contains 7 dimensional 'q) 'qs ' ; ',– ' 8<br />

parameters: . According<br />

to the PI-theorem, the number of dimensionless parameters can be as low as<br />

three. A conventional choice of these parameters is:<br />

¢ )<br />

„Œz<br />

A!<br />

¢<br />

;<br />

8


s<br />

"<br />

<br />

X<br />

@<br />

s ã<br />

><br />

°<br />

P<br />

)<br />

158 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

¢ 8 –<br />

<strong>and</strong> (C.20) becomes:<br />

º(x<br />

R ã<br />

R ã<br />

R ã<br />

A! 5º_x<br />

„Œz<br />

" ><br />

A!<br />

ã<br />

) ã<br />

) >çã<br />

R ã<br />

¦ì¦<br />

„Œz<br />

m ¦ 1<br />

¦f<br />

m ¦ 4


)<br />

r<br />

s<br />

)<br />

V<br />

¦<br />

)<br />

– ;<br />

P<br />

Ä<br />

r<br />

s<br />

¦<br />

s<br />

s<br />

> 8 — R ) r<br />

® R<br />

F<br />

)<br />

1<br />

®<br />

)<br />

<br />

1<br />

®<br />

F<br />

Ä<br />

)<br />

F<br />

R<br />

s<br />

r<br />

)<br />

r<br />

<br />

)<br />

F<br />

V<br />

T<br />

T<br />

T<br />

Ä<br />

<br />

:<br />

T<br />

V<br />

<br />

F<br />

159<br />

Chapter 3<br />

Problem 3.7.1: Couette flow equations<br />

Show how to obtain equation (3.7) <strong>and</strong> (3.8).<br />

Solution<br />

Consider (3.2) <strong>and</strong> (3.3):<br />

(C.21)<br />

(C.22)<br />

(C.23)<br />

>^)Wj)<br />

¦f<br />

j<br />

9 )<br />

1 <br />

s]>])<br />

¦f<br />

j*j<br />

¦ì¦ <br />

¦G ¦f<br />

;Ãé F<br />

`<br />

º<br />

¦<br />

¦ 4 °<br />

>])<br />

y ¦ 6 ¦/<br />

Using the definition (2.17), <strong>and</strong> since the only non-zero velocity<br />

¦§%<br />

components<br />

¦Cµ%<br />

- are , <strong>and</strong> considering<br />

)<br />

continuity (2.4) <strong>and</strong> constancy º<br />

of pressure , we arrive at (3.7) <strong>and</strong> (3.8).<br />

<strong>and</strong> the only independent variable is ¦$<br />

-<br />

Problem 3.7.2: Couette plates solutions<br />

Solve equations (3.7) <strong>and</strong> (3.8) with the boundary conditions u(0) = 0 <strong>and</strong><br />

u(H) = U, s <strong>and</strong> s <strong>and</strong> .<br />

Solution<br />

1}T 4 <br />

1} 4 <br />

(C.24)<br />

r <br />

(C.25)<br />

R ®<br />

° R<br />

Aligning the coordinate origin with the lower plate (y=0) <strong>and</strong> solving the<br />

equation (C.24) with the boundary conditions: u(0) = 0 <strong>and</strong> u(H) = U, we have<br />

R ®<br />

4 <br />

®C>\Ä<br />

1UT 4 <br />

T <br />

1U 4 <br />

v<br />

Ä<br />

4 <br />

®


@<br />

1 4<br />

®<br />

s<br />

1<br />

®<br />

¥<br />

<br />

s<br />

1<br />

®<br />

` 8 4<br />

°<br />

Ä<br />

F<br />

s<br />

®<br />

r<br />

<br />

s<br />

F<br />

`<br />

®<br />

l<br />

s<br />

r<br />

F<br />

V<br />

s<br />

`<br />

<br />

s<br />

Ä<br />

F<br />

V<br />

<br />

V<br />

<br />

<br />

À<br />

V<br />

160 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

Using this solution we can solve equation (C.25):<br />

>\Ä<br />

®C>\Ä<br />

1UT 4 <br />

8<br />

° l<br />

V ><br />

(C.26)<br />

` 8 4<br />

°<br />

8<br />

° l<br />

— s<br />

®»>às<br />

V ><br />

l ><br />

Problem 3.7.3: Flow of a liquid film<br />

Ï ¿ , flowing steadily<br />

Consider a wide fluid film of ’ @‰I<br />

constant<br />

T<br />

thickness,<br />

due to the gravity down the inclined plate Î<br />

{<br />

at angle . Find an analytical<br />

expression of a fluid velocity distribution as a function of a distance from the plate<br />

) surface:<br />

. Assuming the viscosity <strong>and</strong> the density of the fluid are 8 <br />

é ° 1 ¹. 4 T WT‰T °<br />

, <strong>and</strong> ’ :< <br />

respectively, find the maximum flow velocity<br />

:<br />

<strong>and</strong> the volumetric ¢ flow rate, , per 1m of the plate. Atmospheric … pressure can<br />

be considered constant.<br />

@‰I,[<br />

ž<br />

Figure C.2: Flow of a liquid film.<br />

Solution<br />

Selecting the coordinate parallel to the plate <strong>and</strong> in the direction of the<br />

steepest decent, <strong>and</strong> ® normal to the plate, we can conclude that the steady


(C.27) )<br />

6<br />

<br />

1<br />

®<br />

T<br />

<br />

’<br />

…<br />

<br />

)<br />

r<br />

)<br />

T 8 Y ) 4<br />

® Y<br />

)<br />

V<br />

|<br />

<br />

V<br />

Î<br />

Î<br />

¤«<br />

4<br />

4<br />

<br />

r<br />

><br />

Î<br />

’<br />

Î<br />

4<br />

¡<br />

4<br />

`<br />

Î<br />

<br />

’<br />

4<br />

r<br />

l<br />

®<br />

4<br />

¤<br />

=<br />

Î<br />

4<br />

1 4<br />

Î<br />

<br />

’<br />

4<br />

)<br />

1 4<br />

®<br />

<br />

‹<br />

<br />

T<br />

state solution should satisfy the following )<br />

constraints: ,<br />

. Then the momentum equation (3.2) reduces to:<br />

161<br />

, <strong>and</strong><br />

Y X : Y <br />

Y X : Y ®<br />

8 R<br />

r ` ; defÆ 1<br />

with the solution:<br />

R ®<br />

` ; defÆ 1<br />

l 8 ®<br />

®ƒ><br />

where the constants ¡ <strong>and</strong> ¤ can be determined from the boundary conditions:<br />

1}T 4 T v<br />

T<br />

The second condition is the negligible shear stress at the free surface:<br />

` ; defÆ 1<br />

1 l<br />

,¯<br />

> 8 ¡<br />

thus:<br />

O¯Y~<br />

’‰defÆ 1<br />

¡£ ;<br />

Substituting <strong>and</strong> into (C.27) we obtain the final solution for the axial flow<br />

velocity distribution with :<br />

1 l<br />

; defÆ 1<br />

l 8 ®<br />

The maximum velocity will correspond to ®<br />

’ :<br />

; deKÆ 1<br />

The volumetric flow rate is obtained by integrating the velocity along the<br />

thickness <strong>and</strong> the width of the film:<br />

)œ|aw<br />

l 8 ’<br />

. #<br />

Ö F<br />

Ö ~<br />

;<br />

) R ® R ›<br />

À 8


¾<br />

r<br />

ë<br />

X<br />

<br />

ë<br />

r<br />

r )<br />

<br />

r<br />

)<br />

l<br />

T<br />

4<br />

162 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

A cubic dependence of the flow-rate on ’ means that the draining of the film is<br />

strongly dependent on the film thickness.<br />

The values of )W|­w<br />

<strong>and</strong> …<br />

can be obtained from the following Matlab solution:<br />

h=1.5e-3 % m: film thickness<br />

theta=pi/6 % RAD: inclination of the plate<br />

mu=1.6e-3 % kg/(m s): viscosity<br />

rho=8e2 % kg/mˆ3: fluid density<br />

g=9.8 % m/sˆ2: gravity acceleration<br />

umax=(rho*g*hˆ2*sin(theta))/(2*mu) %=2.756 m/s<br />

Q=(rho*g*hˆ3*sin(theta))/(3*mu) %=0.002756 mˆ3/s<br />

Problem 3.7.4: Couette solution for non-Newtonian fluids<br />

How will the solution (3.9) change for a non-Newtonian fluid?<br />

Solution:<br />

For non-Newtonian fluids the relation between stress <strong>and</strong> strain is nonlinear,<br />

thus, instead of (3.5) we have:<br />

B<br />

which after the assumptions of Couette flow between parallel plates (Sec.3.2.1)<br />

leads to a modified equation (3.7):<br />

F ã 6 ¦<br />

j ;Ãé<br />

<br />

j<br />

¦f<br />

j*j<br />

9 1<br />

B <br />

8 — R<br />

which for any non-zero 8 <strong>and</strong> = has the same solution as (3.7). So, the solution<br />

for a non-Newtonian fluid will be the same.<br />

Problem 3.7.5: Momentum equation for Couette flow between concentric cylinders<br />

Using the assumptions on the Couette velocity profile between the rotating<br />

concentric cylinders (Sec.3.2.3) <strong>and</strong> the expression for the momentum equation<br />

<strong>and</strong> the Laplacian operator in cylindrical coordinates:<br />

R ®<br />

x_><br />

¢ >^)GH,)GFH­><br />

¢<br />

)œx5)IF<br />

<br />

)IF­>])Wx5)GF<br />

¼<br />

F`<br />

)GFœ)GFF<br />

)œx<br />

; ¢ > 9 1 ¾<br />

¢ r 4<br />

)GF­><br />

¢ r ` )GF<br />

@<br />

¢<br />

<br />

x<br />

4<br />

x_><br />

@ r ë ¢<br />

F„F­>Þë<br />

F<br />

F<br />

HH<br />

1 ¢


that )<br />

<br />

)<br />

¦<br />

)<br />

<br />

¢<br />

A<br />

R<br />

R<br />

V<br />

¤<br />

r<br />

¢<br />

V<br />

<br />

x<br />

6F<br />

x<br />

A<br />

R<br />

R<br />

V<br />

)<br />

<br />

6<br />

V<br />

¼ F¥6¥x<br />

)<br />

¢<br />

<br />

)<br />

¢<br />

<br />

L)<br />

¢M<br />

x<br />

xñx_><br />

A<br />

V<br />

6<br />

V<br />

V<br />

Ç<br />

'<br />

T<br />

F<br />

<br />

T<br />

163<br />

Derive equation (3.18):<br />

T<br />

¢ r ><br />

¢ ¢ M L)GF<br />

Solution<br />

)IF, <strong>and</strong><br />

, the equation (C.28) simplifies to:<br />

)IF<br />

Renaming for simplicity: )<br />

using the assumptions of (Sec.3.2.3)<br />

1 ¢ 4<br />

1 ¢<br />

<br />

x<br />

4<br />

x<br />

<br />

x<br />

<br />

x€x_><br />

)<br />

¢ r `<br />

)<br />

¢ r `<br />

(C.28)<br />

where we used the identity:<br />

<br />

x<br />

)<br />

¢ r `<br />

)<br />

L)<br />

¢‚M<br />

Problem 3.7.6: Rotation torque <strong>and</strong> power<br />

In the system of two rotating cylinders (Sec.3.2.3) consider the torque applied<br />

to the inner rotating cylinder when the outer cylinder is (n fixed ). What<br />

is the power required to rotate the inner cylinder?<br />

Solution:<br />

In this case we need to multiply the force applied at the surface of the cylinder<br />

by its radius:<br />

(C.29)<br />

l§UA<br />

¡£<br />

where is the length of the cylinder. Since we are using a curvilinear coordinate<br />

