Sunteți pe pagina 1din 468

lntroduction to

Fluid Mechanics

STEPHEN WHITAKER

Professor of Chemica/ Engineering


University of California at Davis

ROBERT E. KRIEGER PUBLISHING COMPANY


MALABAR, FLORIDA
Original Edition 1968
Reprint Edition 1981 w /corrections

Printed and Published by


ROBERT E. KRIEGER PUBLISHING COMPANY, INC.
KRIEGER ORIVE
MALABAR, FLORIDA 32950 Preface
Copyright 1968 by
PRENTICE HALL, INC.
Transferred to Stephen Whitaker 1976
Reprinted by arrangement with author

All rights reserved. No part of this boolc may be reproduced


in any form or by ay electronic or mechanical means in-
cluding information slorage and retrieval systems without
permission in writing from the publisher. This book is intended for use in an
The student is expected to have com:]
Printed in the United S tates of America
and to be familiar with ordinary diffe
multiple integrals, Taylor series, and
The book is based primarily on
in which the author participated wh
each department (with the excepti
advanced undergraduate courses in
to provide a rigorous foundation
eight chapters are the result of ex
can be covered satisfactorily in app
used in either a three-unit semester e
Chapters 9, 10, and 11 were adde
who may wish to use the book in
Library of Congress Cataloeinlln Publieation Data stances it may be necessary to delete s
certain sections must be covered if s
Whitaker, Stephen. These sections are marked with an
lntroduction to Fluid Mechanics.
Vector notation is used freely th~
Reprint. Originally published: Englewood Cliffs, N. J. : elegance or rigor but simply because
Prentice-Hall (Prentice-Hall international series in the
physical and chemical engineering sciences. in a form which attempts to connec
lncludes bibliographical references and index. A variety of people contributed
l. Fluid dynamics. l. Title. 11. Series: Prentice- of this text; they have the autho ,
Hall international series in the physical and chemical engi- Professor John C. Slattery of Nort
neering sciences.
text rests largely on an endless ser
(QA91l.WS8 1981) 532'.05 81-1620
ISBN 0-89874-337-0 AACR2 the problems of teaching fluid mee
10 9 8 7 6 S Davis, California
.J. A..\'-oQ.~\o Oc~oa.. \~f~O..
Uo..\l~ s. SuiMQ..'<' \'i. ~q

,INC.

Preface
.,_

may be reproduced
echanical means in-
systems without
This book is intended for use in an introductory course in fluid mechanics.
The student is expected to have completed two years of college mathematics
and to be familiar with ordinary differential equations, partial differentiation,
multiple integrals, Taylor series, and the basic elements of vector analysis.
The book is based primarily on a common core fluid mechanics course
in which the author participated while at Northwestern University. There,
each department (with the exception of Electrical Engineering) offered
advanced undergraduate courses in fluid mechanics, and it was necessary
to provide a rigorous foundation in the common core course. The first
eight chapters are the result of experience in teaching that course. They
can be covered satisfactorily in approximately 40 lectures, and thus can be
used in either a three-unit semester course or a four-unit quarter course.
Chapters 9, 10, and 11 were added to provide flexibility for those persons
who may wish to use the book in a terminal course. Under these circum-
Data stances it may be necessary to delete sorne material in Chapters 1-8; howevcr,
certain sections must be covered if subsequent material is to be understood.
These sections are marked with an asterisk.
Vector notation is used freely throughout the text, not because it leads to
Cliffs, N. J. :
series in the elegance or rigor but simply because fundamental concepts are best expressed
in a form which attempts to connect them with reality.
index. A variety of people contributed in innumerable ways to the completion
Series: Prentice- of this text; they have the author's thanks. Special appreciation is due
and chemical engi-
Professor John C. Slattery of Northwestern University, for the origin of the
81-1620 text rests largely on an endless series of conversations with him regarding
AACR2 the problems of teaching fluid mechanics to undergraduates.
Davis, California STEPHEN WHITAKER

V
Contents
t

l. lntroduction

* The continuum postulate, l. * Typesof ftow, 3. The solution of


ftow problems, 6. Units, lO. *Fluid properties, 12. * Vectors, 22.
Problems, 28.

2. Fluid Statics and One-Dimensional Laminar Flow 32

* The material volume, 32. *Fluid statics, 36. Barometers, 43. Ma-
nometers, 44. * Forces on submerged plane surfaces, 46. * Forces on
submerged curved surfaces, 49. * Buoyancy forces, 53. One-dimensional
laminar ftows, 56. Problems, 68.

3. Kinematics 75

*Material and spatia/ coordina/es, 76. * Time derivatives, 77. * The


divergence theorem, 84. * The transport theorem, 88. * Conserva/ion
of mass, 92. * Streamlines, path fines, and streak lines, 97. Problems,
104.

* Sections marked with an asterisk must be covered if subsequent material is to be


understood.

vii
Contents
viii Contenu

4. Stress in a Fluid 1O. Compressible Flow


107
The governing equations for e
The stress vector, 107. The stress tensor, 110. Symmetry of the
equation and the entropy equt
stress tensor, 119. The stress equations o[ motion, 121. Problems, 124.
tropic nozzle jfow, 404. Shoc

S. The Differential Equations of Motion 128


11 . Flow Around lmmersed B
The viscous stress tensor, 128. t Newton's law o[ viscosity, 134. t Qua/i-
tative description of the rateo[ strain, 147. The equations of motion, 153. Description of jfow, 421. Tll
Dimensional analysis, 158. Applications o[ the differential equations boundary /ayer on a jfat pi
of motion, 166. Problems, 179. Problems, 453.

6. Turbulent Flow 186 lndex


Time averages, 187. Time-averaged equations of continuity and
motion, 189. A qualitative description of turbulent ftow, 195. The eddy
viscosity, 200. Turbulent ftow in a tube, 203. Relative magnitude of
molecular and eddy viscosity, 201. Problems, 209.

7. Macroscopic Balances: lnertial Effects 211


The macroscopic mass balance, 212. The macroscopic momentum
balance, 217. The macroscopic mechanica/ energy balance, 221.
Bernoulli's equation, 230. Sudden expansion in a pipeline, 235. Sudden
contraction in a pipeline, 244. The nozzle and the Borda mouthpiece, 241.
Applications of the momentum balance, 254. Moving control volumes and
unsteady ftow problems, 261. Differentia/-macroscopic balances, 273.
Problems, 215.

8. Macroscopic Balances: Viscous Effects 285


Friction factors: definition, 285. Friction factors: experimental, 293.
Pipeiine systems, 306. Unsteady ftow in closed conduits, 320. Flow rafe
measurement, 331. Prob/ems, 341.

9. Open Channel Flow 348

Uniform ftow, 349. Gradual/y varied ftow, 355. The solitary wave, 370.
Flow over bumps, crests, and weirs, 377. Hydraulic jump, 382. Problems,
386.

t The general form of Newton's law of viscosity is listed in the previous section.
t This section requires supplementary discussion if it is presented independently of
the previous section.
Contents ix
Contents

107
1O. Compressible Flow 390

The governing equations for compressible jlow, 390. The thermal energy
tensor, 110. Symmetry of the
of motion, 121. Problems, 124. equation and the entropy equation, 395. The speed of sound, 398. Jsen-
tropic nozzle jlow, 404. Shock waves, 416. Prob/ems, 419.

128
11. Flow Around lmmersed Bodies 421
s law of viscosity, 134. t Qua/i-
47. The equations ofmotion, 153. Description of jlow, 421. The suddenly accelerated jlat plate, 422. The
of the differential equations boundary /ayer on a jlat plate, 430. Externa/ jlows and wakes, 440.
Problems, 453.

186 lndex 455


equations of continuity and
of turbulent flow, 195. The eddy
tube, 203. Relative magnitude of
209.

211
The macroscopic momentum
mechanical energy balance, 221.
in a pipeline, 235. Sudden
and the Borda mouthpiece, 247.
254. Moving control volumes and
balances, 273.

285
Friction factors: experimental, 293.
in c/osed conduits, 320. Flow rafe

348

jlow, 355. The solitary wave, 370.


r Hydraulic jump, 382. Problems,

viscosity is listed in the previous section.


fussion if it is presented independently of
N omenclaturet
1

Roman Letters e constant of integration


a acceleration vector (82) (41)
A cross-sectional area, portion Cn drag coefficient (304)
of a closed surface (33) Ca discharge coefficient (333)
A,(t) area of entrances and exits Ce contraction coefficient (334)
(214) e wave speed (372), velocity of
A,(t) area of solid moving surfaces sound (399)
(214) e"' constant pressure heat ca-
A, area of solid fixed surfaces pacity per unit mass (402)
(214) e, constant volume heat ca-
A characteristic area (287) pacity per unit mass (402)
d area of a closed surface fixed D tube diameter (5)
in space (33) d rate of strain tensor in Gibbs
da(t) area of an arbitrary closed notation (133)
surface moving in space d; 1 rate of strain tensor in index
(33) notation (133)
dm(t) area of a closed material e interna! energy per unit mass
surface (33) (392)
b width (47) e(il unit base vectors for rec-
b arbitrary constant vector tangular, Cartesian coor-
(86) dinate system (26)

t Page number in parentheses indicates where the symbol is first defined.


xi
xii Nomenclature
Nomenclature

Ev rate of viscous dissipation NFr Froude number (165) T;; total stress tensor in in
(227) Nca cavitation number (22) notation (115)
Esp specific energy (360) p absolute pressure (38) :r torque vector (48)
f friction factor (287) Pv'J) vapor pressure (22) u0 characteristic velocity (1 5
Fn drag force (286) pg gauge pressure (45) u 00 velocity far removed fron1
F force vector (10) p0 stagnation pressure (407) immersed body (430)
g gravity vector (34) p0 ambient or atmospheric v fluid velocity vector (77)
g magnitude of gravity vector pressure (43) U dimensionless fluid velo
(41) P dimensionless pressure (160) vector (160)
gc gravitational constant (11) 9 dimensionless pressure which v.,, v11 , scalar components of v
h fluid depth (43), enthalpy includes the body force vz rectangular Cartesian
per unit mass (394) term (160) ordinales (24)
h1 friction head Ioss (308) q volumetric flow rate per unit v magnitude of fluid velo
hm minor head loss (308) width (67) vector (222)
Hw change in head caused by a q heat flux vector (392) v+ dimensionless fluid vel
pump or turbine (308) Q volumetric flow rate (61) (206)
i, j, k unit base vectors for rec- Q rate of heat transfer (392) v, relative fluid velocity ve
tangular, Cartesian coor- r, O, z cylindrical coordinates (94) (156)
dinate system (24) r, O,,, spherical coordinates (94) V volume (54)
1 unit tensor (129) r spatial position vector (76) ..Y control volume fixed in Sf1
k thermal conductivity (404) r position vector locating the (33)
K head loss coefficient (311) center of stress (49) ..Ya<t) arbitrary volume moving
KE characteristic kinetic energy R material position vector space (33)
per unit volume (287) (76)
..Ym(t) material volume (33)
1 Iength (47), Prandtl mixing R dimensionless radius (162), w arbitrary velocity vector
Iength (201) gas constant (402) W rate of work (227)
L Jength (42) R,. hydraulic radius (159) x, y, z rectangular, Cartesian e
L, entrance length (171) Yl universal gas constant (402), dinates (24)
m mass (402), Ostwald-de Wael dimensionless ratio of like
X, Y,Z dimensionless rectangu
model parameter (20) quantities (162) Cartesian coordinates (1
m mass flow rate (265) s entropy per unit mass (397), y+ dimensionless distance (2
M mass (10), Mach number are length (98)
(403) S wetted perimeter (351) Greek Letters
MW molecular weight (13) 8 scalar function (93) rx;; direction cosine (118)
n number of moles (9), t time (10) fJ coefficient of expansion (1
Ostwald-de Wael model t(n) stress vector (35) y specific gravity (55), rati
parameter (20), Manning t(':, 1 net stress vector (257) specific heats (402)
roughness factor (352) T0 Bingham model yield stress boundary !ayer thickness (
n outwardly directed unit nor- (19) ;; Kronecker delta (130)
mal (35) T absolute temperature (9) ef D relative roughness (293)
NRe Reynolds number (5) T0 stagnation temperature (407) 11 length (47)
NRe,z length Reynolds number T total stress tensor in Gibbs o angle (47)
(431) notation (112) 0 dimensionless time (160)
Nomenclature
Nomenclatu re xiii
Froude number (165)
T;; total stress tensor m index 1( compressibility (13), bulk
cavitation number (22)
notation (115) coefficient of viscosity (133)
p absolute pressure (38)
vapor pressure (22)
:r torque vector (48) wave Iength (374)
u0 characteristic velocity (159) unit tangent vector (99)
gauge pressure (45)
u 00 velocity far removed from an shear coefficient of viscosity
stagnation pressure (407)
immersed body (430) (14)
ambient or atmospheric
pressure (43) v fluid velocity vector (77) llapp apparent viscosity (19)
p U dimensionless fluid velocity Po Bingham model viscosity
dimensionless pressure (160)
vector (160) (19)
dimensionless pressure which
includes the body force
v.,, vv, scalar components of v in eddy viscosity (200)
term (160) v, rectangular Cartesian co- V kinematic viscosity (16)
ordinates (24) 7T 3.1416 .. .
q volumetric flow rate per unit
width (67) v magnitude of fluid velocity p density (2)
q heat flux vector (392) vector (222) a surface tension (21)
Q volumetric flow rate (61) v+ dimensionless fluid velocity viscous stress tensor in Gibbs
(206) notation (130)
Q rate of heat transfer (392)
r, IJ, z cylindrical coordinates (94) Vr relative fluid velocity vector viscous stress tensor in index
(156) notation (130)
r, IJ, </> spherical coordinates (94)
V volume (54) turbulent stress tensor (194}
r spatial position vector (76)
r position vector locating the "Y control volume fixed in space wall shear stress (205)
(33) gravitational potential
center of stress (49)
"Ya(t) arbitrary volume moving in function (40)
R material position vector
(76) space (33) viscous dissipation function
"Y m(t) material volume (33) (223)
R dimensionless radius (162),
gas constant (402) w arbitrary velocity vector (79) 'P stream function (102)
R 11 hydraulic radius (159) W rate of work (227) w angular velocity (178)
tJt universal gas constant (402), x,y, z rectangular, Cartesian coor- w vorticity vector (152)
dimensionless ratio of like
dinates (24) n vorticity tensor in Gibbs
quantities (162) X, Y,Z dimensionless rectangular, notation (152)
s entropy per unit mass (397), Cartesian coordinates (164) !1;; vorticity tensor in index nota-
are Jength (98) y+ dimensionless distance (206) tion (138)
S wetted perimeter (351) Greek Letters Mathematical Symbols
S scalar function (93) a.;; direction cosine (118) V "del" vector operator (40)
t time (10) f1 coefficient of expansion (13) v2 the Laplacian (154)
t(n) stress vector (35) y specific gravity (55), ratio of D
t~> net stress vector (257) material derivative (78)
specific heats (402) Dt
T0 Bingham model yield stress <'l boundary !ayer thickness (426) d
(19) total derivative (77)
<'lii Kronecker delta (130) dt
T absolute temperature (9) ef D relative roughness (293) o
r stagnation temperature (407) 17 length (47) ar
partial derivative (79)
T total stress tensor in Gbbs o angle (47) () area or volume average (109)
notation (112) 0 dimensionless time (160) time average (187)
Introduction to
Fluid Mechanics
lntroduction 1

This chapter is devoted to a brief discussion of the fundamental postulates


governing the motion of fluids, the types of flow that are to be investigated,
the physical properties of fluids, and vector notation. The first five sections
are qualitative and may be read quickly; however, Sec. 1.6 must be studied
carefully for we will draw upon that material throughout the remaining
chapters.

*1.1 The Continuum Postulate

The object of this text is to formulate the equations governing the motion
of a continuum and apply them to the problem of fluid motion. In treating
a fluid as a continuum we postulate that functions such as velocity, pressure,
density, etc. are continuous point functions. In actual fact this is not true,
for the materials we wish to study are made up of molecules. We may speak
of the velocity of a molecule with sorne assurance that this quantity is well
defined; however, the velocity ata fixed Jocation in space is rather meaningless
from the molecular point of view. We need not be concerned with this
dilemma, because the cases we wish to study represent a class of problems for
which the distance between molecules is so small that they represent a con-
tinuous system.
Sec. 1.2 Types of Flow
2 lntroduction Chap. 1

the earth). 1 In the second case, the m


The density of a fluid may be defined as
very small (on the order of I0-6 cm),
p = lim
4V-+o
(tlM)
flV
(1.1-1)
may also be extremely small and coro~
of these examples, the velocity is not a
these to be cases of slip jlow. We use th
where tlM is the mass contained in a small volume tl V. As defined by Eq. intimate contacting of the fluid mole<
1.1-1, the density p might be represented by the curve shown in Fig. 1.1-1. velocity of the fluid at the solid-fluid i

*1.2 Types of Flow

We shall examine severa! types o


which divide them into various classe
differences exist, however, and it will

Compressible and incompressible fl

Very .often a fluid is considered i


"negligible" changes for "appreciable'
The words negligible and appreciable
only in terms of our experience. Thm
than 5 per cent in 100C and less th1
inclined to consider water as incompr(
be an incompressible fluid simply bec~
occur with water are satisfactorily tre~
~v
equations of motion. However, if we
~V
note that circulation patterns are set 11
the bottom of the pan is less dense th:
Fig. 1.1-1. Density as a function of volume. buoyancy effects give rise to convectiv'
density of the fluid which causes the f
"compressible," a term which genen
The volume tl V* is the same order of magnitude as the cube of the mean free fluid velocity approaches or exceeds s'
patht for gases, and is comparable to the volume of a molecule for liquids. This situation is more likely to occu
For the continuum approach tQ hold, we must be dealing with systems that about 1100 ft/sec at normal temperat1
have dimensions much larger than either the mean free path or the molecular caused by high-speed jet aircraft is an e
diameter. Since both these quantities are generally quite small, we can expect density (and pressure) that occur ata
satisfactory results for a great many practica! situations. Two examples Sonic velocity in water is about 4
where the continuum approach must be used with caution are the following: compressible flows are less likely to occ1
the motion of a spacecraft through the upper atmosphere; the motion of a gas of "water hammer" that occurs wher
through the pores of a catalyst pellet such as those currently used in petroleum
refining processes. In the first case, the pressure is very low; thus, the mean l. E. J. Opik, Physics of Meteor Flight in t
lishers, Inc., 1958), p. 13.
free path of the gaseous molecules is large (on the order of 1 ft at 70 mi from 2. P. Emmett, ed., Catalysis (New York
p. 126.
t The mean free path is the average distance traveled by molecules between collisions.
lntroduction Chap. 1 Sec. 1.2 Types of Flow 3

as the earth). 1 In the second case, the mean free path of the molecules may be
very small (on the order of I0-6 cm), yet the pore size in the catalyst pellet
may also be extreme!y small and comparable to the mean free path. 2 In both
(1.1-1)
of these examples, the velocity is not a continuous function, and we consider
these to be cases of slip jiow. We use the term "slip flow" because there is not
a small volume .V. As defined by Eq. intimate contacting of the fluid molecules with the solid surfaces, and the
by the curve shown in Fig. 1.1-1. velocity of the fluid at the solid-fluid interface need not be zero.

*1.2 Types of Flow

We shall examine severa! types of flow in this text, and the boundaries
which divide them into various classes are not always clear. Sorne distinct
differences exist, however, and it will be helpful to discuss them.

Compressible and incompressible flow

Very often a fluid is considered incompressible if its density undergoes


"negligible" changes for "appreciable" changes in temperature and pressure.
The words negligible and appreciable are rather vague, and they have meaning
only in terms of our experience. Thus, the density of water changes by less
than 5 per cent in 100C and less than 1 per cent in 100 atm, and we are
inclined to consider water as incompressible. In actual fact water is taken to
be an incompressible fluid simply beca use the types of flows which general/y
occur with water are satisfactorily treated by the incompressible form of the
equations of motion. However, if we heat a pan of water on the stove, we
note that circulation patterns are set up. They occur because warm water at
the bottom of the pan is less dense than the cooler water at the surface, and
a function of volume. buoyancy effects give rise to convective flows. Although it is the nonuniform
density of the fluid which causes the flow, such a flow is usually not termed
"compressible," a term which general usage reserves for flows where the
magnitude as the cube of the mean free
fluid velocity approaches or exceeds sonic velocity (i.e., the speed of sound).
the volume of a molecule for liquids.
This situation is more likely to occur in gases where the sonic velocity is
we must be dealing with systems that
about 1100 ft/sec at normal temperatures and pressures. The "sonic boom"
the mean free path or the molecular
caused by high-speed jet aircraft is an obvious example of the rapid changes in
are generally quite small, we can expect
density (and pressure) that occur ata shock wave.
y practica! situations. Two examples
Sonic velocity in water is about 4700 ft/sec; thus, we might expect that
be used with caution are the following:
compressible flows are less likely to occur. However, the common phenomenon
upper atmosphere; the motion of a gas
of "water hammer" that occurs when a valve is suddenly closed in a water
as those currently used in petroleum
pressure is very low; thus, the mean l. E. J. Opik, Physics of Meteor Flight in the Atmosphere (New York: Interscience Pub-
(on the order of 1 ft at 70 mi from lishers, lnc., 1958), p. 13.
2. P. Emmett, ed., Catalysis (New York: Reinhold Publishing Corp., 1955), Vol. 2,
traveled by molecules between collisions. p.126.
Sec. 1.2 Types of Flow
lntroduction Chap. 1

line is a case of compressible fl.ow in liquids. The point here is that we must because of molecular diffusion.
consider whether a given.ftow may be treated as incompressible, not whether dispersed throughout the tube
the fluid is incompressible. Compressible fl.ows are treated in Chap. 10. entrance of the tube underwe
dispersion of the dye streak is si
the thin stream of smoke given
Laminar and turbulent flow

The distinction between laminar and turbulent fl.ows is somewhat easier


to make than the distinction between compressible and incompressible fl.ows.
Laminar fl.ow is characterized by
smooth motion of one lamina offl.uid
past another, while turbulent fl.ow is
characterized by an irregular and
nearly random motion superimposed
on the main motion of the fluid. The
two types of flow can be observed in
the trail of smoke leaving the burning
cigarette shown in Fig. 1.2-1. The
smoke rises from the cigarette in a
smooth, laminar stream for perhaps
1 or 2 in.; however, at that point it
generally becomes unstable and a
transition to turbulent flow takes
place. This is characterized by whirls
and a more random motion of the
smoke stream as it rises into the air.
The transition from laminar to
turbulent flow in tubes was first
Fig. 1.2-2. Reynolds experir
investigated by Osborne Reynolds, 3 turbulence.
and a sketch of the apparatus used
by Reynolds is illustrated in Fig.
1.2-2. The system consisted essen- transition conditions could be e<
tially of a bell-mouthed glass tube now known as the Reynolds nu
into which a dye streak was injected
with the water that entered the tube N
from a reservoir. Reynolds observed
two distinct types of flow. In the first, where p = density
the dye streak maintained its identity (v,) = average velocity in th
Flg. 1.2-1. Laminar and turbulent flow D = tube diameter
from a burning cigarette.
and remained in the center of
the tube, although it spread slowly p, = viscosity
3. O. Reynolds, "Ail Experimental Investigation of the Circumstances which Determine Reynolds found that the transiti
whether the Motion of Water Shall be Direct or Sinuous and the Law of Resistance in 2100, regardless ofthe specific val
Parallel Channels." Phi/. Trans. Roy. Soc. (London) Ser. A, 1883, 174: 935.
Sec. 1.2 Types of Flow 5
lntroduction Chap. 1

The point here is that we must because of molecular diffusion. In the second, the dye streak was soon
treated as incompressible, not whether dispersed throughout the tube when the laminar flow that existed at the
ble flows are treated in Chap. 10. entrance of the tube underwent the transition to turbulent flow. The
dispersion of the dye streak is similar in sorne respects to the dispersion of
the thin stream of smoke given off by the cigarette. Reynolds found that the

turbulent flows is somewhat easier


and incompressible flows.
Laminar flow is characterized by
smooth motion of one lamina offluid
past another, while turbulent flow is
characterized by an irregular and
nearly random motion superimposed
Laminar flow
on the main motion of the fluid. The
two types of flow can be observed in
the trail of smoke leaving the burning
cigarette shown in Fig. 1.2-1. The

~--
smoke rises from the cigarette in a
smooth, laminar stream for perhaps
1 or 2 in.; however, at that point it
generally becomes unstable and a
transition to turbulent flow takes
place. This is characterized by whirls
and a more random motion of the
smoke stream as it rises into the air.
Turbulent flow
The transition from laminar to
turbulent flow in tubes was first
Fig. 1.2-2. Reynolds experimental investigation of the transition to
investigated by Osborne Reynolds,a turbulence.
and a sketch of the apparatus used
by Reynolds is illustrated in Fig.
1.2-2. The system consisted essen- transition conditions could be correlated by a dimensionless group which is
tially of a bell-mouthed glass tube now known as the Reynolds number, defined as follows.
into which a dye streak was injected
with the water that entered the tube _ p(v,) D
N Re- (1.2-1)
from a reservoir. Reynolds observed fl
two distinct types of flow. In the first, where p = density
the dye streak maintained its identity (v,) = average velocity in the z-direction
and remained in the center of D = tube diameter
the tube, although it spread slowly p, = viscosity

of the Circumstances which Determine Reynolds found that the transition took place for values of NRe of about
or Sinuous and the Law of Resistance in 2100, regardless of the specific values of p, (v,), D, and p,. In Chap. 5, we will
Ser. A, 1883, 174: 935.
Sec. 1.3 The Solution of Flow Pro
6 lntroduction Chap. 1