system, the expression for the shear stress tensor used in the momentum<br />

(A.45), which in cylindrical coordinate system has a form:<br />

Ç<br />

equation is 6 <br />

°<br />

V<br />

<strong>and</strong> evaluating the expression above at ¢ƒ<br />

¦Gä% T<br />

Considering the specific form of velocity )<br />

dependence:<br />

, we have A<br />

'*)GF1 ¢ 4<br />

- (Sec.3.2.3),


I<br />

6<br />

V<br />

º<br />

<br />

R<br />

¢ OxZYVÚ`<br />

R<br />

)GF<br />

A<br />

V<br />

)GF<br />

6<br />

V<br />

<br />

ø<br />

¢ M<br />

x<br />

¢L)GF<br />

<br />

n<br />

`<br />

V<br />

A<br />

n<br />

V<br />

V<br />

¤<br />

A<br />

…<br />

V<br />

r A<br />

…<br />

F<br />

<br />

><br />

<br />

r<br />

A<br />

` V<br />

<br />

F<br />

A<br />

ÑU8n<br />

n<br />

<br />

V<br />

r<br />

V<br />

A<br />

)<br />

Ç<br />

<br />

Ñ<br />

rF<br />

V<br />

A<br />

ˆ<br />

><br />

r A<br />

1 4<br />

®<br />

`<br />

V<br />

r<br />

@<br />

A<br />

A<br />

` V<br />

ÑU8n<br />

Ä<br />

ø<br />

¢ M<br />

x<br />

ù<br />

¢L)IF<br />

F<br />

A<br />

r<br />

F<br />

V<br />

r<br />

<br />

r<br />

F<br />

Ç<br />

<br />

A<br />

`<br />

r A<br />

V<br />

n<br />

n<br />

r<br />

V<br />

A<br />

` V<br />

V<br />

A<br />

A<br />

F<br />

A<br />

r<br />

: A<br />

: F<br />

rA<br />

F<br />

A<br />

r<br />

F<br />

rV<br />

V<br />

V<br />

V<br />

`<br />

@<br />

Ï<br />

`<br />

`<br />

A<br />

A<br />

rA<br />

rF<br />

F<br />

V<br />

V<br />

`<br />

A :<br />

A :<br />

@<br />

<br />

F<br />

F<br />

<br />

164 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

¢ ù<br />

)GFF<br />

xZYVÚ<br />

Computing the derivative from (3.20), we have:<br />

xZYVlÚ<br />

— A<br />

— A<br />

>\A<br />

from which we obtain M:<br />

The power needed rotate the cylinder is:<br />

°<br />

V<br />

† <br />

V°<br />

V<br />

Problem 3.7.7 Flow between parallel plates under pressure<br />

A viscous fluid with viscosity (<br />

parallel plates<br />

8<br />

` l<br />

ωÏ<br />

) is driven between two<br />

`ÞÑ•° : 1 ¹. 4<br />

Ï @‰I<br />

¿ apart by an imposed pressure gradient of º : R R<br />

volume flow rate per 1m of the plates’ width. What pressure gradient will cause<br />

the flow to reverse?<br />

£ :< . The upper plate is moving with velocity<br />

<br />

Solution<br />

– ¹:‰.<br />

. Find the<br />

Adapting the general expression for the volumetric flow rate (3.29) to the<br />

case of the flow in a square duct of unit width <strong>and</strong> height , we have:<br />

which is a volumetric flux per 1m width of the duct. Substituting the solution<br />

(3.36), (3.37) into the above, <strong>and</strong> performing the integration, we obtain:<br />

V<br />

ÖJ—<br />

R ®<br />

à1<br />

À<br />

4 :[


ë<br />

Y<br />

r<br />

Ä<br />

<br />

r<br />

Î<br />

¥<br />

r<br />

º<br />

<br />

<br />

Î<br />

4<br />

`<br />

À<br />

A<br />

¢<br />

r<br />

><br />

ë<br />

<br />

A<br />

r<br />

r<br />

T<br />

<br />

165<br />

where<br />

l<br />

The pressure gradient causing the flow reversal at the lower plate can be<br />

computed from (3.38).<br />

8<br />

Below is the Octave (Matlab) solution to this problem.<br />

% Q=Integrate[u(y),y]<br />

% where u(y) is given by (CFM:Sec.3.2.5):<br />

% u = U/H*y*(C*y/H+1-C)<br />

% where C=Hˆ2/(2*mu*U)*dpdx<br />

% Integrating, we obtain:<br />

% Q=U*H*(3-C)/6<br />

% Express all values in SI units:<br />

U=0.15 % m/s<br />

H=0.005 % m<br />

mu=1.4e-4 % kg/(m s)<br />

dpdx=-2.55 % kg/(mˆ2 sˆ2)<br />

C=Hˆ2/(2*mu*U)*dpdx<br />

Q=U*H*(3-C)/6 %=0.0005647 mˆ2/s =(mˆ3/s)/m<br />

% The pressure gradient that will case the<br />

% reverse flow:<br />

dpdx=2*mu*U/Hˆ2 %=1.68 kg/(mˆ2 sˆ2)<br />

Problem 3.7.8 Verifying the Stokes solution<br />

Verify that the solution (3.73):<br />

(C.30)<br />

rdeKÆ<br />

l ¢<br />

',Î<br />

@ 4<br />

Ñ£<br />

A —<br />

¢<br />

1 ¢<br />

satisfies the equation (3.72):<br />

(C.31)<br />

1 "<br />

r Y<br />

¢<br />

—\Y<br />

¢ r ><br />

r ` –<br />

Solution:<br />

@ r Y<br />

¢<br />

Î Y<br />

Y Îò


166 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

Using the symbolic manipulation library GiNaC (www.ginac.org), we can<br />

easily check the consistency of the solution. Below is the example of a C++<br />

code using the GiNaC library.<br />

/********************************************<br />

This function computes the Laplacian<br />

of the Stokes solution in spherical<br />

coordinates.<br />

It should be compiled as<br />

c++ diff.cc -o diff -lcln -lginac<br />

Using GiNaC system (www.ginac.org)<br />

Running the executable produces<br />

the following output:<br />

Laplace(F)=3/2*sin(t)ˆ2*a*U*rˆ(-1)<br />

Laplace2(F)=0<br />

********************************************/<br />

#include <br />

#include <br />

using namespace std;<br />

using namespace GiNaC;<br />

ex Laplace<br />

(<br />

const ex & F,<br />

const symbol & r,<br />

const symbol & t<br />

)<br />

{ return normal<br />

( F.diff(r,2)<br />

+ F.diff(t,2)/pow(r,2)<br />

- cos(t)*F.diff(t)/(sin(t)*pow(r,2))<br />

);<br />

}<br />

int main()<br />

{


(C.34)<br />

6 x„F<br />

ë<br />

<br />

Î<br />

rdeKÆ<br />

x<br />

—<br />

`<br />

r<br />

Î<br />

<br />

A<br />

A<br />

<br />

<br />

`<br />

`<br />

À<br />

A<br />

¢<br />

<br />

><br />

<br />

A<br />

Â<br />

r<br />

r<br />

<br />

Î<br />

167<br />

symbol r("r"), t("t"), U("U"), a("a");<br />

ex<br />

F=normal<br />

(<br />

U/4*pow(a*sin(t),2)*(a/r-3*r/a+2*pow(r/a,2))<br />

),<br />

LF=Laplace(F,r,t),<br />

LF2=Laplace(LF,r,t);<br />

cout<br />


dropped into oil of density ; {<br />

<br />

T<br />

IWeW;—rZ²<br />

168 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

This function computes velocity components<br />

<strong>and</strong> viscous stress of the Stokes solution.<br />

It should be compiles as:<br />

c++ stokes.cc -o stokes -lcln -lginac<br />

Using GiNaC system (www.ginac.org)<br />

Running the executable produces<br />

the following output:<br />

Ur=1/2*(2*rˆ3+aˆ3-3*a*rˆ2)*cos(t)*U*rˆ(-3)<br />

Ut=3/4*sin(t)*a*U*rˆ(-1)-sin(t)*U+1/4*sin(t)*aˆ3*U*rˆ(-3)<br />

tau=-3/2*mu*sin(t)*aˆ3*U*rˆ(-4)<br />

********************************************/<br />

#include <br />

#include <br />

using namespace std;<br />

using namespace GiNaC;<br />

int main()<br />

{ symbol r("r"), t("t"), U("U"), a("a"), mu("mu");<br />

ex F=normal<br />

(<br />

U/4*pow(a*sin(t),2)<br />

*(a/r-3*r/a+2*pow(r/a,2))<br />

),<br />

Ur= F.diff(t)/(pow(r,2)*sin(t)),<br />

Ut=-F.diff(r)/(r*sin(t)),<br />

dU= Ur.diff(t)/r + Ut.diff(r) - Ut/r;<br />

cout<br />


¤<br />

Ü<br />

F<br />

<br />

T<br />

Ï<br />

º<br />

<br />

¤<br />

¤<br />

¤<br />

R<br />

R<br />

<br />

U[<br />

R<br />

<br />

Ü<br />

UÑ R<br />

<br />

r<br />

<br />

r<br />

¤<br />

F<br />

¤<br />

l V<br />

Ñ : l<br />

A»ˆ<br />

!<br />

@•><br />

Ð[<br />

V<br />

F<br />

1<br />

R<br />

r<br />

V<br />

r<br />

!<br />

!<br />

V<br />

4<br />

4<br />

T<br />

@<br />

Ñ<br />

I<br />

169<br />

Solution<br />

The balance of forces on the sphere when it reached the terminal velocity<br />

is: ¤<br />

<strong>and</strong> viscosity 8 T<br />

diameter is (a) R<br />

. £<br />

. Estimate the terminal velocity of the sphere<br />

<br />

if<br />

T<br />

its<br />

<strong>and</strong> (c) R ¿ .<br />

IK@<br />

IO@ ¿ , (b) R<br />

<br />

@ ¿<br />

where<br />

is the buoyancy <strong>and</strong><br />

¤is a drag force:<br />

A<br />

1 ;<br />

` ;<br />

Ĥ;<br />

;<br />

are densities of the oil <strong>and</strong> sphere respectively, <strong>and</strong> the drag coefficient<br />

Ĥcan be first assumed for a laminar<br />

'<br />

case:<br />

where ; V<br />

Ä R<br />

;<br />

8<br />

Or, alternatively the expression for Stokes drag force can be used:<br />

¤<br />

A»ˆ<br />

These equations can be solved for velocity:<br />

ÀU8 ½<br />

` ;<br />

! R<br />

1 ;<br />

@W8<br />

If the computed velocity leads to<br />

Ĥshould be used:<br />

A#<br />

then the turbulent approximation for Û¥@<br />

l Ñ<br />

Ä R<br />

A»ˆ<br />

4 ><br />

><br />

Below is the Octave solution:<br />

A!