be able to prove that the Reynolds number is, indeed, the governing parameter Iaws or fundamental postulates
has pointed out that the funda
for the transition to turbulent flow.
The notation in Eq. 1.2-1 deserves sorne comment, for it will be used more appropriately attributed
mechanics as follows.
consistently throughout the text. Dimensionless numbers or groups will
always be denoted by the letter N with a subscript appropriate to the name l. The time rate of change <
ofthe number; area and volume averages will be denoted by angular brackets, acting on the body.
( ); and the scalar components of a vector will be denoted by either an 2. The time rate of change
alphabetical or a numerical subscript. Thus Vz represents the scalar com- mentum) of a body equal
ponent of v in the z-direction, and not the derivative of v with respect to z. the torque and the mome
The latter interpretation is commonly encountered in mathematics texts, but point.
rarely found in books on mechanics.
These laws apply not only to e
Steady and unsteady flow continuous body. Euler's first la\
second law which states :t
These two designations are fairly obvious, and we only need to clarify
their meaning in the case of turbulent flow. If a laminar flow is steady, the "The change in motion is prc
three components of the velocity-v.,, v11 , and Vz-and the pressure p are and it takes place along the ri
independent of time t. Turbulent flows are naturally unsteady; however, we
shall refer to a turbulent flow as- steady if the time-averaged components of Euler's second law is very de
velocity and pressure-v.,, v11 , vz, and ji-are independent of time. A careful continuum mechanics; however
treatment ofthe time-averaged equations of motion for turbulent flow appears stood. Confusion arises because
in Chap. 6. law if all forces are continuous fu
torques are the moments of th
present, however, the second la~
One-dimensional flow While Euler's first law is mos
second Iaw of motion, its devel
By one-dimensional flow we mean that the velocity v is a function of only efforts of a number of scientists
one spatial coordinate. One-dimensional turbulent flow, of course, implies ending with Euler. To avoid conf
that the time-averaged velocity v is a function of only one spatial coordinate. linear momentum principie and th
The flow in the Reynolds' apparatus is one-dimensional sorne distance down- In addition to the two laws o
stream from the entrance (i.e., vis only a function of r), but at the entrance, v principies of conservation of mas
is a function of r and z and the flow is two-dimensional. Often we approxi- lems of fluid motion. The latter
mate two- and three-dimensional flows by one-dimensional models, because compressible; we shall not enco
the velocity field is easily determined for a one-dimensional flow. reasons which will become appar
these fundamental postulates, we
the principies of conservation o
*1.3 The Solution of Flow Problems momentum, or the balance of mom
tum. According to the linear mo
In attempting to formula te the equations of fluid motion, we need a clear served for it may increase or decre
understanding of the fundamental postulates governing this motion. The 4. C. Truesdell, "A Program towa
student has aiready made use of Newton's second law to solve problems in Age of Reason," Archives for the Hist
statics and dynamics, and it would seem natural to include it as one of the t See Ref. 4, page 6.
Chap. 1
Sec. 1.3 The Solution of Flow Problems 7
lntroduction

laws or fundamental postula tes which govern the motion of fluids. Truesdell 4
is, indeed, the governing parameter
has pointed out that the fundamental postulates of continuum mechanics are
more appropriately attributed to Euler. We may state Euler's two laws of
sorne comment, for it will be used
mechanics as follows.
numbers or groups will
a subscript appropriate to the name l. The time rate of change of the momentum of a body equals the force
will be denoted by angular brackets, acting on the body.
a vector will be denoted by either an 2. The time rate of change of the angular momentum (moment of mo-
Thus v. represents the scalar com- mentum) of a body equals the torque acting on the body, where both
the derivative of v with respect to z. the torque and the moment are taken with respect to the same fixed
encountered in mathematics texts, but point.

These laws apply not only to discrete particles, but also to any arbitrary
continuous body. Euler's first law is, of course, the generalization ofNewton's
second law which states:t
obvious, and we only need to clarify
flow. If a laminar flow is steady, the "The change in motion is proportional to the motive force impressed,
and it takes place along the right line in which the force is impressed."
vy, and v.-and the pressure p are
are naturally unsteady; however, we
if the time-averaged components of Euler's second law is very definitely a separate fundamental postulate of
p-are independent of time. A careful continuum mechanics; however, this concept is not always clearly under-
of motion for turbulent flow appears stood. Confusion arises because the second law may be derived-from the first
law if all forces are continuous functions of mass (i.e., "body forces") and all
torques are the moments of these body forces. When shear stresses are
present, however, the second law is independent of the first.
While Euler's first law is most often referred to and known as Newton's
second Iaw of motion, its development can actually be attributed to the
that the velocity v is a function of only efforts of a number of scientists and engineers starting with Newton and
turbulent flow, of course, implies ending with Euler. To avoid confusion, we shall refer to these two laws as the
uu-ulJu of only one spatial coordinate. linear momentum principie and the angular momentum principie.
one-dimensional sorne distance down- In addition to the two Iaws of mechanics, we must also make use of the
a function of r), but at the entrance, v principies of conservation of mass and conservation of energy to solve prob-
is two-dimensional. Often we approxi- lems of fluid motion. The latter principie is required only when the flow is
by one-dimensional models, because compressible; we shall not encounter this condition until Chap. 10. For
for a one-dimensional flow. reasons which will become apparent when we explore the various forms of
these fundamental postula tes, we often refer to the two Iaws of mechanics as
the principies of conservation of momentum and conservation of angular
momentum, or the balance of momentum and the balance of angular momen-
tum. According to the linear momentum principie, momentum is not con-
of fluid motion, we need a clear served for it may increase or decrease depending u pon the force that acts upon
governing this motion. The 4. C. Truesdell, "A Program toward Rediscovering the Rational Mechanics of the
second law to solve problems in Age of Reason," Archives for the History of Exact Science, 1961, 1:31.
natural to include it as one of the t See Ref. 4, page 6.
8 lntroduction Chap. 1
Sec. 1.3 The Solutlon of Flow P
the body; however, to apply a force to a body is to supply momentum to the
body, and with this in mind we might state these laws as follows: to understand the first two ite
mechanics. Constitutive equa
the time rate of } {the rate at which momentum) applied forces. Thus Hooke's
change of momentum = is supplied to the body (by (1.3-1)
{of the body application of a force) stress, is a constitutive equatioJ
of viscosity (rate of strain pr
the time rate of } {the rate at which angular } conduction (heat flux propor
change of angular = momentum is supplied to the (1.3-2) 1aw (electrical current proporti~
{momentum of the body body (by application of a torque) equations describing idealized
relationships are called "laws"
In both Eqs. 1.3-1 and 1.3-2 the word "body" refers to a fixed quantity of linear momentum principie is a
material; therefore, a body always contains the same mass points and is al/ materials, while constitutive
sometimes referred to as a "system." If we compare these equations with the law are merely "rules" which
principie of conservation of energy, which states Constitutive equations are gene
observation, although kinetic ti
the time rate of change} = (the rate at which energy} (1.3-3)
(of energy of the body is supplied to the body viscosity for dilute gases.
An equation of state deter
or the principie of conservation of mass, pressure, the density, and the te

the time rate of change} = (the rate at which mass} (1.3-4)


(of mass of the body is supplied to the body
is a well-known example of an e
we see that the words "mass," "energy," "momentum," and "angular tions of state are available for
momentum" are interchangeable. For this reason we often speak of the and ~ensity for nonideal gases 1
principies of conservation of mass, energy, momentum, and angular momen- mined by experimental observ
tum, and we make use of mass balances, energy balances, and momentum may be derived by kinetic theo
balances to sol ve engineering problems. Although the fundamental postula tes The last category of knowleq
appear to be very similar when stated in words, we must remember that Eqs. is intuition. In. general, simple
1.3-1 and 1.3-2 are actually vector equations, whereas Eqs. 1.3-3 and 1.3-4 intuition, while the more comf
represent scalar equations. require a great deal to obtain e
The solution of problems only starts with the fundamental postulates, written the following regarding
and the four main tools which an engineer uses to reach a final solution are " ... the boundary value pr
as follows: exceedingly difficult, and p1
rigorous mathematics had n
l. the fundamental postulates; intuitive hypotheses. Of tht
2. mathematical analysis; suggestive:
3. constitutive equations, equations of state, and other experimental l. Intuition suffices for det
information; consideration.
4. intuition. 11. Small causes produce sm
infinitesimal effects.
The fundamental postulates are the easiest to deal with beca use they are com- 111. Symmetric causes produ
paratively well understood and limited in number. The mathematical analysis IV. The flow topology [i.e
required to put these fundamental postulates into useable form is not exces- field] can be guessed by
sively difficult and has wide applications. For this reason, the student's effort
5. G. Birkhoff, Hydrodynamics, A
Princeton University Press, 1960), p.'
lntroduction Chap. 1
Sec. 1.3 The Solutlon of Flow Problems 9
a body is to supply momentum to the
state these laws as follows: to understand the first two items has applications far beyond the area of fluid
mechanics. Constitutive equations describe the response of materials to
rate at which momentum} applied forces. Thus Hooke's law, which relates the strain to the applied
supplied to the body (by (1.3-1)
lilPP'm;uun of a force)
stress, is a constitutive equation for linearly elastic materials. Newton's law
of viscosity (rate of strain proportional to stress), Fourier's law of heat
rate at which angular } conduction (heat flux proportional to temperature gradient), and Ohm's
mOl'llentum is suppJied tO the (1.3-2) law (electrical current proportional to potential gradient) are all constitutive
(by application of a torque) equations describing idealized linear behavior. It is unfortunate that these
relationships are called "laws" for they are not Iaws in the sense that the
"body" refers to a fixed quantity of
linear momentum principie is a law. The linear momentum principie applies to
contains the same mass points and is
al/ materials, while constitutive equations such as Hooke's law and Newton's
If we compare these equations with the
law are merely "rules" which describe the behavior of a particular material.
states
Constitutive equations are generally formulated on the basis of experimental
observation, although kinetic theory can be used to derive Newton's law of
= (the rate at which energy) (1.3-3)
is supplied to the body viscosity for dilute gases.
An equation of state determines the relationship of the thermodynamic
pressure, the density, and the temperature. The ideal gas law,
pV= niT (1.3-5)
(1.3-4)
is a well-known example of an equation of state; however, many other equa-
," "momentum," and "angular tions of state are available for describing the relationship between pressure
this reason we often speak of the and density for nonideal gases and for liquids. Equations of state are deter-
momentum, and angular momen- mind by experimental observation, although once again the ideal gas law
energy balances, and momentum may be derived by kinetic theory for gases at low pressures.
Although the fundamental postulates The Iast category of knowledge that we require in order to sol ve problems
in words, we must remember that Eqs. is intuition. In general, simple problems can be solved with practically no
uations, whereas Eqs. 1.3-3 and 1.3-4 intuition, while the more complex problems encountered later in this text
require a great deal to obtain even an approximate solution. Birkhoff has
with the fundamental postulates, written the following regarding intuition and problem solving:
uses to reach a final solution are " ... the boundary value problems of rational hydrodynamics are
exceedingly difficult, and progress would have been much slower if
rigorous mathematics had not been supplemented by various plausible
intuitive hypotheses. Of these, the following have been especially
suggestive:
of state, and other experimental l. Intuition suffices for determining which physical variables require
consideration.
11. Small causes produce small effects and infinitesimal causes produce
infinitesimal effects.
to deal with because they are com- 111. Symmetric causes produce effects with the same symmetry.
in number. The mathematical analysis IV. The flow topology [i.e., the salient characteristics of the flow
,... .... u .......... into useable forro is not exces- field] can be guessed by intuition."
. For this reason, the student's effort
5. G. Birkhoff, Hydrodynamics, A Study in Logic, Fact, and Similitude (Princeton, N.J.:
Princeton University Press, 1960), p. 4.
10 lntroduction Chap. 1 Sec. 1.-4 Units

He goes on to say, "The preceding plausible assumptions are usually made acceleration gives rise to one unit
tacitly as a matter of course." Although Birkhoff devotes the first two chap- of g-cmfsec2 , but it is far more con
ters of his book to an examination of the fallibility of these assumptions, he Note that "gram" is a measure o
does not imply that we should refrain from using intuition in solving fluid a wise choice ofwords. In the Engll
mechanics problems. Rather he implies that we should understand clearly we make use of the word "pound
when intuition is being used. an unfortunate choice of words. 1
Throughout this text it will be pointed out to the student how easy it is to consist of pounds-mass (lbm),
to make use of these intuitive hypotheses without ever realizing that it is analogous to the cgs system we
being done. In attacking a problem it is most important to realize clearly in
what area one is working, and the student should constan ti y ask himself the ~o
following questions. Am I working with a fundamental postulate at this dt
point? Is this step simply mathematical -analysis? Do I need sorne experi-
(1 lbm-ft/!
mental information here, or is this step I am now making purely intuitive?
If it is intuitive, then go on with it, but do not be surprised if the "wrong" Thus, the force required to give a
answer is obtained. Intuition is only acquired by experience, and that 1 poundal; however, we generally
experience requires solving problems and getting the wrong answer in addi- We choose to work instead with
tion to solving problems and getting the right answer. tional field on 1 pound-mass. Th

unit of force =
1A Units

The simplest manner of handling units is to define mass, Iength, and


time arbitrarily, and then use the linear momentum principie to define a unit We call this unit of force, 1 poun~
of force. In the cgs system, 1 lbr =
M= mass (g)
If we wish to use as our units, po
v = velocity (cmfsec)
Eq. 1.4-3 must be rewritten
t = time (sec)
1d
We may write Eq. 1.3-1 as
d
- (Mv) =F (1.4-1)
dt where gc = 32.
where Mv is the momentum of the body, and F is the force acting on the Note that it is simply the Iack of a
body.t We may use Eq. 1.4-1 to define a "dyne" as the unit offorce required acceleration, and force which Ieads
to give 1 g an acceleration of 1 cmfsec2 the remainder of this text we will
d work in absolute units (dyne, gra
-(Mv) = F mass, foot, second). In working
dt (1.4-2)
2
gc is not incorporated into the eq
(1 g-cmfsec = 1 dyne) equations take a somewhat simple
Note that there is a simple "one-to-one" correspondence between the various ally required in pounds-force rath
terms in the equation; thus, one unit of mass multiplied by one unit of sion factors to accomplish this e
Eq. 1.4-5 that
t Note that throughout the text boldface type is one of the methods used to indicate
vectors. 1 = 32.
lntroduction Chap. 1 Sec. 1.4 Units 11

plausible assumptions are usually made acceleration gives rise to one unit of force. We could speak offorce in terms
Birkhoff devotes the first two chap- of g-cm/sec 2, but it is far more convenient to refer to force in terms of dynes.
of the fallibility of these assumptions, he Note that "gram" is a measure of mass, and "dyne" is a measure of force-
from using intuition in solving fluid a wise choice ofwords. In the English system ofunits, confusion arises because
that we should understand clearly we make use of the word "pound" to indicate both mass and force-surely
an unfortunate choice of words. Let us consider the English system of units
v u1uu;u out to the student how easy it is to consist of pounds-mass (lbm), feet (ft), and seconds (sec). In a manner
without ever realizing that it is analogous to the cgs system we may define a unit of force as,
it is most important to realize clearly in
should constantly ask himself the ~(Mv) = F
with a fundamental postulate at this dt (1.4-3)
.analysis? Do I need sorne experi- 2
I am now making purely intuitive?
(1lbm-ft/sec = 1 poundal)
but do not be surprised if the "wrong" Thus, the force required to give a mass of 1 lbm an acceleration of 1 ft/sec 2 is
acquired by experience, and that 1 poundal; however, we generally do not use the poundal as a unit of force.
and getting the wrong answer in addi- We choose to work instead with that force exerted by the earth's gravita-
the right answer. tional field on 1 pound-mass. Thus,

unit of force = (llbm) (32.2 ft/sec 2)