170 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

g=9.8 % gravity<br />

denw=1000 % density of water<br />

den0=0.88*denw % density of oil<br />

den1=7.8*denw % density of the sphere<br />

mu=0.15<br />

% (a):<br />

d=0.1e-3<br />

V=dˆ2*(den1 - den0)*g/(18*mu) % =0.00025<br />

Re=den0*V*d/mu<br />

% =0.00014735: laminar<br />

% (b):<br />

d=1e-3<br />

V=dˆ2*(den1 - den0)*g/(18*mu) % =0.025<br />

Re=den0*V*d/mu<br />

% =0.147: laminar<br />

% (c):<br />

d=1e-2<br />

V=dˆ2*(den1 - den0)*g/(18*mu) % =2.5117<br />

V1=0.0<br />

% It’s a turbulent case so we loop until convergence:<br />

while (abs(V1-V)>0.1*abs(V))<br />

V1=V<br />

Re=den0*V*d/mu<br />

Cd=24/Re + 6/(1+sqrt(Re)) + 0.4<br />

Fb=pi*dˆ3/6*(den1 - den0)*g<br />

Fd=Fb<br />

V=sqrt(Fd/(pi/4*dˆ2*Cd*den0/2))<br />

end<br />

which results in<br />

Re = 48.569<br />

V = 0.78984<br />

% m/s


ß<br />

V<br />

Ö˜±<br />

( 1 ß : <br />

1<br />

<br />

T<br />

I<br />

) <br />

<br />

R Î l<br />

R<br />

<br />

( 1 ß : <br />

@<br />

R ) 8<br />

® R<br />

<br />

@<br />

`<br />

`<br />

R l<br />

R<br />

Î<br />

<br />

@<br />

L®<br />

ßM<br />

l<br />

V<br />

T<br />

I<br />

(Äkõm A"<br />

ß<br />

)<br />

<br />

<br />

À<br />

<br />

ß<br />

171<br />

Problem 3.7.11: Boundary layer analysis for cubic velocity profile<br />

Repeat the boundary layer analysis of Sec.3.5.1 with assumed velocity profile:<br />

À<br />

L®<br />

l<br />

4m A! ),<br />

4m A! ),<br />

4m A" ),<br />

Compute ( Î : <br />

1 :‰9<br />

. ©<br />

(Ĥm<br />

where A!<br />

<br />

Hint: Assume )<br />

Solution<br />

¯m¯<br />

1}T 4 T<br />

.<br />

ßM<br />

`<br />

Using the definition of momentum thickness <strong>and</strong> introducing dimensionless<br />

¢<br />

coordinate , we obtain:<br />

),<br />

A" ),<br />

® : ß<br />

` )<br />

(C.35)<br />

Ï `<br />

Ï 4 1<br />

Ï 4<br />

À‰Ô<br />

£Ö˜±<br />

<br />

L@<br />

® MƒR<br />

@‰I<br />

@‰I<br />

lyWT<br />

Similarly, for displacement thickness:<br />

Ï<br />

»><br />

R<br />

<br />

(C.36)<br />

ß” ß Ö 1<br />

ß :W<br />

R ®<br />

) :4<br />

Using the definition of the wall shear stress:<br />

(C.37)<br />

6Q<br />

@JI<br />

Ï 8: ß<br />

Thus, for the friction coefficient:<br />

<br />

Äkõ<br />

l<br />

;r 6Q<br />

8: ß<br />

À<br />

;r<br />

(C.38)<br />

lyWT<br />

— ÀJÔ


s<br />

<br />

@<br />

4<br />

<br />

@<br />

1<br />

@<br />

ß<br />

ß<br />

<br />

Ð<br />

`<br />

Î<br />

ß<br />

R<br />

@<br />

T 8 R<br />

Ñ<br />

<br />

@EÀ ;<br />

l Ö 6QR<br />

4<br />

<br />

T<br />

T<br />

T<br />

Ñ<br />

l<br />

¡<br />

<br />

4<br />

@<br />

r<br />

172 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

<strong>and</strong><br />

ß <br />

<strong>and</strong> solving this for ß , we have:<br />

(C.39)<br />

I4[Ñ<br />

<br />

Using (C.35), <strong>and</strong> (C.36), we obtain:<br />

A! Ð<br />

Substituting ß 1 <br />

from (C.39) into (C.38), we obtain:<br />

<br />

Ð<br />

<br />

A! A!<br />

@JI¢ÑT<br />

I4[Ñ[<br />

Problem 3.7.12: Drag force on a triangle<br />

<br />

<br />

@‰I Ô‰À A!<br />

ÄkõÐ<br />

A thin equilateral triangle plate (Fig.C.3) with the edge length , is<br />

l £Þ ¹:‰. l<br />

at<br />

ĤРI4[Ñ[<br />

temperature<br />

immersed parallel to a stream of air with the velocity<br />

<br />

this plate.<br />

l {<br />

Ä <strong>and</strong> pressure º<br />

Solution<br />

Using (3.125), we have:<br />

£ "<br />

. Assuming laminar flow, estimate the drag on<br />

F}É<br />

— ;Ž8<br />

l<br />

<br />

; šgg1}T 4<br />

IÁÀ‰À<br />

The drag force on one side of the triangle is: ¤<br />

l<br />

<br />

9 o<br />

6Q<br />

<br />

The area element R is , ’ <strong>and</strong><br />

. For two sides of the triangle, we have for the total drag force:<br />

R <br />

: ’<br />

, where <br />

Ç<br />

',’<br />

УdeKÆ 1U: À<br />

¡ÐУ<br />

1}T<br />

4 <br />

£mÀ : l


¤<br />

l Ö 6QR<br />

V<br />

1<br />

@<br />

l 4WÀ<br />

Üm<br />

`<br />

4<br />

r<br />

F}É Ý<br />

¤<br />

r<br />

<br />

<br />

lV<br />

l 4WÀ<br />

r<br />

r<br />

`<br />

<br />

r<br />

r<br />

`<br />

V<br />

@<br />

’<br />

:<br />

r<br />

F}É<br />

l<br />

ÀŽ’<br />

T<br />

1<br />

@<br />

<br />

V<br />

`<br />

r<br />

Ñ<br />

F}É : ÀŽ’<br />

À<br />

l<br />

r§Üm<br />

r<br />

4<br />

r<br />

F}É £Ý<br />

r<br />

r<br />

173<br />

Figure C.3: Triangular plate pulled in a viscous fluid<br />

£Ö ~<br />

¡¼ l 1}T<br />

IÁÀ‰À<br />

l 4 1 ;Ž8 4<br />

é F}É<br />

F}É<br />

R <br />

Computing the integral:<br />

: ’<br />

Ö ~<br />

Ö ~<br />

: ’<br />

Ö ~<br />

é F}É<br />

R <br />

R <br />

é F}É<br />

R <br />

F}É<br />

’ F}É<br />

£ 1 ;Ž8 4<br />

F}É<br />

1UT<br />

IÕÀJÀ<br />

Tœ1 ;Ž8 £ 4<br />

1UT<br />

1 ;Ž8 £ 4<br />

À<br />

l<br />

IW‰l<br />

IÁÀ‰À<br />

F}É<br />

À‰Ô<br />

F}É<br />

a=2 % m<br />

U=12 % m/s<br />

rho=1.2 % kg/mˆ3<br />

mu=1.8e-5 % kg/(m s)<br />

F=0.332*8/3*(sqrt(3)/2)ˆ(1/2)*(rho*mu*(a*U)ˆ3)ˆ(1/2) %=0.45 N<br />

Problem 3.7.13: Boundary layer equations<br />

Derive Pr<strong>and</strong>tl boundary layer equations (3.110) - (3.113):


F<br />

ã<br />

<br />

<br />

<br />

<br />

F<br />

)<br />

r<br />

)<br />

¾<br />

F<br />

¾<br />

F<br />

–<br />

P<br />

)<br />

F<br />

)<br />

F<br />

F<br />

r<br />

F<br />

r<br />

<br />

<br />

)<br />

F<br />

ã<br />

F<br />

¦<br />

s<br />

r<br />

F<br />

)<br />

)<br />

<br />

F<br />

ã s<br />

F<br />

¦<br />

)<br />

)<br />

F<br />

F<br />

<br />

F<br />

<br />

r<br />

F<br />

<br />

V<br />

<br />

)<br />

)<br />

F<br />

s<br />

ã r<br />

s<br />

¦f“ 9 )<br />

¾<br />

F<br />

¾<br />

F<br />

<br />

V<br />

¾<br />

)<br />

F<br />

)<br />

F<br />

F<br />

s ã<br />

ã<br />

)<br />

F<br />

F<br />

r<br />

T<br />

r<br />

)<br />

r<br />

)<br />

><br />

) ã<br />

<br />

F<br />

X ã<br />

1<br />

)<br />

X<br />

><br />

¦ <br />

> ;<br />

F<br />

)<br />

)<br />

¤<br />

ã<br />

<br />

<br />

F<br />

F<br />

T x<br />

r<br />

T<br />

V<br />

T<br />

¦<br />

`<br />

r ¾ r*<br />

X<br />

) ;<br />

¾<br />

ã<br />

r<br />

X<br />

F<br />

X<br />

F<br />

<br />

V<br />

V<br />

"<br />

¦<br />

V<br />

174 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

(C.40)<br />

(C.41)<br />

<br />

) ã<br />

¦f ¦G<br />

r“<br />

rðr `<br />

(C.42)<br />

X ã<br />

) >äã<br />

) ã<br />

) ã<br />

) >äã<br />

) ã<br />

r•<br />

(C.43)<br />

<br />

sà>«ã ) ã<br />

r 4<br />

rðr<br />

) ã<br />

r<br />

using equations (3.107) - (3.107):<br />

) >çã<br />

„©z<br />

º_x<br />

¦f ¦G<br />

¦K€ `<br />

>])<br />

°<br />

;<br />

¦ 4 <br />

1<br />

s]>])<br />

¦ì¦<br />

m ¦ 1<br />

¦f<br />

m ¦ 4<br />

>^)<br />

> 9 )<br />

<strong>and</strong> scaling transformations (3.108) - (3.109):<br />

(C.44)<br />

¾ r<br />

" )<br />

r“<br />

V ã ¾<br />

ã <br />

V<br />

A F}É<br />

(C.45)<br />

A F}É<br />

9 ¾ r ¾ r<br />

r ¾ r<br />

9 ¾<br />

>\)<br />

>^)<br />

; ><br />

(C.47)<br />

¾ r<br />

r ¾ r<br />

r<br />

r `<br />

Jr<br />

9 ¾<br />

>\)<br />

>^)<br />

> 9 ¾ r ¾ r<br />

; >


)<br />

r<br />

F<br />

F<br />

<br />

<br />

F<br />

r V<br />

F<br />

V<br />

@<br />

É<br />

<br />

r<br />

<strong>and</strong> A F}É<br />

dividing by<br />

@ r <br />

) ã<br />

A!<br />

><br />

@<br />

)<br />

F<br />

A<br />

F<br />

T<br />

F<br />

r<br />

F<br />

ã ¾<br />

F<br />

r ã ¾<br />

><br />

F<br />

F<br />

@<br />

F<br />

K<br />

F<br />

F<br />

r<br />

) ã<br />

F<br />

)<br />

,<br />

><br />

r ã ¾ r<br />

F<br />

><br />

)<br />

<br />

<br />

V<br />

V<br />

r<br />

V<br />

F<br />

1<br />

r ã ¾ r<br />

r<br />

@<br />

<br />

><br />

r<br />

V<br />

><br />

¾<br />

F<br />

)<br />

r<br />

F<br />

r<br />

V<br />

4<br />

r V<br />

@ r ã ¾ r ã ¾ r<br />

F}É A<br />

@<br />

A<br />

<br />

r ã ¾<br />

F<br />

F<br />

ã ¾<br />

F<br />

<br />

F<br />

F<br />

F<br />

><br />

F<br />

F<br />

r<br />

F<br />

F<br />

r<br />

><br />

F<br />

><br />

`<br />

><br />

r<br />

)<br />

<br />

r<br />

V<br />

ã<br />

V<br />

V<br />

F<br />

r<br />

@ ã ¾ r ã<br />

A!<br />

r<br />

<br />

r<br />

V<br />

ã<br />

F<br />

F<br />

`<br />

¤<br />

@<br />

¤<br />

@<br />

x<br />

x<br />

ã<br />

F<br />

V<br />

F<br />

r<br />

ã<br />

F<br />

F<br />

<br />

V V r<br />

)<br />

V<br />

ã<br />

F<br />

¤<br />

x<br />

175<br />

substituting dimensionless variables from (C.44), (C.45), we obtain:<br />

<br />

) ã<br />

ã ¾<br />

r ã ¾ r<br />

A F}É<br />

) ã<br />

) ã<br />

) ã<br />

) ã<br />

(C.48)<br />

9 )<br />

A F}É<br />

9 )<br />

A F}É<br />

` )<br />

) ã<br />

) ã<br />

XH><br />

ã ¾<br />

ã ¾<br />

¾ r ã ¾ r ã<br />

ã ¾<br />

Using the definition of the Reynolds number (2.110), we have:<br />

<br />

) ã<br />

r ã ¾ r<br />

A!<br />

ã<br />

ã ¾ r ã<br />

A"<br />

¾ r<br />

ã ¾<br />

) >äã<br />

ã ¾<br />

X> ã<br />

) ã<br />

) >çã<br />

) ã<br />

ã ¾<br />

) ã<br />

Considering the limit<br />

we have:<br />

A!<br />

<strong>and</strong> using the definition of Froude number (2.111),<br />

) ã<br />

(C.49)<br />

¾ r ã ¾ r ã<br />

ã ¾<br />

A"<br />

b<br />

<br />

) ã<br />

) >çã<br />

X> ã<br />

) ã<br />

) >çã<br />

) ã<br />

) ã<br />

Similarly, introduce the dimensionless variables into the momentum equation for<br />