(1.4-4)
= 32.2 lbm-ft/sec2
units is to define mass, length, and
momentum principie to define a unit We call this unit of force, 1 pound-force (lbr)

1 lbr = 32.2 lbm-ft/sec 2 (1.4-5)

If we wish to use as our units, pound-force, pound-mass, foot, and second,


Eq. 1.4-3 must be rewritten

_!_ ~ (Mv) = F
g. dt
(1.4-1) (1.4-6)
where g. = 32.2lbm-ft/lbr-sec2
body, and F is the force acting on the Note that it is simply the lack of a one-to-one correspondence between mass,
a "dyne" as the unit of force required acceleration, and force which leads to the conversion factor, g. Throughout
the remainder of this text we will not use the preceding form but will always
work in absolute units (dyne, gram, centimeter, second; or poundal, pound-
mass, foot, second). In working with this set of units, the conversion factor
(1.4-2)
g. is not incorporated into the equations written in the text; therefore, the
= 1 dyne) equations take a somewhat simpler form. However, final answers are gener-
correspondence between the various ally required in pounds-force rather than poundals, and we must use conver-
of mass multiplied by one unit of sion factors to accomplish this change of units. lt is useful to note from
Eq. 1.4-5 that
type is one of the methods used to indicate
1 = 32.2 lbm-ft/lbr-sec2
12 lntroductlon Chap. 1 Sec. 1.5 Fluid Properties

Table 1.4-1 Density


UNITS AND CoNVERSION FACTORS
Mass Length Time Table 1.5-1 lists the physi
llbm = 453.6 g 1 in. = 2.54 cm 1 min = 60sec densities (except that for mercu
1 ton (short) = 2000 lbm 1ft= 12 in. 1 hr = 60min
1 ton (long) = 2240 lbm 1 yd =3ft 1 day = 24hr
1 kg = 1000 g 1 mi= S280ft
Al'PROXIMATE PHYSICAL
1 slug = 32.2 lbm 1 angstrom ()= 10- cm
1 micron (p) = 1o-' cm Substance T, op

Force Area Volume Ethyl alcohol 68


1 dyne = 1 in.' = 6.45 cm 1 in.' = 16.4 cm Benzene 68
2.25 X lQ-1 lbt 1 ft1 = 144 in.' 1 ft 1= 1728 in.' Carbon tetrachloride 68
1 dyne = 7.23 x lO-' 1 acre = 4.35 x 10' ft 1 1 gal = 231 in.' Mercury 68
poundal 1 qt (dry) = 1101 cm SAE 30 oil 100
1 qt (Iiquid) = 0.25 gal Water 60
1 bl = 31.5 gal
Pressure Viscosity Power compressibility K, defined as

1 atm = 14.7lbr/in.S 1 poise = 1 g!cm-sec 1 hp = 3.3 x 10' lbr-ft/min


1 atm = 29.9 in. of Hg 1 poise = 6. 72 x 10- llbr-ft/min = 1.29 x to-Btu/min
1 atm = 33.8 ft of H 10 lbm/ft-sec 1 Btu (39F) = 1055 watt-sec and the coefficient of expansion
1 atm = 1.013 x 10 1 poise = 101 centipoise
dyne/cm
fl=
and, similarly,
1 = 12 in./ft are so small for liquids, we gen
1 = 1.8Foc temperature and pressure on dem
1 = 3600 sec/hr depends strongly on the temper
etc. (298K and 1 atm), the density,
can be estimated by the ideal ga
Since g. = 32.2lbm-ft/lbr-sec2 is also equal to unity (dimensionless), we may
multiply any quantity by this number without changing its value-i.e.,
multiplication by one does not change the value of the quantity.
Table 1.4-1 lists a variety of conversion factors from one set of units to
another. A more thorough set of conversion tables can be obtained in the
Handbook of Chemistry and Physics. 6

*1.5 Fluid Properties p

In the study of single-phase incompressible flow, we need know only the where MW is the molecular wei~
density p and the viscosity p. if the fluid is Newtonian. For non-Newtonian Whenever the temperature or ~
fluids, we require additional parameters to characterize the viscous behavior the pressure and temperature at t
of the fluid. gas behavior becomes significan
sophisticated correlations must be
6. C. D. Hodgman, ed., Handbook of Chemistry and Physics (Cleveland, Ohio:
Chemical Rubber Publishing Co., 1956).
and experimental information pe1
lntroductlon Chap. 1 Sec. 1.5 Fluid Properties 13

Density

Time
Table 1.5-1 lists the physical properties of sorne common liquids. The
1 min = 60sec densities (except that for mercury), range from 50 to 100 lbm/ft3 Since the
1hr=60min
1 day = 24hr Table 1.5-1
APPROXIMATE PHYSICAL PROPERTIES OF SoME COMMON LIQUIDS

Substance T, F p,lbm/ft3 K, atm- 1 {1, K- 1


Volume Ethyl alcohol 68 49.2 120 X IO- 1.12 X 10-a
Benzene 68 54.8 90 X 10- 1.20 X 10-a
1 in. = 16.4 cm
Carbon tetrachloride 68 128 98 X 10- 1.24 X 10-a
1 ft 1 = 1728 in.
1 gal = 231 in. Mercury 68 846 3.9 X IO- 0.18 X I0- 8
SAE 30 oil 100 57.5 50 X 10- 0.90 x to-a
1 qt (dry) = 1101 cm
Water 60 62.4 52 X 10- 0.21 X I0-3
1 qt (liquid) = 0.25 gal
1 bl = 31.5 gal
Power compressibility K, defined as
1 hp = 3.3 x 10' lbr-ft/min
1 Ibt-ft/min = 1.29 x 10-Btu/min
K= !(op)
P op T
(1.5-1)

1 Btu (39F) = 1055 watt-sec


and the coefficient of expansion {3, defined as
centipoise

f3 = _l(ap) (1.5-2)
Par"'
are so small for liquids, we generally do not need to consider the effect of
temperature and pressure on density. On the other hand, the density of gases
depends strongly on the temperature and pressure. At standard conditions
(298K and 1 atm), the density, compressibility, and coefficient of expansion
can be estimated by the ideal gas law as
equal to unity (dimensionless), we may
without changing its value-i.e., p(MW)
the value of the quantity. p=-- (1.5-3a)
&IT
factors from one set of units to
~nT11VI'I'~It1m tables can be obtained in the 1
K=- (1.5-3b)
p
1
{3=- (1.5-3c)
T

"ble fiow, we need know only the where M W is the molecular weight.
fluid is Newtonian. For non-Newtonian Whenever the temperature or pressure are close to the critica! values-i.e.,
to characterize the viscous behavior the pressure and temperature at the critica/ point-the deviation from ideal
gas behavior becomes significant and either experimental data or more
sophisticated correlations must be used. There is a great deal of theoretical
of Chemistry and Physics (Cleveland, Ohio:
and experimental information pertaining to the densities and viscosities of
14 lntroduction Chap. 1 Sec. 1.5 Fluid Properties

gases as a function of pressure and temperature. A summary of this informa- moving at a constant velocity u0
tion has been compiled by Reid and Sherwood 7 and an elaborate treatment of dicated. This idealized flow can 1J
the subject is given by Hirschfelder, Curtiss, and Bird. 8 viscometer, which consists of a
of coaxial cylinders arranged as ill
Viscosity trated in Fig. 1.5-2. The inner cyl
der is fixed and a torque is appli
The distinctive mechanical characteristic of a fluid is that it deforms con- to the outer cylinder causing it
tinuously under the action of a shear stress. Solids, of course, suffer a rotate at a steady rate. As h/r0 b
deformation (or "strain") under the action of a shear stress; however, this comes small, the curvature of
deformation is independent of time except for the very slow changes from annular region can be neglected a1
"creep" phenomena. Fluids, on the other hand, will continue to deform as the flow in a Couette viscometer t
long as the shear stress is applied. sembles the rectilinear flow shov
The rate at which a fluid deforms is given by the rate of change in distance in Fig. 1.5-1.
between two neighboring points moving with the fluid divided by the distance For many fluids, experiments
between the points. Thus, the rate of deformation is given by the "change in this kind Iead to results similar
length per unit length per unit time." In solid mechanics, the change in those shown in Fig. 1.5-3, where t
length per unit Iength is called the strain; it is therefore natural to refer to the force per unit area, F/A, required
rate of deformation as the rate of strain. However, the student should be maintain the motion is plotted as
forewarned that the terms rate of strain, strain rate, and shear rate (used function of the velocity u0 divided ~
especially with simple shearing flows) may all be used in discussing the rate of the plate spacing h. In a Couette v
deformatioil. of a fluid. cometer, the force F is given by t
For Newtonian fluids there is a simple linear relationship between the applied torque divided by the Iever
applied shear stress and the rate of strain. The coefficient relating the shear the outer cylinder. These results m
stress and the rate of strain is called the coefficient of viscosity or, more F
commonly, "the viscosity." The nature of viscosity can best be introduced A
by describing ~n experiment often performed to measure the coefficient of
viscosity, .. Consider the fluid motion between two long parallel plates such
as those illustrated in Fig. 1.5-1. One of the plates is fixed and the other is

Moving plote

Fixed plote

Fig. 1.5-1. Velocity distribution between two parallel plates.

7. R. C. Reid and T. K. Sherwood, The Properties of Gases and Liquids (New York:
McGraw-Hill Book Co., lnc., 1958).
8. J. O. Hirschfelder, C. S. Curtiss, and R. B. Bird, Molecular Theory of Gases and Flg. 1.5-l. Typical resul
Liquids (New York: John Wiley and Sons, Inc., 1954). experiment.
lntroduction Chap. 1 Sec. 1.5 Fluid Properties
15
temperature. A summary ofthis informa- moving at a constant velocity u0 giving rise to the simple shearing flow in-
Sherwood7 and an elaborate treatment of dicated. This idealized flow can be approximated quite closely in a Couette
, Curtiss, and Bird. 8 viscometer, which consists of a pair
of coaxial cylinders arranged as illus-
trated in Fig. 1.5-2. The inner cylin-
der is fixed and a torque is applied
~forof'~of a fluid is that it deforms con- to the outer cylinder causing it to
stress. Solids, of course, suffer a rotate at a steady rate. As h/r0 be-
action of a shear stress; however, this comes small, the curvature of the Cup rotales
except for the very slow changes from annular region can be neglected and
other hand, will continue to deform as the flow in a Couette viscometer re- L rro+h
sembles the rectilinear flow shown
B~b
Fluid to be
is given by the rate of change in distance in Fig. 1.5-1. tested in gap
1
with the fluid divided by the distance For many fluids, experiments of (stationary)
deformation is given by the "change in this kind lead to results similar to 1

" In solid mechanics, the change in those shown in Fig. 1.5-3, where the
; it is therefore natural to refer to the force per unit area, F/A, required to
strain. However, the student should be maintain the motion is plotted as a
strain, strain rate, and shear rate (used function of the velocity u0 divided by
may all be used in discussing the rate of the plate spacing h. In a Couette vis- Fig. 1.5-2. Couette viscometer.
cometer, the force F is given by the
simple linear relationship between the applied torque divided by the lever arm r0 +
h, and the area A is the area of
strain. The coefficient relating the shear the outer cylinder. These results may be expressed as
the coefficient of viscosity or, more
of viscosity can best be introduced (1.5-4)
r~n. r-~.- to measure the coefficient of
between two long parallel plates such
of the plates is fixed and the other is

o Experimental paints
between two parallel plates.