, (C.47):<br />

<br />

ã ) r<br />

ã ¾<br />

r ã<br />

@<br />

F}É A<br />

@ r ã )<br />

F}É A<br />

@ r ã )<br />

F}É A<br />

¾ r<br />

) ã<br />

) ã<br />

r `<br />

¾ r<br />

ã<br />

ã<br />

:<br />

¾ r<br />

ã ¾ r<br />

r `<br />

@ r<br />

F}É A<br />

ã<br />

ã<br />

) ã<br />

In the limit A!<br />

, we have:<br />

(C.50)<br />

b<br />

T<br />

¾ r<br />

X ã


T<br />

V<br />

s )<br />

<br />

><br />

V<br />

><br />

V<br />

P<br />

–<br />

9 P<br />

9 l<br />

–<br />

–<br />

P<br />

<br />

ã 1<br />

) s]>äã<br />

) 9<br />

s –<br />

P<br />

><br />

) 9<br />

r<br />

V<br />

–<br />

P<br />

ã s<br />

F<br />

V<br />

<br />

V V<br />

)<br />

V<br />

<br />

F F<br />

F<br />

ã s<br />

r F<br />

V<br />

<br />

F<br />

s<br />

F<br />

1<br />

V<br />

¦<br />

s<br />

)<br />

)<br />

s<br />

r<br />

F<br />

<br />

F<br />

)<br />

F<br />

><br />

<br />

F<br />

<br />

F F<br />

V<br />

<br />

F<br />

F<br />

ã r<br />

s<br />

F<br />

l<br />

)<br />

ã r<br />

s<br />

rq<br />

F<br />

F<br />

rq<br />

F<br />

r<br />

s<br />

4<br />

F F<br />

r<br />

s<br />

F<br />

)<br />

– ;<br />

P<br />

><br />

°<br />

– ;<br />

1<br />

)<br />

°<br />

s<br />

– ;<br />

°<br />

<br />

F<br />

V<br />

)<br />

P<br />

F<br />

F<br />

V<br />

<br />

r V<br />

F<br />

1<br />

s<br />

1<br />

)<br />

P<br />

r<br />

)<br />

r V<br />

ã 1<br />

s<br />

><br />

1<br />

s<br />

<br />

FðF<br />

F<br />

r<br />

ã 1<br />

s<br />

><br />

<br />

FðF<br />

)<br />

><br />

<br />

FðF<br />

<br />

FðF<br />

1<br />

)<br />

@ 1<br />

A"<br />

@ 1<br />

A!<br />

4<br />

F<br />

rq<br />

r<br />

4<br />

F<br />

rq<br />

r<br />

4<br />

F<br />

<br />

<br />

F<br />

176 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

This means that the pressure only changes in the direction of<br />

Now consider the energy equation (C.44):<br />

.<br />

1<br />

sà>])<br />

°<br />

s ;<br />

¦ 4 <br />

¦ì¦<br />

m ¦ 1<br />

¦f<br />

m ¦ 4<br />

> 9 )<br />

>^)<br />

Writing it out explicitly, we have:<br />

1<br />

sà>^)<br />

°<br />

;9<br />

r 4 <br />

rðr 4<br />

>])<br />

>]s<br />

rq<br />

r<br />

rq<br />

(C.51)<br />

>C)<br />

>])<br />

>^)<br />

>])<br />

r 1<br />

rq<br />

r 4<br />

rq r 1<br />

rq r<br />

rq r 4<br />

>C)<br />

>])<br />

>^)<br />

>])<br />

Multiplying it by 9ú: –<br />

, <strong>and</strong> rearranging the RHS, we obtain:<br />

1<br />

s]>])<br />

r 4 <br />

rðr 4<br />

>]s<br />

(C.52)<br />

P<br />

R1<br />

¦f ¦ 4<br />

r<br />

rq<br />

F<br />

r 4<br />

rq 4<br />

F<br />

>])<br />

¦K ¦G<br />

where )<br />

(C.44) <strong>and</strong> (C.45), we obtain:<br />

is zero by continuity (2.4). Transferring, to dimensionless variables<br />

rT<br />

ã s >^A!<br />

r 4 <br />

rðr 4<br />

(C.53)<br />

) >çã<br />

— l<br />

r<br />

r 4<br />

) ã<br />

) ã<br />

) ã<br />

) ã<br />

Multiplying the latter by<br />

(2.110), we have:<br />

: s<br />

<strong>and</strong> using the definition of the Reynolds number<br />

>^A!<br />

1<br />

<br />

sà>çã ) ã<br />

r“<br />

rðr 4<br />

) >äã<br />

(C.54)<br />

ã s >\A"<br />

— l<br />

r<br />

r 4<br />

) ã<br />

) ã<br />

) ã<br />

) ã<br />

>^A!<br />

1


@<br />

><br />

ã s<br />

F<br />

<br />

F<br />

ã s<br />

F<br />

l<br />

F<br />

)<br />

F<br />

ã<br />

<br />

F<br />

ã s<br />

F<br />

T<br />

) ã<br />

<br />

F<br />

rq<br />

F<br />

ã r<br />

s<br />

rq<br />

F<br />

)<br />

ã r<br />

s<br />

><br />

ã r<br />

s<br />

1<br />

@<br />

qº(x A!<br />

F<br />

F<br />

s ã<br />

r<br />

s ã<br />

r<br />

<br />

FðF<br />

><br />

ã 1<br />

s<br />

><br />

<br />

FðF<br />

><br />

1<br />

@ 1<br />

A!<br />

) ã r<br />

A<br />

><br />

s ã<br />

1<br />

4<br />

r<br />

F<br />

F<br />

rq<br />

r<br />

4<br />

F<br />

r<br />

<br />

¢<br />

)<br />

r<br />

V<br />

P<br />

s<br />

V<br />

177<br />

using the definition of Pr<strong>and</strong>tl (1.30) <strong>and</strong> Eckert numbers „§z (3.114),<br />

we obtain:<br />

: –<br />

,<br />

<br />

s]>çã ) ã<br />

r“<br />

rðr 4<br />

ã s >\A"<br />

(C.55)<br />

) >çã<br />

r<br />

r 4<br />

— l<br />

) ã<br />

) ã<br />

Collecting the terms with the same powers of A#<br />

A! q„Œz<br />

we have:<br />

) ã<br />

) ã<br />

>\A"<br />

1<br />

r“<br />

rðr<br />

,<br />

<br />

s]>çã ) ã<br />

>çã<br />

r 4<br />

A! qº_x<br />

rq<br />

º(x<br />

r<br />

) ã<br />

<strong>and</strong> considering the limit A#<br />

A! 5„©z<br />

, we obtain:<br />

„Œz<br />

„Œz<br />

r 4<br />

rðr<br />

) ã<br />

b<br />

<br />

s]>çã ) ã<br />

r<br />

) >«ã<br />

„©z<br />

º_x


¤<br />

¡©¦<br />

P<br />

j<br />

„<br />

„<br />

j<br />

P<br />

¦<br />

¤<br />

><br />

<br />

Ä j<br />

j<br />

P<br />

æ<br />

Ä<br />

P<br />

Ä<br />

¤<br />

¤<br />

P<br />

½<br />

Ä<br />

æ<br />

½<br />

j<br />

¤<br />

<br />

<br />

T<br />

¤<br />

P<br />

j<br />

¦<br />

'<br />

j<br />

P<br />

¡<br />

<br />

j<br />

<br />

T<br />

½<br />

'<br />

¡<br />

178 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

Chapter A<br />

Problem A.4.1: Check tensor expressions<br />

Check if the following Cartesian tensor expressions violate tensor rules:<br />

¡©¦/ ¤<br />

Answer: term (1): ik = free, term (2): pk=free<br />

P æ<br />

jê><br />

½ <br />

Answer: (1): ijq=free (2): p=free (3): kp=free<br />

P æ<br />

æÉæ<br />

jÛ><br />

¦/<br />

¡“Ù¤<br />

¦/¡©¦M¤Œ ¦/É<br />

`à½<br />

Answer: (1): i=free (2): none, (3): i=free, j = tripple occurrence<br />

Problem A.4.2: Construct tensor expression<br />

¡t¦/ ¤§¦/ ½ ¦7<br />

j


£<br />

¦<br />

¦<br />

j<br />

j<br />

¦<br />

><br />

£<br />

¦<br />

£<br />

£<br />

£<br />

¦<br />

j<br />

<br />

¦<br />

<br />

<br />

j<br />

j<br />

<br />

<br />

<br />

<br />

<br />

¦<br />

j<br />

¦<br />

<br />

¦<br />

j<br />

j<br />

<br />

¦<br />

><br />

£<br />

<br />

<br />

<br />

><br />

£<br />

Problem A.4.3: Cartesian identity<br />

Prove identity (A.15).<br />

Proof<br />

Integrating (A.5) in the case of constant transformation marix coefficients,<br />