Properties of Gases and Liquids (New York:

R. B. Bird, Molecular Theory of Gases and Fig. 1.5-l. Typical results from a Couette viscometer
Inc., 1954). experiment.
16 lntroduction Chap. 1 Sec. 1.5 Fluid Propertles

where .t is the slope of the line drawn through the experimental points. We
will find it convenient to write the force per unit area acting on a surface for
whichy is constant (henceforth referred toas the y-surface) in the x-direction
1000
80 Ot--~
6 00
ttc-
""o.. e-
1\~
as 40o ., f- 1
30 o ~
F
-=T.
A vo:
(1.5-5) \o-~ <S)
20o
The velocity gradient may be expressed as a derivative \
10o
u0 = dv., 8o 1
(1.5-6)
h dy 60
and Eq. 1.5-4 takes the form
T. _ dv.,
4 o~"o.
3 Or--~o~
11.,-.t dy (1.5-7)
2O K ~o t-

which is Newton's law of viscosity for this particular simple fl.ow. ~-t
The subscripts used here deserve sorne attention. Throughout the text, T 10 f\
with two subscripts will be used to denote the total stress (force per unit area) 8
acting on a surface. The first subscript will indicate the surface upon which 6
the stress is acting; the second will indicate the direction of the stress. Thus, 4 lre,-1-
T11., is the stress acting on the y-surface in the x-direction, while T., 11 is the 3 .... 08111)e-
stress acting on the x-surface in the y-direction. The scalar components of a
2
!'---~ ~;
Mercury-'1"'1) 1
11
vector such as the velocity vector v may be denoted by alphabetic subscripts.
Thus, v.,, v11 , and v. represent the components of the velocity vector in the GosoJ"
x-, y-, and z-directions, respectively. .o
1
r-;..!" }'
The viscosity of a wide variety of commonly encountered fl.uids is pre- o.8
Go 10).~
O. 6
sented as a function oftemperature in Fig. 1.5-4. The fl.uids we may encounter
in practice have viscosities ranging over severa! orders of magnitude. Com- O. 4 Joj3
o

mon Iubricating oils are in the range of lOO centipoise. Water, kerosene, o. 3 ~
gasoline, and other similar fl.uids have viscosities on the order of 1 centipoise.
Gases such as air, saturated steam, and hydrogen have viscosities nearly
o. 2 ~))
~
r-~~~
three orders of magnitude less than that of water. As we shall see in Iater
0.10
r-..,....
studies, the kinematic viscosity, or the ratio of viscosity to the density, is a 0.08
key physical property in understanding fluid motion. In Fig. 1.5-5, the kine- 0.06
matic viscosity is given as a function of temperature. Here we see that such
diverse fl.uids as water and air have nearly the same order of magnitude for 0.04
the kinematic viscosity, and, in general, the range of kinematic viscosities is 0.03
somewhat Iess than that of the absolute viscosity. 0.02 A ir t-

0.0 1
Methone CH4
1
j'
Non-Newtonian fluids
0.008
10 20 3040
Fluids which exhibit a linear relationship between the shear stress T11.,
and the shear rate (dv.,fdy) are called Newtonian fl.uids. Many fl.uids exhibit Fl. l.
lntroduction Chap. 1 Sec. 1.5 Fluid Propertles 17

through the experimental points. W e


force per unit area acting on a surface for
toas the y-surface) in the x-direction

(1.5-5)

as a derivative
u0 = dv., (1.5-6)
h dy

dv.,
=.t- (1.5-7)
dy
for this particular simple flow.
sorne attention. Throughout the text, T 2
denote the total stress (force per unit area)
will indicate the surface upon which
1 x 10-4
indicate the direction of the stress. Thus, 8
in the x-direction, while T.,11 is the 6
y-direction. The scalar components of a
may be denoted by alphabetic subscripts. 4
3
components of the velocity vector in the

of commonly encountered fluids is pre-


in Fig. 1.5-4. The fluids we may encounter
over severa! orders of magnitude. Com-
of 100 centipoise. Water, kerosene,
viscosities on the order of l centipoise.
, and hydrogen have viscosities nearly
that of water. As we shall see in later
the ratio of viscosity to the density, is a
fluid motion. In Fig. 1.5-5, the kine-
of temperature. Here we see that such
nearly the same order of magnitude for
the range of kinematic viscosities is
viscosity.

"''"uu'"~'ll'~ between the shear stress T11 ., Temperature, "F


Newtonian fluids. Many fluids exhibit Fl1. 1.5-4. Viscosity of fluids.

/
Sec. 1.5 Fluid Properties
18 lntroduction Chap. 1

this type of behavior; how


do not. These are called n
rate relationship for sorne fl

"'Q)
-"'-
0
c
11

l
...~
E
u
i
;;
o
bl
;;
u
.,
:;
E
Q)
e: Flg. 1.5-6.
~

The curves shown in Fi


non-Newtonian fluid beha
The slope of these curves is
by f'app

The Blngham model

The Bingham model d


the material will not defor
sorne critical value, T0 T
given by
Mercury
' ~,..

o.ooooo1 ._____~~~;f;:;-L;::L;;~:Lhr--~!:::......,~::L~W..+lil:l 01
10 20 30 40 60 80 100 200 400 600 102
Temperolure, F d
Fig. 1.5-5. Kinematic viscosity of fluids. d
lntroduction Chap. 1 Sec. 1.5 Fluid Propertles 19

this type of behavior; however, there is an important class of fiuids which


do not. These are called non-Newtonian fiuids, and the shear stress-shear
rate relationship for sorne fluids of this type is indicated in Fig. 1.5-6.

"'Ql
r,...
.JL
o
c
11

~
"'uE
...>.
;;
o
u
"'
;;
u
:; (~)
"'Ql
E dy
.S Flg. 1.5-6. Behavior of non-Newtonian fluids.
:.::

The curves shown in Fig. 1.5-6 indicate just a few of the many types of
non-Newtonian fluid behavior which have been observed experimentally.
The slope of these curves is often called the apparent viscosity, and denoted
by /l-app

The Bingham model

The Bingham model describes fluids which possess a yie/d stress, i.e.,
the material will not deform unless the magnitude of the shear stress exceeds
sorne critical value, T0 Thus, the shear stress-shear rate relationship is
given by
if ITv:r:l > T0 (1.5-Sa)

dvz = 0
if 17;,:~:1 ::;; To (1.5-Sb)
dy
20 lntroductlon Chap. 1 Sec. 1.5 Fluid Properties

Although no real fluids are described exactly by the Bingham model, severa! The material presented in this
fit the model reasonably well. Ordinary paint behaves somewhat like a mation has been given for the stu
Bingham fluid. This characteristic is beneficia!, for it means that there is a laminar flow problems. Several 1
critica! film thickness below which paint will no longer flow under the action thorough discussion of non-Newt
of gravity, allowing it to be applied to vertical walls without excessive runoff.
Surface tenslon
The Ostwald-de Wael model ("power-Jaw" fluid)
For two phase flows, the surfa
Fluids for which the apparent viscosity decreases with increasing shear property; however, situations whet
rate are called pseudoplastic. When the apparent viscosity increases with not be covered in this text. Even
shear rate the fluid is called dilatant. This type of behavior may be described briefly.
The molecules at an interface
by 1d ,..-1 d
T. =m~ ~ (1.5-9) from that of the molecules in t
""' dy dy interface, for example, the water 1
by water molecules on only one
Table 1.5-3
completely surrounded. Consequ
BINOHAM MODEL PARAMETERSa
surface molecules differs from th
Fluid To, lbr/ft1 p 0 , lbr-sec/ft1
exists in a state of tension. This te
tension and has units of force p
Water o 2.09 X1Q-1 drop shown in Fig. 1.5-7 indicate
Honey 1.2 2.40 X1Q-1
Mayonnaise 0.8 1.32 XJO-I
Ketchup 0.3 1.74 x10-l
Enamels, glossy 0--0.06 2.93 x 10-~- 8.15 x 10-l
Enamels, semiglossy 0.10--0.25 2.09 X JO-I - 7.32 X JO-I
Flat or matte paints 0.04-0.21 1.25 x 1o- - 2.09 x 1o-
Water-dispersible paints 0.02~.21 0.42 X JO-I- 2.93 X 1Q-1
Varnishes o 1.88 X J0-1 - 6.06 X 1Q-1
a Data laten from F. R. Elrlcb, ed., Rlteology-Tireory lUid Pracllce (New York: Academlc
Presa. 1960), Vol. 3, Chap. 6.

When n = 1, this equation reduces to Newton's law ofviscosity with m= p.


If n < 1, it describes a pseudoplastic fluid, and if n > 1, the behavior is
dilatant. Sorne representative values of the Bingham model parameters and
the Ostwald-de Wael model parameters are given in Tables 1.5-3 and 4.
I forces P; 1Tt~
Table 1.5-4
STWALD-DE WAEL MODEL PARAMETERSa t
lntenor
Are'
Composition by Weight, pressure
Fluid per cent m, lbr-secw/ft1 n (dimensionless)
Flg. 1.5-7. Force
Carboxymethyl cellulose
in water 3.0 0.194 0.566 9. A. B. Metzner, Advances in Ch.
Lime in water 33.0 0.150 0.171 1960), Vol. l.
Napalm in kerosene 10 0.089 0.52 10. R. B. Bird, W. E. Stewart, and
Paper pulp 4 0.418 0.575 John Wiley and Sons, Inc., 1960), Cha
11. A. Fredrickson, Principies and
a Data takell from il.-B. Blrd, W. E. Stewart, aDd E. N. Llhtfool, Tr11111port PINMIMtlll (New York: Jolm Wlley
aDd Soal, IDC., 1960). Prentice-Hall, Inc., 1964).
lntroductlon Chap. 1 Sec. 1.5 Fluid Properties 21

exactly by the Bingham model, severa! The material presented in this section has been brief, but sufficient infor-
paint behaves somewhat like a mation has been given for the studeot to solve sorne simple non-Newtonian
beneficia!, for it means that there is a laminar flow problems. Severa! sources9- 11 are available which provide a
will no longer flow under the action thorough discussion of non-Newtonian fluids.
vertical walls without excessive runoff.
Surface tenslon

For two phase flows, the surface tension a may be an important physical
decreases with increasing shear property; however, situations where surface tension plays a significant role will
the apparent viscosity increases with not be covered in this text. Even so, it is worthwhile to discuss the subject
type of behavior may be described briefly.
dv
-"'
n-l -"'
dv
(1.5-9)
The molecules at an interface between two fluids exist in a state different
from that of the molecules in the interior of the fluid. At an air-water
dy dy
interface, for example, the water molecules are (on the average) surrounded
by water molecules on only one side, whereas the interior molecules are
1.5-3 completely surrounded. Consequently, the configurational energy of the
surface molecules differs from that of the interior molecules and the surface
P.o. lbr-sec/ft1 exists in a state of tension. This tension is known as the surface or interfacial
tension and has units of force per length. A force balance on the spherical
2.09 X 1Q-1
2.40 X 1Q-1 drop shown in Fig. 1.5-7 indicates that the pressure in the interior of the drop
1.32 X JO-I
1.74 X 10-1
2.93 X JQ-1 - 8.15 X J0-1
2.09 X JO-I- 7.32 X JQ-1
1.25 X JO-- 2.09 X 10-
0.42 X JO-I- 2.93 X J0- 1
1.88 X JO-- 6.06 X 1Q-1

Newton's law of viscosity with m = p.


fluid, and if n > 1, the behavior is
the Bingham model parameters and
are given in Tables 1.5-3 and 4.
I torces P; TTr2 - Po rrr 2 (j 2rrr0
to to t
Weight,
PARAMETERSr. t
Interior
Areo t
Ambient
Areo t Perimeter
Surface
1 pressure pressure tension
m, lbr-sec"/ft n (dimensionless)
Fl. 1.5-7. Force balance on a spherical drop.

0.194 0.566 9. A. B. Metzner, Advances in Chemical Engineering (New York: Academic Press,
O.J50 0.171 1960), Vol. l.
0.089 0.52 JO. R. B. Bird, W. E. Stewart, andE. N. Lightfoot, Transport Phenomena (New York:
0.418 0.575 John Wiley and Sons, lnc., 1960), Chaps. J, 3.
Llahtfoot. Tr11111p~~rt I'Mrfi>IMIIII (New York: Jolm Wlley
Jl. A. Fredrickson, Principies and Applications of Rheology (Englewood Cliffs, N.J.:
Prentice-Hall, Inc., 1964).
Sec. 1.6 Vectors
22 lntroduction Chap. 1

p, is given by
2a
P;=Po+- (1.5-10)
ro
where p 0 = ambient pressure
r0 = radius of the drop
For an arbitrary surface it can be shown12 that the pressure "jump" or
discontinuity is
(1.5-11)

where r1 and r 2 are the principal radii of curvature. There are sorne fasci-
nating and important situations where surface tension effects domnate the
fluid motion. These cases are best illustrated by the brilliant movie prepared
by Professor L. Trefethen, 13 which is both educational and entertaining.