we have:<br />

179<br />

(C.56)<br />

¦<br />

j<br />

£<br />

>A¦<br />

where the transformation matrix is given by (A.4):<br />

ã<br />

(C.57)<br />

¦<br />

j<br />

¢ Y ã<br />

Y j<br />

By the definition of the Cartesian coordinates (C.58) we have:<br />

(C.58)<br />

ß ¦/<br />

j ¦ £ j £<br />

Y ã<br />

¦ Y ã<br />

<br />

¦<br />

Let’s multiply the transformation rule (C.56)<br />

<br />

by . Then we get:<br />

£<br />

Y <br />

Y <br />

£<br />

ß <br />

¦<br />

<br />

¦<br />

£ <br />

¦<br />

A¦<br />

¦<br />

A¦<br />

¦<br />

A¦<br />

ã<br />

j<br />

Differentiation this ã over<br />

, we have:<br />

¦<br />

üY <br />

Now rename index¨into ° :<br />

Y ã<br />

Comparing this with (C.57), we have<br />

üY <br />

Y ã<br />

which proves (A.15).<br />

Y ã<br />

Y <br />

<br />

Y <br />

Y ã


(C.60)<br />

~<br />

<br />

j<br />

p<br />

j<br />

P<br />

¤<br />

<br />

P<br />

ß<br />

j<br />

j<br />

p5¦<br />

Pæ<br />

Ä<br />

P<br />

j<br />

~<br />

ß<br />

j<br />

æ ¡ P<br />

æ<br />

Ä<br />

j<br />

p<br />

<br />

æ<br />

j<br />

æ<br />

æ<br />

Ä<br />

j<br />

æ<br />

P<br />

`<br />

j<br />

ß<br />

j<br />

j<br />

¤<br />

j<br />

æ<br />

æ<br />

Ä<br />

Ä<br />

P<br />

Ä<br />

P<br />

æ<br />

ß<br />

Ä<br />

j<br />

j<br />

P<br />

Ä<br />

Ä<br />

æ<br />

æ<br />

æ<br />

4<br />

æ<br />

ß<br />

j<br />

P<br />

4Û`<br />

Ä~<br />

'<br />

j<br />

j<br />

j<br />

¤<br />

¤<br />

ß ¡<br />

j<br />

P<br />

P<br />

¤<br />

P<br />

Ä<br />

Ä<br />

P<br />

¤ P<br />

j<br />

Ä<br />

æ<br />

æ<br />

P<br />

æ<br />

180 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

Problem A.4.4: Tensor identity<br />

Using the tensor identity:<br />

(C.59)<br />

p5¦/<br />

ß <br />

`]ß <br />

prove the vector identity (A.30):<br />

1~<br />

Í ¤ 4 ¡ ¤ ¡‡ž Í 1~ ~ ~ ~ 1~<br />

¡‡ž ¤ 4 ~<br />

Proof<br />

Applying (A.26) twice to the RHS of (C.60), we have:<br />

Í ¡ ¤ Í ~ 1~<br />

pm¦/<br />

¡p<br />

pm¦/<br />

æ ¡“ ¤ P<br />

From (A.24) it follows that p5¦/ j<br />

. Then we have:<br />

P æ<br />

` p5¦<br />

kp<br />

¦7<br />

(C.61)<br />

¡¤<br />

æ P<br />

Now rename the dummy indexes: b # °<br />

, so that the expression<br />

looks like one in (A.29):<br />

p5¦/<br />

¡¤<br />

¼p<br />

¦/p<br />

P æ<br />

# b<br />

¨‰'¨<br />

b °<br />

p5¦<br />

4 ¡<br />

1 p5¦/<br />

4 ¡<br />

Pæ<br />

(C.62)<br />

1 ß <br />

`àß <br />

ß <br />

`àß <br />

Using (A.10), <strong>and</strong> since ¡â邏 is the same as ¡§¦“朗¦ the latter can be<br />

rewritten as:<br />

j¡Ä<br />

(C.63)<br />

¼¤k ¡<br />

,¡<br />

P


¤ 1~ ~<br />

F Y<br />

ã Y<br />

<br />

r<br />

r<br />

¡ž<br />

Ä~<br />

<br />

®<br />

<br />

Y<br />

Î Y<br />

<br />

Y<br />

¢<br />

Y<br />

Y ® <br />

¢<br />

Y<br />

üY ® r<br />

Î Y<br />

<br />

<br />

<br />

¢<br />

¦<br />

4a`<br />

Ä~<br />

¦<br />

+<br />

¢<br />

¢<br />

<br />

Î<br />

›<br />

<br />

a/cedWÎ<br />

Î<br />

GF<br />

¢<br />

®yF£¢a/cedWÎ<br />

Y + <br />

+ Y<br />

¢<br />

<br />

Î<br />

@<br />

¦<br />

¦<br />

181<br />

which is the same as<br />

¡‡ž ¤ 4 ~<br />

1~<br />

Problem A.4.5: Metric tensor in cylindrical coordinates.<br />

%§¢<br />

ã ',Î'å›ð- Cylindrical coordinate system (C.64) is given by the following<br />

transformation rules to a ('*®å',+- Cartesian coordinate system, :<br />

ä%<br />

£¢a/cedWÎ<br />

£¢defÆ<br />

Obtain the components of the metric tensor (A.40) <strong>and</strong> its inverse<br />

(A.38) in cylindrical coordinates.<br />

J¦7<br />

Solution:<br />

First compute the derivatives of<br />

'*Î'W›ñ- :<br />

¦/<br />

ä%<br />

Ú%§¢<br />

with respect to ã <br />

('*®å',+-<br />

œx<br />

F<br />

Y <br />

defÆ<br />

Y ã<br />

®


^<br />

<br />

®<br />

<br />

+<br />

ë<br />

<br />

r @<br />

@<br />

@<br />

@<br />

182 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

xñx<br />

œxqœxa>^® xq® x<br />

F³F<br />

GFWGFa>^®”Fœ®yF£¢<br />

HZH<br />

xñx <br />

F³F<br />

@ r ¢<br />

HZH<br />

Problem A.4.6: Metric tensor in curvilinear coordinates<br />

Using Mathematica, write a procedure to compute metric tensor in curvilinear<br />

coordinate system, <strong>and</strong> use it to obtain the components of metric tensor, ,<br />

(A.40) <strong>and</strong> its conjugate, , (A.38) in spherical coordinate system ( ' ë_',Î ):<br />

¢<br />

¼¢defÆ<br />

ÎŒa/cydåë<br />

(C.65)<br />

ÎŒdeKÆ<br />

¢arcedåÎ<br />

¢defÆ<br />

Solution with Mathematica<br />

NX = 3<br />

(* Curvilinear cooridnate system *)<br />

Y = Array[,NX] (* Spherical coordinate system *)<br />

Y[[1]] = r; (* radius *)<br />

Y[[2]] = th; (* angle theta *)<br />

Y[[3]] = phi; (* angle phi *)<br />

(* Cartesian coordinate system *)<br />

X = Array[,NX]<br />

X[[1]] = r Sin[th] Cos[phi];<br />

X[[2]] = r Sin[th] Sin[phi];<br />

X[[3]] = r Cos[th];<br />

(* Compute the Jacobian: dXi/dYj *)


^2×%Ž%<br />

T<br />

'<br />

T<br />

'<br />

T<br />

T<br />

'<br />

'<br />

¢<br />

r<br />

'<br />

r<br />

é ¢<br />

T<br />

'<br />

T<br />

'<br />

T<br />

'<br />

T<br />

'<br />

r<br />

1 ¢ 4<br />

rdefÆ<br />

r Î<br />

Î<br />

4<br />

183<br />

J = Array[,{NX,NX}]<br />

Do[<br />

J [[i,j]] = D[X[[i]],Y[[j]]],<br />

{j,1,NX},{i,1,NX}<br />

]<br />

(* Covariant Metric tensor *)<br />

g = Array[,{NX,NX}] (* covariant *)<br />

Do[<br />

g [[i,j]] = Sum[J[[k,i]] J[[k,j]],{k,NX}],<br />

{j,1,NX},{i,1,NX}<br />

];<br />

g=Simplify[g]<br />

(* Contravariant metric tensor *)<br />

g1 =Array[,{NX,NX}]<br />

g1=Inverse[g]<br />

With the result:<br />

âä%Ž%<br />

% T<br />

% T<br />

@‰'<br />

-Q'<br />

-Q'<br />

-Ž-<br />

% T<br />

% T<br />

'ard=a1<br />

@‰'<br />

-Q'<br />

-Q'<br />

¢ r -Ž-<br />

Problem A.4.7: Christoffel’s symbols with Mathematica<br />

Using the Mathematica package, write the routines to compute Christoffel’s<br />

symbols<br />

Solution<br />

(************* File g.m *************<br />

The metric tensor<br />

<strong>and</strong> Christoffel symbols<br />

*************************************)


184 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

DIM = 3<br />

(*<br />

The metric tensor<br />

*)<br />

g = Array[,{DIM,DIM}] (* covariant *)<br />

g1 =Array[,{DIM,DIM}] (* contravariant *)<br />

Do[<br />

g [[i,j]] = 0;<br />

g1[[i,j]] = 0<br />

,<br />

{j,1,DIM},{i,1,DIM}<br />

]<br />

(*<br />

Cylindrical coordinates<br />

*)<br />

Z=Array[,DIM]<br />

Z[[1]] = r<br />

Z[[2]] = th<br />

Z[[3]] = z<br />

g [[1,1]] = 1<br />

g [[2,2]] = rˆ2<br />

g [[3,3]] = 1<br />

g1[[1,1]] = 1<br />

g1[[2,2]] = 1/rˆ2<br />

g1[[3,3]] = 1<br />

(*<br />

Christoffel symbols of the first <strong>and</strong> second type<br />

*)<br />

Cr1 = Array[,{DIM,DIM,DIM}]<br />

Cr2 = Array[,{DIM,DIM,DIM}]<br />

Do[<br />

Cr1[[i,j,k]] = 1/2<br />

(<br />

D[ g [[i,k]], Z[[j]] ]<br />

+ D[ g [[j,k]], Z[[i]] ]<br />

- D[ g [[i,j]], Z[[k]] ]<br />

),<br />

{k,DIM},{j,DIM},{i,DIM}<br />

]<br />

Do[<br />

Cr2[[l,i,j]] =


185<br />

]<br />

Sum[<br />

g1[[l,k]] Cr1[[i,j,k]],<br />

{k,DIM}<br />

],<br />

{j,DIM},{i,DIM},{l,DIM}<br />

Problem A.4.8: Covariant differentiation with Mathematica<br />

Using the Mathematica package, write the routines for covariant differentiation<br />

of tensors up to second order.<br />

solution<br />

(************** File D.m *******************<br />

Rules of covariant differentiation<br />

********************************************)<br />

(*<br />

B.Spain<br />

Tensor Calculus, 1965<br />

Eq.(22.2)<br />

*)<br />

D1[N_,A_,k_,X_,j_]:=<br />

(*<br />

Computes covariant derivative<br />

of a mixed tensor of second order<br />

with index k - covariant (upper)<br />

*)<br />

Module[<br />

{i,s},<br />

s = Sum[Cr2[[k,i,j]] A[[i]],{i,N}];<br />

D[A[[k]],X[[j]]] + s<br />

]<br />

Dl1[N_,A_,l_,X_,t_]:=<br />

(*<br />

Computes covariant derivative<br />

of a mixed tensor of second order


186 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

with index l - covariant (lower)<br />

*)<br />

Module[<br />

{s,r},<br />

s =Sum[Cr2[[r,l,t]] A[[r]],{r,N}];<br />

D[A[[l]],X[[t]]] - s<br />

]<br />

D1l1[N_,A_,m_,l_,X_,t_]:=<br />

(*<br />

Computes covariant derivative<br />

of a mixed tensor of second order<br />

with index m - contravariant (upper) <strong>and</strong><br />

index l - covariant (lower)<br />

*)<br />

Module[<br />

{s1,s2,r},<br />

s1 =Sum[Cr2[[m,r,t]] A[[r,l]],{r,N}];<br />

s2 =Sum[Cr2[[r,l,t]] A[[m,r]],{r,N}];<br />

D[A[[m,l]],X[[t]]] + s1 - s2<br />

]<br />

D2[N_,A_,i_,j_,X_,n_]:=<br />

(*<br />

Computes covariant derivative<br />

of second order tensor with<br />

both m <strong>and</strong> l contravariant (upper)<br />

indexes<br />

B.Spain<br />

Tensor Calculus, 1965<br />

Eq.(23.3)<br />

*)<br />

Module[<br />

{s1,s2,k},<br />

s1 =Sum[Cr2[[i,k,n]] A[[k,j]],{k,N}];<br />

s2 =Sum[Cr2[[j,k,n]] A[[i,k]],{k,N}];<br />

D[A[[i,j]],X[[n]]] + s1 + s2<br />

]<br />

D2l1[N_,A_,i_,j_,k_,X_,n_]:=<br />

(*<br />

Computes covariant derivative<br />

of third order tensor with<br />

i <strong>and</strong> j contravariant (upper)