Vapor pressure

Under sorne conditions, the fluid pressure can become less than the vapor
pressure,t p1111 , and cavitation may occur. The most common cases are in
high-speed pumps and in the wake of a bluff body, such as that shown in
Fig. 1.5-8. The cavitation number, Nca, is defined as,
N ca-
-Po- 1
Pv'
2
IJ (1.5-12)
"[PUo
and is an important parameter in establishing the probability that cavitation
will occur. Figure 1.5-8 shows a missile which has been launched under
water and has just reached the surface. The cavitation numbers range from
1.10 to 0.054, and the results indicate that we can expect severe cavitation for
this type of flow when Nca is less than 0.10. Since cavitation can drastically
alter the operating characteristics of a centrifuga! pump14 and lead to serious p.= 1.00 atm Po
corrosion, knowledge of the vapor pressure of a fluid is an important factor u. = 44.2 ft/sec u.
in pump design and operation. Nca = 1.10 N ca
Fig. 1.5-8. Cavitation behind
courtesy of U.S. Naval Ordnan
*1.6 Vectors cussion see NAVWEPS Report
Missiles, by J. G. Waugh and G.
The task of describing the complex world that surrounds us in terms of
it is wise to use symbols which con
abstract symbo/s is an immense one. In addressing ourselves to this problem,
which relate this information to t
12. R. Aris, Vectors, Tensors, and the Basic Equations of Fluid Mechanics (Englewood Inasmuch as velocity is a quantity
Cliffs, N.J.: Prentice-Hall, lnc., 1962), Chap. 10.
vector), we choose a notation whi
13. L. Trefethen, "Surface Tension in Fluid Mechanics," distributed by Educational
Services, lnc., 47 Galen St., Watertown, Mass. 02172. velocity as v. Instead of using bold
t The vapor pressure is the pressure exerted by the vapor in equilibrium with the liquid. just as well place an arrow over th<
Thus, the vapor pressure of water at lOO"F is 0.065 atm, and at 212"F it is 1.00 atm. v*, etc. Use of boldface type is th
14. R. H. Sabersky and A. J. Acosta, Fluid Flow-A First Course in Fluid Mechanics
(New York: The Macmillan Company, 1964), Chap. 10.
t The use of boldface type to indicate
lntroduction Chap. 1 Sec. 1.6 Vectors 23

2a
Po + - (1.5-10)
ro

shown12 that the pressure "jump" or

(1.5-11)

There are sorne fasci-


surface tension effects domnate the
by the brilliant movie prepared
both educational and entertaining.

can become less than the vapor


The most common cases are in
a bluff body, such as that shown in
is defined as,

(1.5-12)

~v"''"u;
the probability that cavitation
which has been launched under
The cavitation numbers range from
we can expect severe cavitation for
0.10. Since cavitation can drastically
centrifuga! pump 14 and lead to serious
Po = 1.00 atm p.= 0.51 atm p. = 0.10 atm
of a fluid is an important factor u. 44.2 ft/sec
= u. = 47.8 ft/sec u.= 55.5 ft / sec
Nca = 1.10 Nca = 0.47 Nca = 0.054
Fig. 1.5-8. Cavitation behind an underwater missile. Photographs
courtesy of U.S. Naval Ordnance, Pasadena, Calif. For a detailed dis-
cussion see NAVWEPS Report 7735, Part 1, Water-Exit Behavior of
Missiles, by J . G. Waugh and G. W. Stubstad.
world that surrounds us in terms of
addressing ourselves to this problem, it is wise to use symbols which contain as much information as possible and
which relate this information to the physical world as closely as possible.
Equations of Fluid Mechanics (Englewood
10. Inasmuch as velocity is a quantity having both magnitude and direction (a
Mechanics," distributed by Educational vector), we choose a notation which symbolizes this fact and represent the
02172. velocity as v. Instead of using boldface typet to represent a vector, we could
by the vapor in equilibrium with the liquid. just as well place an arrow over the symbol, v, or mark it with an asterisk,
0.065 atm, and at 212F it is 1.00 atm.
v*, etc. Use of boldface type is the traditional method of communicating
Flow-A First Course in Fluid Mechanics
Chap. 10. t The use of boldface type to indicate vectors is known as Gibbs notation.
24
lntroduction Chap. 1
Sec. 1.6 Vectors
information, and, like Pavlov's dogs, we respond toa symbol in boldface type
wth the thought, "lt's a vector." Asirle from the physical
Our physical notion of a vector is that of a directed line segment, such as has certain definite transfo
that iUustrated in Fig. 1.6-1. The vector v may be represented in terms ofits two-dimensional form beca
three-dimensional rotation
z sions is shown in Fig. 1.6-2,
y' y

Vz --------- V

X
Fig. 1.6-J. A vector in rectangular Cartesian coordinates. Fig. 1.6-2

three scalar components v.,, V11 , and Vz by the relationship x,y-coordinate system, and th
associated with the transform
v = iv., 11 + jv + kvz (1.6-1)
where i,j, and k are the three unit base vectors. The three scalar components l. 0.,.,,, angle between the
are the projections of v on the three coordinate axes. The scalar (or dot) 2. 011.,,, angle between the .
product between two vectors, A and B, is defined as 3. ()""''' angle between the
A B = AB cos () (1.6-2) 4. ()1/ll,, angle between the .
where A and B are the magnitudes of the two vectors, and () is the angle The subscripts of () obviously
between them. The scalar product of an arbitrary vector and a unit vector helpful in constructing a com
yields the projection of the vector on a Iine defined by the unit vector. For for vectors. By the constructio:
example,
V~= t
V i = V COS ()""' = V., (1.6-3) and by simple geometrical cons
wh(!re ()""' is the angle between the vector v and the x-axis.
ot=
lntroduction Chap. 1
Sec. 1.6 Vectors 25
respond to a symbol in boldface type
Aside from the physical significance of a directed line segment, a vector
has certain definite transformation properties. These are best examined in
that of a directed line segment, such as
two-dimensional form because the algebraic effort associated with a fully
v may be represented in terms of its
three-dimensional rotation can obscure the analysis. A vector in two dimen-
sions is shown in Fig. 1.6-2, and the scalar components are noted for both the
y' y

- - - - - - - - - eJ. - - ~
V
yy 1/ co"' _..
\
__./V 1
\ 1 1
\ 1 X

1
1

Cartesian coordinates. Fig. 1.6-2. Transformation of a vector.

x,y-coordinate system, and the rotated coordinates, x', y'. The four angles
associated with the transformation are given by the following expressions:
(1.6-1)
l. 0.,.,., angle between the x-coordinate axis and the x'-coordinate axis;
vectors. The three scalar components 2. 011.,, angle between the y-coordinate axis and the x'-coordinate axis;
coordinate axes. The scalar (or dot)
is defined as 3. ()ZII., angle between the x-coordinate axis and the y'-coordinate axis;
4. 01111 , angle between the y-coordinate axis and the y'-coordinate axis.
(1.6-2)
The subscripts of () obviously have a definite meaning, and they will be
the two vectors, and () is the angle helpful in constructing a compact formulation of the transformation law
an arbitrary vector and a unit vector for vectors. By the construction shown in Fig. 1.6-2 we may write,
a line defined by the unit vector. For
V~ = V., COS ().,.,, +V 11 COS ex (1.6-4)
(1.6-3) and by simple geometrical considerations, the angle ex may be expressed
v and the x-axis. (1.6-5)
Sec. 1.6 Vectors
26 lntroduction Chap. 1

Substitution of Eq. 1.6-5 into Eq. 1.6-4 yields Index notation is most usef
ofvectors. For example, the
v; = v., cos O.,.,. + VIl cos o""'' (1.6-6a) Eqs. 1.6-7 as
An expression for v~ may be derived in a similar manner. V~= V1 C
v; = v., cos 0., + v cos 011 (1.6-6b)
11 1111 V~= V1 CO
The extension of this result to three dimensions is straightforward but V~= V1 CO
tedious; we shall list only the results.
where it is understood that
v; = v., cosO.,.,. + V 11 cos (J""', + v. coso.,. (1.6-7a)
(') coordinate system. The :
v; = V., COS 0., + V
11 01111
11 COS 11 + v. COS 0. (1.6-7b) Eqs. 1.6-12 toa more compa
v; = v., cosO., . + V 11 cos 011 + v. coso...
(1.6-7c)
The terms, cosO.,.,,, cos 0.,,,, etc., are called the direction cosines for the
rotation, x, y, z---+- x', y', z'. We note in these equations an ordered arrange-
ment of subscripts, the first always identical to the subscript on the velocity
component in the x, y, z system, and the second always identical to the Although i was used as the r
subscript on the velocity in the x', y', z' system. This order will allow us to k,j, n, etc.--could have been
express Eqs. 1.6-7 in a much more compact forro if we make use of index
index may be represented by
notation rather than Gibbs notation.
single, or unrepeated, index is
from 1 to 3. Using the con
lndex notation 1.6-13 further by writing

Traditionally, we denote the three coordinate axes of a rectangular


Cartesian coordinate system as x, y, and z, and the three scalar components Certainly Eq. 1.6-14 is a far n
of the velocity vector as v.,, v11 , and v. It would be just as satisfactory to presented in Eqs. 1.6-12; how
denote the coordinate axes as x 1 , x 2 , and x 3 , and the scalar components as is not to produce a "slick" m
V, v2 , artd v3 . Similarly, the base vectors can be expressed as stress (Chap. 4), in Newton's
(1.6-Sa) energy equation (Chap. 7), w
e< 1 l = i
because it saves us from gettin
_e(2) = j (1.6-Sb)
In Gibbs notation, the ve
e< 3 , =k (1.6-Sc) index notation may be used t
Using this notation, we may express the vector vas remembered here is that the
relative to the x,y,z-coordinat
v = iv., + jv11 + kv. = e(l,v1 + e< 2 ,v2 + e< 3 ,v3 (1.6-9) components just as well by
Use of numerical subscripts allows us to express a vector in terms of its scalar Gibbs notation is that by tra
components and base vectors as of a directed quantity, while
3
three numbers. The former i
v = !e(i)v; (1.6-10)
i= l while the Iatter is extremely u
Further simplification results if we accept the convention that repeated If we jump ahead briefiy
subscripts (or indices, as the subscripts i are called) are summed from 1 to 3. viscous dissipation to interna!
This is called the summation convention, and allows us to write Eq. 1.6-10 as t T represents the total stress t
(1.6-10 sans serif type whenever possible.
lntroduction Chap. 1 Sec. 1.6 Vectors 27

Index notation is most useful in dealing directly with the scalar components
of vectors. For example, the use of numerical subscripts allows us to express
(1.6-6a) Eqs. 1.6-7 as
v{ = v1 cos en + v2 cos e21
+ Va cos ea (1.6-12a)
1

V~ = V1 COS e12 + V2 COS e22 + Va COS ea 2


(1.6-6b) (1.6-12b)
dimensions is straightforward but
v~ = v cos e a+ v cos e a+ Va cos e
1 1 2 2 (1.6-12c)
33

where it is understood that the second subscript on e refers to the primed


cos e + v. cos e.,.
11 .,. (1.6-7a)
(') coordinate system. The summation convention can be used to reduce
cos 0 + v. cos e.
1111 11 (1.6-7b) Eqs. 1.6-12 toa more compact form.
cos el/. + v. cose... (1.6-7c)
v{ = vi cos eil (1.6-13a)
called the direction cosines for the
these equations an ordered arrange- v~ = vi cos Oi 2 (1.6-13b)
to the subscript on the velocity V~ = Vi COS eia (1.6-13c)
the second always identical to the
Although i was used as the repeated index in each case, any index-such as
z' system. This order will allow us to
l.llrrln<>,rt form if we make use of index
k,j, n, etc.-<:ould have been used to obtain the same result. Since a repeated
index may be represented by any letter, it is often called dummy index. A
single, or unrepeated, index is called a free index, and it may assume any value
from 1 to 3. Using the convention for a free index, we can simplify Eqs.
1.6-13 further by writing
(1.6-14)
coordinate axes of a rectangular
z, and the three scalar components Certainly Eq. 1.6-14 is a far more compact representation ofthe information
It would be just as satisfactory to presented in Eqs. 1.6-12; however, our objective in introducing this notation
and Xa, and the scalar components as is not to produce a "slick" method of writing equations. In the treatment of
can be expressed as stress (Chap. 4), in Newton's law ofviscosity (Chap. 5), and in the mechanical
=i (1.6-8a) energy equation (Chap. 7), we shall find index notation invaluable, primarily
because it saves us from getting bogged down in a ponderous algebraic effort.
=j (1.6-8b)
In Gibbs notation, the velocity vector would be expressed as v; however,
=k (1.6-8c) index notation may be used to express the same vector as vi. The'point to be
vector vas remembered here is that the symbol v represents three scalar components
relative to the x,y,z-coordinate system, and we can represent these three scalar
e<uV1 + e< 2lv2 + e(a)Va (1.6-9)
components just as well by the symbol vi (or V;, etc.). The advantage of
express a vector in terms of its scalar Gibbs notation is that by tradition (certainly not by logic) it gives the feeling
of a directed quantity, while index notation gives the feeling of a group of
three numbers. The former is more closely related to our physical intuition,
(1.6-10)
while the latter is extremely useful in treating complex vector operations.
the convention that repeated If we jump ahead briefty to Chap. 7, we find that in Gibbs notation the
i are called) are summed from 1 to 3. viscous dissipation to interna! energy is represented by Vv:T,t while the rate
and allows us to write Eq. 1.6-10 as
t T represents the total stress tensor. Tensors in Gibbs notation will be represented by
(1.6-10 sans serif type whenever possible.
28 lntroduction Chap. 1 Problens