187<br />

<strong>and</strong> k contravariant (lower)<br />

indexes<br />

B.Spain<br />

Tensor Calculus, 1965<br />

Eq.(23.3)<br />

*)<br />

Module[<br />

{s1,s2,s3,m},<br />

s1 =Sum[Cr2[[i,m,n]] A[[m,j,k]],{m,N}];<br />

s2 =Sum[Cr2[[j,m,n]] A[[i,m,k]],{m,N}];<br />

s3 =Sum[Cr2[[m,k,n]] A[[i,j,m]],{m,N}];<br />

D[A[[i,j,k]],X[[n]]] + s1 + s2 - s3<br />

]<br />

D4l1[N_,A_,i1_,i2_,i3_,i4_,i5,X_,i6_]:=<br />

(*<br />

Computes covariant derivative<br />

of 5 order tensor with<br />

4 first indexes contravariant (upper)<br />

<strong>and</strong> the last one contravariant (lower)<br />

B.Spain<br />

Tensor Calculus, 1965<br />

Eq.(23.3)<br />

*)<br />

Module[<br />

{k,s1,s2,s3,s4,s5},<br />

s1= Sum[Cr2[[i1,k,n]] A[[k,i2,i3,i4,i5]],{k,N}];<br />

s2= Sum[Cr2[[i2,k,n]] A[[i1,k,i3,i4,i5]],{k,N}];<br />

s3= Sum[Cr2[[i3,k,n]] A[[i1,i2,k,i4,i5]],{k,N}];<br />

s4= Sum[Cr2[[i4,k,n]] A[[i1,i2,i3,k,i5]],{k,N}];<br />

s5=-Sum[Cr2[[k,i5,n]] A[[i1,i2,i3,i4,k]],{k,N}];<br />

D[A[[i1,i2,i3,i4,i5]],X[[i6]]]+s1+s2+s3+s4+s5<br />

]<br />

Problem A.4.9: Divergence of a vector in curvilinear coordinates<br />

Using the Mathematica package <strong>and</strong> the solution of Problem A.4.8, write the<br />

routines for computing divergence of a vector in curvilinear coordinates.<br />

Solution


188 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

Using the algorithms of covariant differentiation developed in Problem A.4.8<br />

we have:<br />


189<br />

(\cite[5.102-5.110]{SyScTC69})<br />

*)<br />

V = Array[,NX]<br />

Do[<br />

V[[i]] = PowerExp<strong>and</strong>[V0[[i]]/g[[i,i]]ˆ(1/2)],<br />

{i,1,NX}<br />

]<br />

(*<br />

Transform vectors<br />

as first order contravariant tensors<br />

*)<br />

U = Array[,NX]<br />

SetAttributes[RV1,HoldAll]<br />

RV1[NX,V,U]<br />

(*<br />

Compute first covariant derivatives<br />

of vectors<br />

*)<br />

DV = Array[,{NX,NX}];<br />

Do[<br />

DV[[i,j]] = D1[NX,V,i,Y,j],<br />

{j,1,NX},{i,1,NX}<br />

]<br />

(* Divergence *)<br />

div=0<br />

Do[<br />

div=div+DV[[i,i]],<br />

{i,NX}<br />

]<br />

div0 = div/.th->0<br />

Problem A.4.10: Laplacian in curvilinear coordinates<br />

Using the Mathematica package, write the routines for computing Laplacian<br />

in curvilinear coordinates.<br />

solution<br />

Using the algorithms of covariant differentiation developed in Problem A.4.8


190 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

we have:<br />


¡<br />

¤©¦<br />

æ P<br />

(C.68) „<br />

„<br />

Ä<br />

j<br />

¤<br />

¦<br />

æ<br />

¦<br />

j<br />

¦<br />

j<br />

`<br />

P<br />

P<br />

><br />

><br />

¤<br />

j<br />

½<br />

½<br />

Éj<br />

P<br />

¦<br />

<br />

Ä<br />

j<br />

P<br />

Pj<br />

Ä<br />

æ<br />

j<br />

<br />

³<br />

j<br />

Ä<br />

æ<br />

<br />

æ<br />

<br />

æ<br />

<br />

<br />

<br />

½<br />

j<br />

<br />

<br />

Ä j æ<br />

¦ÉP<br />

¦<br />

jÉP<br />

¤<br />

P<br />

¦ì³<br />

æ<br />

]<br />

DDQ = Array[,{NX,NX}];<br />

Do[<br />

DDQ[[i,j]] = Sum[DDP[[k,l]] J1[[k,i]] J1[[l,j]],{k,NX},{l,NX}],<br />

{i,1,NX},{j,1,NX}<br />

]<br />

(* Laplacian *)<br />

(*** lap=lap+Sum[g[[i,j]]*Dl1[NX,DS,j,Y,i],{i,1,NX},{j,1,NX}],*)<br />

lap=Sum[DDQ[[i,i]],{i,NX}]<br />

lap0=lap/.th->0<br />

191<br />

Problem A.4.11: Invariant expressions<br />

not:<br />

Check if any of these tensor expressions are invariant, <strong>and</strong> correct them if<br />

(C.66)<br />

¦<br />

<br />

Ä j<br />

<br />

³<br />

j<br />

¡§¦M¤<br />

½ <br />

(C.67)<br />

¦7<br />

<br />

j<br />

î¦<br />

j ¡<br />

¦<br />

<br />

<br />

Answers:<br />

A corrected form of (C.66) is:<br />

¡§¦³¤<br />

½ <br />

Equality (C.68) requires no corrections. A corrected form of (C.68) is:<br />

¦M¤<br />

½<br />

¦<br />

Since there are two combinations for an invariant combination of dummy<br />

indexes (Corollary A.3.5), there can be several different invariant expressions.


¡<br />

¦<br />

¡ X<br />

<br />

¦<br />

¡<br />

<br />

¦<br />

j<br />

<br />

X<br />

<br />

¦<br />

¤<br />

j<br />

<br />

<br />

¦<br />

j<br />

<br />

¦<br />

j<br />

¡<br />

<br />

<br />

¤<br />

¦<br />

j<br />

¤<br />

j<br />

<br />

¦<br />

<br />

192 APPENDIX C. SOLUTIONS TO PROBLEMS<br />

Problem A.4.12: Contraction invariance<br />

¦<br />

¤©¦¼¡§¦M¤<br />

Prove that ¡<br />

Proof<br />

, <strong>and</strong> both are invariant, while ¡§¦M¤©¦ is not.<br />

Using the operation of rising/lowering indexes (A.42), (A.43), we have<br />

¤§¦$‡<br />

¡“ ¦<br />

¤ j <br />

j ß ¡“Ù¤<br />

¡“Ù¤ j ¼¡¤<br />

¦7<br />

¦/<br />

which proves that both forms have the same values. If we now consider the first<br />

form then:<br />

¤§¦$<br />

¦X<br />

Y Wj<br />

ß <br />

£¡<br />

¤Œ¡<br />

¤©¦<br />

Y <br />

which proves the point.<br />

YX<br />

Consider now :<br />

¡§¦³¤§¦ YX<br />

¤§¦$ïY <br />

¡QY Wj<br />

j<br />

which can not be reduced further <strong>and</strong>, therefore is not invariant, since it has a<br />

different form from the LHS.<br />

YX YX<br />

¡§¦X


Appendix D<br />

Midterm Exam Topics: Laminar Flow<br />

Solutions<br />

Items surrounded in brakets: [. . . ] will not be available during the exam. Other<br />

items will be available.<br />

1. Formulate the equations for incompressible flow [(3.1)], [(3.2)], [(3.3)].<br />

2. Give the definition of hydrostatic pressure [(3.4)].<br />

3. Write the expression for the viscous stress tensor for a Newtonian incompressible<br />

fluid [(3.5)].<br />

4. Give the definition of a laminar flow [(Definition 3.1.1)], <strong>and</strong> specify the conditions<br />

when it may occur.<br />

5. Formulate the equations for momentum <strong>and</strong> energy for the incompressible<br />

flow between parallel plates: [(3.7)], [(3.8)].<br />

6. Obtain the solution for velocity <strong>and</strong> shear stress for the flow between parallel<br />

plates: [(3.9)], [(3.10)].<br />

7. Solve Problem 3.7.3.<br />

8. Give the definition of a friction coefficient, ÄŒõ , <strong>and</strong> express it in terms of<br />

Reynolds number for the steady flow between parallel plates: [(3.11)].<br />

9. Give the definition of a Poiseuille number, <strong>and</strong> obtain its value for the flow<br />

between parallel plates: [(3.12)].<br />

10. Give definition of a Brickman number [(3.13)].<br />

193


194 APPENDIX D. MIDTERM EXAM TOPICS: LAMINAR FLOW SOLUTIONS<br />

11. Obtain equation of motion for axially moving concentric cylinders: [(3.14)]<br />

on the basis of general NS equation in cylindrical coordinates.<br />

12. Obtain the solution [(3.16)] to equation (3.14).<br />

13. Show that the problem of pulling an infinite rod does not have a steady-state<br />

solution [Remark 3.2.1].<br />

14. Obtain equations of motion for axially moving concentric cylinders: [(3.17)],<br />

[(3.18)] on the basis of general NS equation in cylindrical coordinates (Problem<br />

3.7.5).<br />

15. Obtain the solution for the flow between axially moving concentric cylinders,<br />

[(3.20)], on the basis of equation (3.18) <strong>and</strong> the appropriate boundary conditions.<br />

16. Obtain solutions for the flow inside <strong>and</strong> outside of a rotating cylinder: Remarks<br />

[3.2.2] <strong>and</strong> [3.2.3].<br />

17. Obtain the pressure distribution in a flow outside a rotating cylinder: [(3.23)]<br />

on the basis of the solution (3.22) <strong>and</strong> equation <strong>and</strong> the appropriate momentum<br />

equation [(3.17)].<br />

18. Solve problem 3.7.6.<br />

19. Derive the equation for axial velocity in a fully developed flow region [(3.25)]<br />

using the appropriate assumptions <strong>and</strong> equation (3.2). Write a non-dimensional<br />

formulation of the boundary value problem [(3.26)].<br />

20. Formulate a boundary value problem for a Poiseuille flow in a circular duct<br />

[(3.27)] on the basis of a NS equation in cylindrical coordinates, <strong>and</strong> obtain<br />

the Poiseuille solution [(3.28)].<br />

21. For a Poiseuille flow through a duct (3.28) compute the volumetric flow rate<br />

[(3.28)], the wall shear stress [(3.30)], the skin friction coefficient [(3.31)]<br />

<strong>and</strong> Poiseuille number [(3.32)].<br />

22. For combined Couette-Poiseuille flows formulate the equation of motion<br />

[(3.33)], obtain the solution [(3.36)], <strong>and</strong> formulate the criterion of separation<br />

[(3.38)].<br />

23. Solve problem 3.7.7.<br />

24. Give a definition of a hydraulic diameter [(3.39)].


4<br />

<br />

<br />

4<br />

<br />

r<br />

l<br />

<br />

Ç<br />

4 :nm<br />

<br />

4<br />

195<br />

25. Formulate the equation for the fluid oscillating above an infinite plate [(3.44)],<br />

<strong>and</strong> obtain the solution for the case of oscillating plate [(3.47)] <strong>and</strong> an oscil-<br />

. @<br />

lating fluid [(3.48)]. Hint: Use the identity: # F}É<br />

, where #­¢ m<br />

26. Obtain the solution for an unsteady flow between two infinite plates<br />

1<br />

[(3.65)].<br />

Hint: relation:x\"defÆ<br />

Use integral . "a/cyd1<br />

<br />

deKÆ 1<br />

1 #<br />

l<br />

`<br />

> @<br />

4a`<br />

R <br />

27. Formulate the assumptions of creeping flow, <strong>and</strong> show how to obtain the<br />

Laplace equations for pressure [(3.67)] <strong>and</strong> vorticity [(3.68)].<br />

28. Using the expressions for pressure (3.76) <strong>and</strong> shear stress (3.77) of a Stokes<br />

flow around a sphere, compute the total drag force [(3.80)], <strong>and</strong> the drag coefficient<br />

as a function of the Reynolds number [(3.81)]. Hint: Use the integral<br />

.<br />

1arced1<br />

relation:xNdefÆ 1<br />

454 : @<br />

4a`<br />

R <br />

À<<br />

Ô{a/ced1<br />

29. Using the assumptions of the lubrication theory, derive the equation for pressure<br />

distribution in a flow between moving plates with a non-uniform gap<br />

[(3.91)].<br />

30. Derive criteria of validity of equation (3.91): [<br />

›<br />

], [(3.92)].