of surface work per unit volume is V (v T). These two abstract symbols (b) Express in square c1
1.6 x I0- 4 poise-ft 3/l
represent two very different physical phenomena, and it is helpful to the student
that these symbols actually look different. In index notation, these terms are Ans: I0- 2 cm 2/sec
represented by (e) Express~ in cubic fec
ov.) -o
gal/min.
(-
OX
~~ and
OX
(v~ 1)
Ans: 0.267 ft 3/sec
respectively. Although the symbols are indeed different and therefore repre- 1-3. Given the following con<l
sent different quantities, the difference is not so apparent, nor so easily
grasped by the reader intent on assembling a stream of information into sorne
coherent whole.
Although index notation lacks something in quick visual communication,
it is an extremely powerful tool for carrying out complex vector and tensor
operations. Therefore, we may sometimes begin an analysis in Gibbs calculate the Reynolds m
notation, switch to index notation when the going gets tough, and then Ans: 6.6 x 104
express our final result in Gibbs notation. This switching of notation should 1-4. On the assumption that
not disturb the student, for it is done in an effort to improve the exposition. 1.5-2, derive a relationshi
Bridgman15 has pointed out that "mathematics is merely a precise verbali- cylinder, and the viscosit
zation," and even the nonmathematician is aware that sorne Ianguages gion), and w 0 (angular ve
(notations) are more suitable than others for expressing certain thoughts or
1-5. G iven that T11:x is a const
ideas. For example, the opera singer prefers to sing in Italian because of the
1.5-1, solve Eq. 1.5-7 subj
soft vowel endings, while the dialectician prefers English for its scope and
variety of meaning. 16 Thus, the opera singer may perform in Italian but B.C. 1:
carry out his contract negotiations in English. Similarly, we will use Gibbs B.C.2:
ntation for quick visual communication and index notation for the more to determine the velocity
difficult analytic steps. Although index notation is really unnecessary until Uo.
we reach Chap. 4, examples and problems will be given throughout the first
1-6. lf the critica! stress T0 fo
three chapters in order to accustom the student to it.
87lbm/ft3 , compute the m
were applied to a vertical
Ans: 0.008 in.
PROBLEMS
1-7. As an approximation, the
1-l. The "weight" of an object is defined as the force (in pounds-force) exerted tires slide on a wet pave
on the object by gravity. lf the mass of an object is 10.0 lbm, what is its trated in Fig. 1.5-1. If th
weight on the earth and on a planet where the gravitational acceleration is pavement is 48 in. 2, and th
10 ft/sec 2 ? exerted on the automobile
Ans: 3.11 lbr relationship is valid even
required for a 2000 lbm c
1-2. (a) Express in terms of atmospheres the pressure p = 12.5 poundals/ft 2 poise for viscosity of wate
Ans: 1.84 x 10-4 atm 1-8. Rework Prob. 1-5 for th
15. P. W. Bridgman, The Nature of Thermodynamics (Cambridge, Mass.: Harvard fluid is
University Press, 1941).
16. D. Conrad, "Lectures on Modem Languages" (from a series ofunpublished lectures
delivered in Chicago, 111., in 1962).
lntroduction Chap. 1 Problems 29

(v T). These two abstract syrnbols (b) Express in square centimeters per second the kinematic viscosity v =
~en<)me:na, and it is helpful to the student 1.6 x 10-4 poise-ft3/lbm.
In index notation, these terrns are Ans: 10- 2 cm 2/sec
(e) Expressj in cubic feet per second the volumetric flow rate Q = 120
a
; - (vT;;)
galfmin.
u X; Ans: 0.267 ft 3/sec
indeed different and therefore repre- 1-3. Given the following conditions for flow in a pipe,
is not so apparent, nor so easily
a strearn of inforrnation into sorne <v.> = 10 ft/sec
D =1 in.
p = 80 lbm/ft3
in quick visual cornrnunication,
1' = 0.015 poise
out cornplex vector and tensor
begin an analysis in Gibbs calculate the Reynolds number.
the going gets tough, and then Ans: 6.6 x 104
This switching of notation should
1-4. On the assumption that h ~ r0 for the Couette viscometer shown in Fig.
an effort to irnprove the exposition.
1.5-2, derive a relationship between the torque :r required to drive the outer
~ulcu ... u-" is rnerely a precise verbali-
cy1inder, and the viscosity p. in terms of r0 , h, L (length of the annular re-
is aware that sorne Ianguages gion), and w0 (angular velocity) .
for expressing certain thoughts or
to sing in Italian because of the 1-5. Given that T11., is a constant for the plane Couette flow illustrated in Fig.
1.5-1, so1ve Eq. 1.5-7 subject to the boundary conditions (B.C.),
prefers English for its scope and
singer rnay perforrn in Italian but B.C. 1: v., =o, y =0
English. Sirnilarly, we will use Gibbs B.C.2: v., = u0 , y =h
and index notation for the more to determine the velocity distribution and the relationship between T11., and
notation is really unnecessary until Uo.
will be given throughout the first
student to it. 1-6. If the critica! stress T0 for a Bingham fluid is 0.058 lb1/ft2 and the density is
87 lbm/ft 3 , compute the mnimum film thickness that would result if this fluid
were app1ied to a vertical wall.
Ans: 0.008 in.
1-7. Asan approximation, the fluid motion that occurs when braked automobile
as the force (in pounds-force) exerted tires slide on a wet pavement might be considered comparable to that illus-
of an object is 10.0 lbm, what is its trated in Fig. 1.5-1 . lf the effective contact area between the tires and the
where the gravitational acceleration is pavement is 48 in. 2, and the water film is 0.003 in. thick, express the drag force
exerted on the automobile as a function of its velocity u0 Assuming that thi~
relationship is valid even when u0 is changing, compute the time and distance
required for a 2000 lbm car to dece1erate from 60 mph to 5 mph. Use 0.011
the pressure p = 12.5 pounda1s/ft 2 poise for viscosity of water.
1-8. Rework Prob. 1-5 for the case where the temperature distribution in the
fluid is

(from a series of unpublished lectures


lntroduction Chap. 1 Problems
30

and the viscosity is related to the temperature by 1-13. The vector e is defined as a sea

~ ~oo exp (~~)


=
Write out the scalar componen
index notation.
This problem requires a numerical integration of Eq. 1.5-7, which is best
handled by forming the dimensionless equation 1-14. Express the scalar S = e (A -t
1-15. Given a second order system a;.
a11 = 6,
a21 = 5,
where U:r =Y;p.00 1hT 11 .c
aal = -4,
y
Y=- Compute the term a 1ka2k.
h
Ans: 24.
which is subject to the boundary condition,
B.C. 1: Ux =0, Y=O
Sorne interesting results can be obtained by setting !lEfRT0 equal to 5.0, and
letting !lTfT0 take on the values 0.0, 0.1, 0.2, and 0.3.
1-9. What is the angle between the vector, A = 1Oi + 5j, and the x-axis?
1-10. Given a counterclockwise rotation
y
of 30 around the z-axis as shown y'
in Fig. 1-10, express i' and j' in x'
terms of i and j.
v'3
Ans: i' = 2 i +ti
v'3
j' = -ti+- 2
j
Fig. 1-10
1-11. The three algebraic equations
x + 5y + 3z = 10
2x + 4y + 4z = 5
3x + 26y = 7
can be expressed in a single equation as

where x 1 = x, x 2 =y, and x 3 = z. List the values of the nine terms rep-
resented by a; 1, if the three terms represented by b 1 are b1 = 10, b2 = 5, and
b3 = 7.
1-12. Write out the three scalar components of the vector A +B. Express the
equation
e=A+B
in index notation.
lntroduction Chap. 1 Problems 31

by 1-13. The vector C is defined as a scalar, S, times the vector, B.


e =SB
Write out the scalar components of the vector C and express this result in
integration of Eq. 1.5-7, which is best index notation.
equation 1-14. Express the scalar S = C (A + B) in index notation.
1-15. Given a second order system a;; such that
a 11 = 6, a12 = 3, a 13 = -3
a21 = 5, a22 =O, a2a = 2
aal = -4, aa2 = 1, a 33 = 7
Compute the term alka2k.

Ans: 24.

by setting !lE/RT0 equal to 5.0, and


0.1, 0.2, and 0.3.
A = lOi + 5j, and the x-axis?

y' y

Fig. 1-10

List the values of the nine terms rep-


by b; are b1 = 10, b2 = 5, and

of the vector A + B. Express the


Sec. 2.1 The Material Volume

The "body" referred to in


material, such as a steel ba
tional field, or all the asp.
polyethylene bag, or a we
around in the turbulent atm ,
Fluid Statics and or bodies, remains the same
It follows immediately fro~
One-Dimensional 2 mass of a material volume i!
In applying Eq. 2.1-1 to
Laminar Flow define our system at sorne ti
radius r0 , we might imagine
system with a smooth clos
molecules. If we wish to wo
we do, we must restrict o
motion can be neglected. A
the gross dimensions of the S
the mean free path of the mo
may correctly speak of a mat
size and shape but exchangin
material volume as "Y m(t), a
The subscript m reminds us
surface, both of which may
have occasion to discuss vq
volumes which move throug
"P'"a(t). The surface areas ass
In this chapter we shall examine the two simplest types of fluid motion: and da(t), respectively. Thes
fluids at rest and steady, one-dimensional, laminar flow . This will serve to are designated by capital se
introduce the student to sorne of the fundamental concepts of fluid mechanics, significance. In addition, it a
without getting involved in the mathematical details required to examine sectional areas and portions
many of the more complex flows encountereq in practice. We will first clearly distinguished by the u
present a general formulation of the linear momentum principie and then use
this result to derive the equations offluid statics, which are simply a special
form ofthe more general equations ofmotion to be presented in Chaps. 4 and The linear momentum pr
5. The second topic covered in this chapter, one-dimensionallaminar flow,
is another example of a special form of the general equations of motion.
Consider now the differen
The mass, dM, contained in
*2.1 The Material Volume

As previously discussed, the linear momentum principie states that and the momentum (mass
the time rate of change ) = {the force acting)
therefore,
{of momentum f the body on the body
(2.1-1)

32
Sec. 2.1 The Materll Volume 33

The "body" referred to in Eq. 2.1-1 consists of sorne fixed quantity of


material, such as a steel hall moving under the action of the earth's gravita-
tional field, or all the asparagus and butter sauce contained in a sealed
polyethylene bag, or a weather balloon filled with helium being bounced
around in the turbulent atmosphere. The material within these three systems,
or bodies, remains the same; therefore, they may be called material volumes.

2
lt follows immediately from the principie of conservation of mass that the
mass of a material volume is constant.
In applying Eq. 2.1-1 to a fluid, sorne conceptual difficulties arise. If we
define our system at sorne time, t = O, as all the material within a sphere of
radius r 0 , we might imagine that at sorne other time we could not locate the
system with a smooth closed surface, owing to the random motion of the
molecules. If we wish to work within the realm of continuum mechanics, and
we do, we must restrict ourselves to cases where this random molecular
motion can be neglected. As we mentioned in Chap. 1, we may do so when
the gross dimensions of the system under consideration are much larger than
the mean free path of the molecules in the fluid. Under these conditions, we
may correctly speak of a material volume which may be continuously changing
size and shape but exchanging no mass with its surroundings. We denote a
material volume as "Y m(t), and the surface of a material volume as d m(t).
The subscript m reminds us that we are discussing a "material" volume and
surface, both of which may be functions of time. Later in the text we shall
have occasion to discuss volumes fixed in space, designated by "Y, and
volumes which move through space in an arbitrary manner, designated by
"Ya(t). The surface areas associated with these volumes are indicated by d
two simplest types of fluid motion: and da(t), respectively. These volumes and the surfaces associated with them
laminar flow. This will serve to are designated by capital script letters to alert the student to their special
concepts of fluid mechanics, significance. In addition, it allows us to use the symbol A to indicate cross-
1"""...'''-al details required to examine sectional areas and portions of c/osed surfaces, while closed surfaces are
in practice. We will first clearly distinguished by the use of d, d m(t), and da(t).
momentum principie and then use
statics, which are simply a special
to be presented in Chaps. 4 and The linear momentum principie
, one-dimensionallaminar flow,
the general equations of motion.
Consider now the differential volume of fluid dV illustrated in Fig. 2.1-1.
The mass, dM, contained in this differential volume is given by

dM = pdV (2.1-2)

and the momentum (mass times velocity) of the differential element is,
= (the force acting)
therefore,
on the body
(2.1-1)
dMv = pvdV (2.1-3)
34 Fluid Statics and One-Dimensional Laminar Flow Chap. 2 Sec. 2.1 The Material Volume

Ll F, Surface force We shall consider only the gravitationa


acting on LIA
the gravity vector. In treating the surf2
in terms of the stress vector, t<n> defin

t(n) = liJ
AA

where ~F is the force exerted by the s


In general, a scalar or a vector
spatial coordinates and time; howev
vector, for it also depends on the orie
of the surface in question. The subs
stress vector. Asan example of this
the solid bar under compression illus

dV, Differential
volume element

r,;, (t), Material volume


Fig. 2.1-1. A material volume.