196 APPENDIX D. MIDTERM EXAM TOPICS: LAMINAR FLOW SOLUTIONS


Appendix E<br />

Final Exam Topics<br />

Items surrounded in brakets: [. . . ] will not be available during the exam. Other<br />

items will be available. Abbreviations used: CFM=”<strong>Concise</strong> <strong>Fluid</strong> <strong>Mechanics</strong>”,<br />

A.Smirnov (http://www.mae.wvu.edu/cfm), VFF=”Viscous <strong>Fluid</strong> Flow”, F.White, 2nd<br />

Edition, McGraw-Hill, 1991.<br />

E.1 Fundamental Laws<br />

1. Definition of substantial derivative <strong>and</strong> derive its expression [(1.7)].<br />

2. Give definitions of strain tensor [(1.9)], vorticity tensor [(1.10)] <strong>and</strong> vorticity<br />

vector [(1.11)].<br />

3. Derive the conservation of mass in explicit form [(2.2)] <strong>and</strong> in substantial<br />

derivative form [(2.3)], <strong>and</strong> for incompressible flow [(2.4)].<br />

4. Define the stream-function in 3D [(2.6)], [2.8)], <strong>and</strong> 2D case [(2.9)]. Describe<br />

its realtion to the mass-flow rate [(2.10)].<br />

5. Derive general equation of momentum [(2.22)] based on the definition of<br />

viscous stress tensor [(2.19)], <strong>and</strong> it’s incompressible limit Navier-Stokes<br />

equation [(2.24)].<br />

6. Derive a vorticity formulation of the incompressible Navier-Stokes equation<br />

[(2.34)].<br />

7. Derive the Poisson equation for pressure for constant density flows [(2.61],<br />

<strong>and</strong> describe the boundary conditions.<br />

197


; R ’<br />

(E.1) "<br />

R<br />

Y<br />

"<br />

Y<br />

~<br />

j<br />

<br />

p5¦<br />

æ P<br />

Ä<br />

X R<br />

" ><br />

R<br />

<br />

P<br />

ß<br />

s<br />

<br />

j<br />

æ<br />

`<br />

Ä<br />

X<br />

¦<br />

æ<br />

ß<br />

j<br />

P<br />

Ä<br />

¦<br />

198 APPENDIX E. FINAL EXAM TOPICS<br />

8. Formulate the energy equation in terms of temperature [(2.74)] <strong>and</strong> enthalpy<br />

[(2.75)]. Explain the meaning of each term.<br />

9. Problem CFM.2.7.5: Using the energy equation (2.77):<br />

ã 6 ¦/<br />

1 °<br />

¦ 4 ¦<br />

m ¦<br />

>^)<br />

<strong>and</strong> momentum equation (2.21):<br />

(E.2)<br />

1 ; )<br />

1 ; )<br />

4 ¦<br />

j<br />

6 ¦<br />

j >Úã<br />

<br />

j<br />

¦ 4<br />

¦<br />

>^)Wj<br />

>\š<br />

derive the strong formulation of the Bernoulli’s equation.<br />

10. Derive the expression for Coriolis forces [(2.93)].<br />

11. Problem CFM.A.4.4: Using tensor identity:<br />

p5¦/<br />

ß <br />

`]ß <br />

prove the vector identity (A.30):<br />

Í ¤ 4 ¡ ¤ ¡‡ž Í 4a` 1~ ~ ~ ~ ~ 1~<br />

¡‡ž ¤ 4 ~<br />

¦<br />

¤§¦$¼¡§¦M¤<br />

12. Problem CFM.A.4.12: Prove that ¡<br />

¡§¦M¤©¦<br />

is not.<br />

, <strong>and</strong> both are invariant, while<br />

1~<br />

E.2 Analytical Solutions<br />

13. Formulate the equations for incompressible flow [(3.1)], [(3.2)], [(3.3)]. Write<br />

the expression for the viscous stress tensor for a Newtonian incompressible<br />

fluid [(3.5)].<br />

14. Formulate the equations for momentum <strong>and</strong> energy for the incompressible<br />

flow between parallel plates: [(3.7)], [(3.8)]. Obtain the solution for velocity<br />

<strong>and</strong> shear stress for the flow between parallel plates: [(3.9)], [(3.10)].


F<br />

<br />

T<br />

¾<br />

¦<br />

)<br />

1 4<br />

®<br />

r<br />

ë<br />

X<br />

<br />

ë<br />

<br />

r<br />

l<br />

E.2. ANALYTICAL SOLUTIONS 199<br />

15. Problem CFM.3.7.3: Consider a wide fluid film of constant ’ thickness, ,<br />

flowing steadily due to the gravity down the inclined plate at Î angle . Find<br />

the velocity )<br />

distribution, , <strong>and</strong> the volumetric flow … rate, . Atmospheric<br />

pressure can be considered constant.<br />

16. Obtain equation of motion for axially moving concentric cylinders: [(3.14)]<br />

on the basis of general NS equation in cylindrical coordinates.<br />

17. Obtain the solution to equation (3.14) for the case of axially moving concentric<br />

cylinders [(3.16)]. Show that the problem of pulling an infinite rod does<br />

not have a steady-state solution [Remark 3.2.1].<br />

18. Problem CFM.3.7.5: Using the assumptions on the Couette velocity profile<br />

between the rotating concentric cylinders (Sec.3.2.3) <strong>and</strong> the expression for<br />

the momentum equation <strong>and</strong> the Laplacian operator in cylindrical coordinates:<br />

x_><br />

¢ >^)GH,)GFH­><br />

¢<br />

)œx5)IF<br />

<br />

)IF­>])Wx5)GF<br />

¼<br />

F`<br />

)GFœ)GFF<br />

)œx<br />

; ¢ > 9 1 ¾<br />

¢ r 4<br />

)GF­><br />

¢ r ` )GF<br />

@<br />

¢<br />

<br />

x<br />

4<br />

x_><br />

@ r ë ¢<br />

F„F_>Þë<br />

1 ¢<br />

Derive equation for )IF, [(3.18)],<br />

rotating concentric cylinders, [(3.20)], on the basis of the appropriate boundary<br />

conditions.<br />

F<br />

F<br />

<strong>and</strong> obtain the solution for the flow between<br />

HH<br />

19. Problem CFM.3.7.6: In the system of two rotating cylinders (Sec.3.2.3) consider<br />

the torque applied to the inner rotating cylinder when the outer cylinder<br />

is (n fixed ). What is the power required to rotate the inner cylinder?<br />

20. Derive the equation for axial velocity in a fully developed flow region [(3.25)]<br />

using the appropriate assumptions <strong>and</strong> the momentum equation (3.2):<br />

¦<br />

Write a non-dimensional formulation of the boundary value problem [(3.26)].<br />

¦f<br />

j<br />

F ã 6 ¦<br />

j ;Ãé<br />

<br />

j<br />

` ;Ãé F º<br />

>])åj)<br />

21. Formulate a boundary value problem for a Poiseuille flow in a circular duct<br />

[(3.27)] on the basis of a NS equation in cylindrical coordinates, <strong>and</strong> obtain<br />

the Poiseuille solution [(3.28)]. Compute the volumetric flow rate [(3.28)], the<br />

wall shear stress [(3.30)], the skin friction coefficient [(3.31)] <strong>and</strong> Poiseuille<br />

number [(3.32)].


Ñ<br />

ˆ<br />

<br />

4<br />

<br />

)<br />

1<br />

®<br />

Ä<br />

<br />

r<br />

®<br />

4<br />

r<br />

<br />

F<br />

<br />

r<br />

l<br />

@<br />

<br />

Ï<br />

I<br />

º<br />

4 :m<br />

<br />

4<br />

200 APPENDIX E. FINAL EXAM TOPICS<br />

22. For combined Couette-Poiseuille flows formulate the equation of motion<br />

[(3.33)], obtain the solution [(3.36)], <strong>and</strong> formulate the criterion of separation<br />

[(3.38)]. Using the obtained solution, consider a viscous fluid<br />

<br />

with<br />

Ï<br />

viscosity<br />

) driven between<br />

¿<br />

two parallel plates<br />

apart<br />

`ÞÑ•° : 1 ¹. 4<br />

(<br />

by an imposed pressure gradient 8 º : R of R<br />

£ :<br />

(Problem 3.7.7).<br />

@‰I<br />

The upper plate is moving velocity<br />

<br />

– ¹:‰.<br />

with . Find the volume flow<br />

rate per 1m of the plates’ width. What pressure gradient will cause the flow<br />

to reverse?<br />

` l<br />

ωÏ<br />

23. Formulate the equation for the fluid oscillating above an infinite plate [(3.44)],<br />

<strong>and</strong> obtain the solution for the case of oscillating plate [(3.47)] <strong>and</strong> an oscil-<br />

. @<br />

lating fluid [(3.48)]. Hint: Use the identity: # F}É<br />

1 #<br />

l<br />

, where #a¢ m<br />

`<br />

24. Obtain the solution for an unsteady flow between two infinite plates<br />

1<br />

[(3.65)].<br />

Hint: relation:x\!defÆ<br />

Use integral . "a/ced1<br />

<br />

defÆ 1<br />

> @<br />

4a`<br />

R <br />

25. Formulate the assumptions of creeping flow, <strong>and</strong> show how to obtain the<br />

Laplace equations for pressure [(3.67)] <strong>and</strong> vorticity [(3.68)].<br />

26. Using the expressions for pressure (3.76) <strong>and</strong> shear stress (3.77) of a Stokes<br />

flow around a sphere, compute the total drag force [(3.80)], <strong>and</strong> the drag coefficient<br />

as a function of the Reynolds number [(3.81)]. Hint: Use the integral<br />

.<br />

1arced1<br />

relation:xÞdefÆ 1<br />

4q4 : @<br />

4a`<br />

R <br />

À <br />

Ôßarced1<br />

27. Using the assumptions of the lubrication theory, derive the equation for pressure<br />

distribution in a flow between moving plates with a non-uniform gap<br />

[(3.91)], <strong>and</strong> derive criteria of its validity: [ ], [(3.92)].<br />

›<br />

Ç<br />

E.3 Boundary Layers<br />

28. Using integral analysis derive the expressions for displacement thickness,<br />

[(3.94)] <strong>and</strong> momentum thickness [(3.95)]. Define the skin friction [(3.96)]<br />

<strong>and</strong> drag, [(3.97)] coefficients, <strong>and</strong> derive their relations to the momentum<br />

thickness [(3.98), (3.99)].<br />

29. Using the parabolic velocity profile:<br />

V<br />

obtain the boundary layer growth rate as a function of [(3.104)], [(3.105)].<br />

>\Ä<br />

®C>\Ä<br />

4


F<br />

ã<br />

<br />

<br />

<br />

–<br />

P<br />

)<br />

F<br />

ã<br />

r<br />

<br />

<br />

‹<br />

¦<br />

s<br />

r<br />

1<br />

)<br />

¦<br />

)<br />

F<br />

F<br />

)<br />

<br />

V<br />

<br />

)<br />

s<br />

¡<br />

<br />

V<br />

¾<br />

r<br />

)<br />

)<br />

r<br />

)<br />

)<br />

X<br />

¦ <br />

> ;<br />

r<br />

<br />

<br />

V<br />

T<br />

¦<br />

X<br />

) ;<br />

r<br />

<br />

V<br />

V<br />

"<br />

¦<br />

E.3. BOUNDARY LAYERS 201<br />

30. Problem CFM.3.7.13: Using equations (3.107) - (3.107):<br />

¦f ¦G<br />

¦f“ 9 )<br />

¦K€ `<br />

>])<br />

°<br />

;<br />

¦ 4 <br />

¦ì¦<br />

m ¦ 1<br />

¦f<br />

m ¦ 4<br />

1<br />

s]>])<br />

> 9 )<br />

>^)<br />

<strong>and</strong> scaling transformations (3.108) - (3.109):<br />

r<br />

¾ r“<br />

r ¾ r/<br />

" )<br />

A F}É<br />

F ã<br />

ã<br />

ã <br />

V<br />

V ã ¾<br />

A F}É<br />

)<br />

r“<br />


p<br />

Ä<br />

202 APPENDIX E. FINAL EXAM TOPICS<br />

35. Problem CFM.3.7.12: A thin equilateral triangle plate with the edge length<br />

of is immersed parallel to a 12m/s stream of air at l T {<br />

<strong>and</strong> 1atm,<br />

as in Fig.C.3. Assuming laminar flow estimate drag of this plate (in N).<br />

l Ç<br />

First give an answer in symbolic form in terms of 8 , ; ,, <strong>and</strong> Ç . And then<br />

compute it to a number.<br />

E.4 Turbulence Modeling<br />

36. What’s the difference between RANS <strong>and</strong> LES turbulence modeling. Apply<br />

Reynolds averaging to the Navier-Stokes equation to obtain the RANS<br />

equation [(4.10)]. Formulate the Boussinesq approximation [(4.11)].<br />

37. What assumptions are used in LES turbulence models. Formulate the governing<br />

equations for the Smagorinsky LES model [(4.2)], [(4.3)], [(4.4)].<br />

38. What assumptions are used in RANS turbulence models. Formulate governing<br />

equations for the turbulent model [(4.17)], [(4.18)], [(4.19)],<br />

[(4.20)], [(4.21)].<br />

°Ó`


Index<br />

Acoustic problems, 13<br />

Algebraic Reynolds stress model, 109<br />

Associate tensor, 130<br />

Bernoulli’s Equation, 24<br />

Bernoulli’s equation, 38, 51, 152, 154,<br />

198<br />

Blasius equation, 90<br />

Blasius stream function, 92, 96<br />

Boundary conditions, 29<br />

Boundary layer, 85, 87<br />

Boundary layer blow-off, 92<br />

Boussinesq approximation, 108<br />

Brinkman number, 56<br />

Bulk modulus, 13<br />

Cartesian Tensors, 119<br />

Christoffel’s symbol, 132<br />

Coefficient of bulk viscosity, 20<br />

Coefficient of thermal expansion, 13<br />

Coefficient of viscosity, 9, 10, 93<br />

Conjugate metric tensor, 129<br />

Conjugate tensor, 40, 130, 134<br />

Continuity equation, 16, 32<br />

Contraction of indexes, 123<br />

Contraction operation, 130<br />

Contravariant index, 117<br />

Contravariant tensor, 133<br />

Contravariant vectors, 116<br />

Convective derivative, 4<br />

Coordinate system, 116<br />

Coriolis force, 42, 43<br />

Couette flows, 54<br />

Covariant differentiation, 132<br />

Covariant index, 117<br />

Covariant vectors, 117<br />

Creeping flow, 71, 76<br />

Cross product, 124<br />

Cylindrical coordinates, 138<br />

Darcy friction factor, 62<br />

Dependent variable, 9<br />

Dependent variables, 1<br />

Direct numerical simulation, 106<br />

Dirichlet boundary, 30, 61<br />

Displacement thickness, 82, 90, 100<br />

Divergence operator, 127, 133, 138,<br />

139<br />

Divergence-free vector field, 16<br />

Dot-notation, 2<br />

Drag coefficient, 83, 98<br />

Drag force, 74<br />

Dummy index restriction, 123<br />

Dummy indexes, 122<br />

Eckert number, 52, 87, 157, 177<br />

Eddy viscosity, 106, 108<br />

Enthalpy, 7<br />

Enthapy, 8<br />

Entrance effect, 60<br />

Entropy, 6<br />

Equation of state, 5, 7, 32<br />

Euler equation, 28, 29<br />

Eulerian description, 1<br />

Falkner-Skan equation, 92<br />

Falkner-Skan wedge flow, 91<br />

First law of thermodynamics, 6<br />

Flow separation, 64<br />

203


204 INDEX<br />

<strong>Fluid</strong> element, 3<br />

<strong>Fluid</strong> particle, 1<br />

<strong>Fluid</strong> particles, 1<br />

<strong>Fluid</strong> properties, 8<br />

<strong>Fluid</strong> velocity, 2<br />

Flux, 9<br />

Fourier series, 70<br />

Fourier’s law, 10, 35<br />

Free boundary, 31<br />

Free indexes, 121<br />

free jet, 95<br />

Friction coefficient, 55, 90, 93<br />

Froude number, 49, 50, 87, 175<br />

Fully developed flow, 60, 61, 65<br />

Fundamental tensor, 129<br />

Gas constant, 6<br />

Gauss theorem, 16, 19, 35, 36<br />

Gibbs rule, 6<br />

Gradient, 9, 127<br />

Gradient approximation, 9–11<br />

Heat conduction coefficient, 10, 39,<br />

49, 93<br />

Heat conduction equation, 38<br />

Heat conductivity, 10<br />

Heat convection equation, 39<br />

Heat dominated flow, 39<br />

Hydraulic diameter, 64<br />

Hydrostatic pressure, 53<br />

Ideal fluid, 28, 152<br />

Ideal gas law, 6<br />

Incompressibility condition, 16<br />

Independent variables, 1, 6, 8<br />

Inlet boundary, 30, 34<br />

Integral momentum relation, 101<br />

Internal energy, 6, 8<br />

Invariance, 39, 116, 128<br />

Invariant, 40, 130, 132<br />

invariant expressions, 39<br />

Invariant forms, 40, 128, 131<br />

Inviscid flow, 28<br />

Irrotational flow, 23, 27, 29<br />

KE turbulence model, 111<br />

Kelvin’s theorem, 152<br />

Kinematic variables, 1<br />

Kinematic viscosity, 10, 43<br />

Kronecker delta tensor, 118<br />

Lagrangian description, 1<br />

Laminar flow, 56<br />

laminar flow, 53<br />

Laplace equation, 24, 27, 28, 71<br />

Laplace equation for pressure, 33<br />

Laplacian, 61, 66, 72, 73, 102, 128,<br />

134, 138, 139, 162, 199<br />

Large eddy simulation, 106<br />

Law of circulation, 152<br />

Law of similarity, 43, 46<br />

Lewis number, 12<br />

Lift force, 74<br />

Lowering indices, 130<br />

Lubrication theory, 76<br />

Mean wall shear stress, 64<br />

Metric tensor, 40, 128, 131<br />

Momentum equation, 131<br />

Momentum flux, 29<br />

Momentum thickness, 83, 90, 100<br />

Momentum transport, 9<br />

Moving boundaries, 30<br />

Nabla, 17, 28, 47, 48, 86, 126, 138<br />

Natural convection, 13<br />

Neuman boundary, 30, 33<br />

No-slip boundary, 30, 61<br />

Non-dimensional parameters, 46, 47,<br />

49<br />

Non-dimensional variables, 48, 49<br />

Non-inertial coordinate systems, 41<br />

Non-Newtonian, 10<br />

Nusselt number, 49, 93


INDEX 205<br />

One equation turbulence model, 111<br />

Order of a tensor, 117, 121<br />

Orthogonal coordinate system, 137<br />

Orthogonal coordinates, 40, 134<br />

Oseen approximation, 76<br />

Outlet boundary, 30, 34<br />

Parabolic equation, 66<br />

Particle trajectory, 2<br />

Permutation tensor, 123<br />

Physical component, 137<br />

Physical components of tensors, 39<br />

PI-group, 46<br />

PI-theorem, 46, 47, 157<br />

Poiseuille flow, 62<br />

Poiseuille number, 56, 62<br />

Poisson equation for pressure, 33<br />

Potential flow, 23<br />

Pr<strong>and</strong>tl number, 11, 38, 92, 177<br />

Primary dimensions, 44, 46, 48<br />

Properties of the fluid, 1<br />

Quasi-equilibrium approximation, 5<br />

Raising indices, 130<br />

Rank of a tensor derivative, 127<br />

Rank of a term, 121<br />

RANS models, 108<br />

Renaming indexes, 123<br />

Renaming of dummy indexes, 123<br />

Reynolds analogy, 92, 93<br />

Reynolds averaged Navier-Stokes equation,<br />

108<br />

Reynolds averaging, 108<br />

Reynolds decomposition, 107<br />

Reynolds number, 44, 49, 54, 55, 60,<br />

61, 75, 85, 94, 105, 175<br />

Reynolds stress model, 108, 111<br />

Reynolds stress tensor, 108<br />

Scalar product, 40, 122, 123, 130, 131<br />

Schmidt number, 12<br />

Second law of thermodynamics, 34<br />

Shear layer, 94<br />

Similarity solution, 88<br />

Skew-symmetric tensor, 124<br />

Skin-friction coefficient, 62, 83<br />

Slip boundary, 30<br />

Smagorinsky constant, 106<br />

Smagorinsky model, 106<br />

Solenoidal vector field, 16<br />

Solid body rotation, 58<br />

Spatial derivative of a tensor, 126<br />

Specific heat, 8, 11, 38, 93<br />

Speed of sound, 12<br />

Stanton number, 93<br />

Stokes flow, 71, 80<br />

Stokes paradox, 76<br />

Stokes theorem, 152<br />

Strain rate tensor, 4, 20<br />

Stream function, 17, 72<br />

Streamline, 2<br />

Stress tensor, 19<br />

Stretching factors, 135, 138<br />

Substantial derivative, 3, 40<br />

Surface area vector, 16<br />

Surface tension coefficient, 31, 50<br />

T-s relations, 7<br />

Taylor number, 105<br />

Tensor, 117<br />

Tensor derivative, 127<br />

Tensor equality, 121<br />

Tensor expression, 120<br />

Tensor identity, 125<br />

Tensor notation, 116, 120<br />

Tensor rules, 120<br />

Tensor terms, 120<br />

Thermodynamic properties, 1<br />

Thermodynamic variables, 1<br />

Time derivative of a tensor, 126<br />

Transformation matrix, 117, 119<br />

Transformation rule, 116


206 INDEX<br />

Transport properties, 1<br />

Transport property, 9<br />

Turbulence dissipation rate, 112<br />

Turbulence model, 111<br />

Turbulent closure, 110<br />

Turbulent kinetic energy, 108<br />

Two equation turbulence models, 111<br />

Two-equation turbulence models, 112<br />

Vector product, 124<br />

Velocity circulation, 50, 151<br />

Velocity potential function, 23, 29<br />

Viscous flow, 27<br />

Viscous stress tensor, 20, 33<br />

Volumetric flow rate, 62<br />

von Karman, 101<br />

Vorticity tensor, 4<br />

Vorticity vector, 5<br />

Wake, 96<br />

Wall shear stress, 62, 90<br />

Weber number, 50<br />

Wedge flows, 91

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!