We may now write the momentum of the material volumet as F


'
{momentum of the material volume} = f pv d V (2.1-4)
"''"m(l)
and Eq. 2.1-1 may be written
D
Dt
f {force acting
pv dV = on the material
}
(2.1-5)
Flg. 2.1-2. St

"!'" m(l) volume ask the following questions. What is


The derivative, D/ Dt, is called the material derivative and will be discussed in What is the velocity at point a? W
detail along with the total and partial time derivatives in Chap. 3. For the the question-What is the stress at 1
present, it is sufficient to say the material derivative simply indicates that we What is the stress at point a for a su
are taking the time rate of change of a material volume. To illustrate this point, we shall
The forces acting on a material volume of fluid consist of body forces- at point a having a normal j. Ifthe
such as gravitational, electrostatic, and electromagnetic forces-that act on is F and the cross-sectional area of t
a for a surface having a normal j
the mass as a whole and surface forces that act on the bounding surface ofthe
material volume. lf the body force per unit mass is represented by g, the
total body force is

{
body force exerted}
on the material =
f pg d V (2.1-6) where the plus or minus sign on t
vol u me depending on whether the normal to
"!'" m<O

t This type of terminology will be used throughout the text instead of the correct
statement, the momentum of the body having a configuration Ym(t).
Chap. 2
Sec. 2.1 The Material Volume 35

We shall consider only the gravitational body force; g will therefore represent
the gravity vector. In treating the surface force it will be. appropriate to work
in terms of the stress vector, t<n> defined by Eq. 2.1-7.

t<n> = lim (~F) (2.1-7)


AA-+0 ~A

where ~F is the force exerted by the surroundings on the area ~A.


In general, a scalar or a vector can be specified simply in terms of its
spatial coordinates and time; however, this is not the case with the stress
vector, for it also depends on the orientation (given by the normal vector n)
of the surface in question. The subscript (n) is thus used in denoting the
stress vector. As an example of this dependence upon n, we might consider
the solid bar under compression illustrated in Fig. 2.1-2. It is appropriate to
dV, Oifferentiol
volume element F

the material volumet as, F

volume} = J pv dV (2.1-4)
..Y,.(t)

force acting } Fig. 2.1-2. Stress in a solid bar.


{
on the material (2.1-5)
volume
ask the following questions. What is the temperatwre in the bar at point a?
derivative and will be discussed in What is the velocity at point a? What is the density at point a? However,
time derivatives in Chap. 3. For the the question-What is the stress at point a?-is incomplete. We must ask:
derivative simply indicates that we What is the stress at point a for a surface having a normal n?
a material volume. To illustrate this point, we shall consider the state of stress for the plane
ume of fluid consist of body forces- at point a having a normal j. Ifthe total compressive force acting on the bar
electromagnetic forces-that act on is F and the cross-sectional area of the bar is A, then the stress vector at point
that act on the bounding surface of the afora surface having a normal j is given by
unit mass is represented by g, the
t(l) = j (~) (2.1-8)
=f pgdv (2.1-6)
where the plus or minus sign on the right-hand side of Eq. 2.1-8 is used
..y,.(1)
depending on whether the normal to the surface is chosen as j. What about
throughout the text instead of the correct
a configura/ion rm(t).
36 Fluid Statics and One-Dimensional Laminar Flow Chap. 2 Sec. 2.2 Fluid Statics

the stress at point a acting on the surface having a normal k? In the absence the stress on a fluid element at rest n
of any ambient pressure, it is obvious that this stress is zero. under consideration . We now wish t
t(k) =o (2.1-9) shown in Fig. 2.2-1, subject to the res
to prove that the static stress in a flui
When discussing the forces exerted by one phase upon another, we can
easily become confused asto which phase we are referring. To eliminate this z
difficulty, we shall adhere to the following convention: t<> ata phase interface
will refer to the force per unit area exerted by the phase into which the normal
n points, on the phase for which n is the outwardly directed normal.
The net surface force acting on the material volume may be represented in
terms of the stress vector

{
surface force exerted }
by the surroundings =
f ~> dA (2.1-10)
on the material volume $ .,(t)

and Eq. 2.1-1 becomes

:t f pv dV =
-r,.(1)
f
-r,.(1)
pg dV + f ~>
$ ,.(1)
dA (2.1-11)

Time rate of change Body force Suace


of momentum force

Equation 2.1-11 represents the application of the linear momentum principie


to an arbitrary body of fluid. In the next section, we will examine the form
of this equation for a fluid at rest. Fig. 2.2-1. Static s

in Fig. 2.2-1 are all normal to the re


*2.2 Fluid Statics Table 2.2-1 indicates the force and
the four planes making up the tetrah1
In studying fluids at rest an engineer is confronted with two problems: Tabl
determination of the pressure field in the fluid and application of this STATIC STRESS O
knowledge to determine forces and moments on submerged surfaces. We
Plan e Area
will consider the first problem initially. Application ofEq. 2.1-11 to any body
of fluid at rest yieldst ABC 6-A.

O= f pg dV + f tea> dA (2.2-1)
BCD
ADC
6-A.
6-A.
-r d ABD 6-A.
Note that in this case the limits of integration for the volume and surface
integrals are independent of time. Application of Eq. 2.2-l to the t,
integrals of the stresses in terms of a
The stress vector for a static fluid
yields
O = pg t:. V - n(p..) t:.A .. + i(p.) .
The commonly accepted definition of a fluid is that it will deform con-
tinuously under the application of a shear stress. In light of this definition, *Here we have used Px to represen! the pr
AAx with a similar interpretation a?plyin
t A scalar zero in a vector equation should be read as the null vector. scalar quantities not to be confused w1th the
and One-Dimenslonal Laminar Flow Chap. 2 Sec. 2.2 Fluid Statics 37

having a normal k? In the absence the stress on a fluid element at rest must always act normal to the surface
that this stress is zero. under consideration . We now wish to apply Eq. 2.2-l to the tetrahedron
(2.1-9) shown in Fig. 2.2-1 , subject to the restriction that all shear stresses are zero,
by one phase upon another, we can to pro ve that the static stress in a fluid is isotropic. The vector forces shown *
we are referring. To eliminate this z
convention: t<a> at a phase interface
by the phase into which the normal
the outwardly directed normal.
material volume may be represented in

(2.1-10)

(2.1-11)

Surface
Coree
X
of the linear momentum principie
next section, we will examine the form
Fig. 2.2-1. Static stress on a tetrahedron.

in Fig. 2.2-l are all normal to the respective surfaces on which they act, and
Table 2.2-l indicates the force and the outwardly directed unit normal for
the four planes making up the tetrahedron.
is confronted with two problems:
in the fluid and application of this Table 2.2-1
moments on submerged surfaces. We STATIC STRESS ON A TETRAHEDRON
Application ofEq. 2.1-11 to any body Plane Are a Normal Stress vector

ABC ~An n -npn


(2.2-1) BCD ~Az -i ipz
ADC ~A. -J jp.
ABD ~A. -k kp.

Application of Eq. 2.2-l to the tetrahedron, and expression of the area


integrals of the stresses in terms of an average stress and an area,
yields

of a fluid is that it will deform con-


O = pg AV- n(pn) AAn + i (p.,) AA., + j (p,) AA + k (p.) AA,
11 (2.2-2)
stress. In Iight of this definition, *Here we have used Px to represent the pressure acting on the surface area designated by
AAx with a similar interpretation applying to Py and AAy, etc. Both Px and AAx are
be read as the null vector. scalar quantities not to be confused with the component of a vector such as nx = in.
38 Fluid Statics and One-Dimensional Laminar Flow Chap. 2 Sec. 2.2 Fluid Statics

To proceed with our analysis and prove that Pn =p.,= p11 =p. we must Air
express the areas dA.,, dA 11 , dA. in terms of dAn. These expressions are given
by, Z=Lr---~----------~
dA., = i n dAn = n., dAn (2.2-3a)
dA 11 = j n dAn = n11 dAn (2.2-3b)
dA. = k n dAn = n. dAn (2.2-3c)

i.e., the areas of the coordinate planes are the projections of the oblique area,
dAn, onto the coordinate planes. The proof of Eqs. 2.2-3 is a straightforward
problem in analytic geometry and will not be given here.
Substitution of Eqs. 2.2-3 into Eq. 2.2-2 yields
(2.2-4)
Wot
Ifwe divide by dAn and take the limit dAn -4- O, we note that dVjdAn -4- O,
and Eq. 2.2-4 reduces to

O= lim {pg d V - [n(pn) - in.,(p.,) - jn 11 (p 11 ) - kn.(p.) 1}


AA,.-+0 dAn
= -npn + .m.,p., + JnvPv
. + k n.p. (2.2-5)

Here the area averages have been replaced with point values in accordance element shown in Fig. 2.2-2 and illust1
with d V -4- O. lf we express the unit normal n in terms of its scalar com- differential cube shown in Fig. 2.2-3 to
planes, and designating the lengths of
ponents,
apply Eq. 2.2-1 to this cubeto yield
n = in., + jn11 + kn. (2.2-6)
we may write Eq. 2.2-5 as O = pg Llx Lly,

O = in.,(pn -p.,) + jnv(pn - p11 ) + kn.(pn -p.) (2.2-7)


Now, if a vector is equal to zero, it readily follows (by forming the dot product
with i, j, and k, respectively) that the scalar components of the vector are
zero. Thus, Eq. 2.2-7 leads us to the conclusion that p.,= Pn p11 = Pn and
p. = Pno and our proof is complete.
Dropping the subscript n, we express the stress vector acting on any
arbitrary surface as
t<> = -np (2.2-8)

where n is the outwardly directed unit normal. Equation 2.2-8 indicates that
r
llz
11
1

L.~-:
the magnitude of the stress vector is given by the pressure, and the direction
............ .1----
is opposite to that of the unit normal. The minus sigo in Eq. 2.2-8 is a mani-
.........
........
festation of the convention that t<> represents the stress exerted on the system ........
by the surroundings when n is the outwardly directed normal for the system. ~+----- llx - -
We now wish to determine the pressure in the fluid illustrated in Fig.
2.2-2. Our first step is to apply Eqs. 2.2-1 and 2.2-8 to the differential volume Fig. 2.2-3. Differet
and One-Dimensional Laminar Flow Chap. 2
Sec. 2.2 Fluid Statics 39

prove that Pn = P~ = p11 = Pz we must A ir


of dAn. These expressions are given
Po= Atmospheric pressure

dAn= n~dAn (2.2-3a) Z=Lr---------------------~--------~


dAn= n11 dAn (2.2-3b)
dAn= n.dAn (2.2-3c)

are the projections of the oblique area,


proof ofEqs. 2.2-3 is a straightforward
not be given here .
. 2.2-2 yields \ Differentiol volume element

(2.2-4)
Water
dAn --.. O, we note that d V/ dAn _.. O,

- in~(p~) - jn11 (p11) - knz<P.>1} Sol id


Fig. 2.2-2. A fluid at rest.
+ kn.pz (2.2-5)

with point values in accordance element shown in Fig. 2.2-2 and illustrated in detail in Fig. 2.2-3. Taking the
normal n in terms of its scalar com- differential cube shown in Fig. 2.2-3 to have its sides parallel to the coordinate
planes, and designating the Iengths of the edges by dx, dy and dz, we may
(2.2-6) apply Eq. 2.2-1 to this cubeto yield

O = pg dx dy dz - Jnp dA (2.2-9)
.111
(2.2-7)
follows (by forming the dot product n= +k
scalar components of the vector are
conclusion that P~ = Pm p11 = Pm and

the stress vector acting on any

(2.2-8)
normal. Equation 2.2-8 indicates that
by the pressure, and the direction
The minus sign in Eq. 2.2-8 is a mani-
the stress exerted on the system
directed normal for the system.
in the fluid illustrated in Fig.
1 and 2.2-8 to the differential volume Fig. 2.2-J. Differential volume element.
Fluid Statics and One-Dimensional Laminar Flow Chap. 2 Sec. 2.2 Fluid Statics

Evaluation of the area integral for the six sides of the cube gives potential energy of the fluid. The gra
Forces on the g = ig.,
O = pg !1x !1y !1z - (e -ip !1y !1z)., + (ip !1y !1z).,+&., x-surfaces
where the scalar components g.,, gil,
Forces on the
+ (-jp !1x !1z) + (jp !1x !1z) +A 1
11 11 y-surfaces
(2.2-10) to b~ given by
e/>= -(xg.,
Forces on the
+ (-kp uX u y). + (kp ux uy)z+Az
A A A A }
.,
z-sur.aces it follows that minus the gradient of

The convention used to describe the surfaces is such that an "x-surface" is


one for which the x-coordinate is constant and the unit outwardly directed
-Ve/>= (i~+i~-fl
ax ay
normal takes on the values i. lfwe divide Eq. 2.2-10 by !1x !1y !1z and take
the Iimits !1x- O, !1y _.O, and !1z _.O, we obtain = ig., + jg11 +k
=g
O= pg- {l . [(Pio:+Ao:-
lm
!1x
Ao:-+o
PI"')]
1
This method of representing g is pos si
Equation 2.2-13 may be written in te
+ lim [i(PI11+A11- Pl11)] (2.2-11)
o= p~
A11-+0 f1y
If the density is constant, this equati<J
+ lim [k(Piz+Az- PI)]}
Az-+0 !1z
Each of the limits is the definition of a partial derivative, and Eq. 2.2-12
results from the limiting process. Since this result holds everywhere in
constant, and we may write
0 = pg _ ( axop + /Poy + k op)
oz (2.2-12) p=
Here we have integrated the equation
This equation may be written in the more compact form, of integration. Use of the potential
O= pg- Vp (2.2-13) of Eq. 2.2-13 is satisfying from th
solution of this equation deserves a t1
where V is a vector operator which takes the form student's notion of static fluid syste
For the system shown in Fig. 2.2-
V .a .a +ka-
=-+J- (2.2-14)
ox oy oz g
in rectangular coordinates. In index notation, this result would be expressed, where g is the magnitude of the gra

o= pgi - -
op (2.2-15)
into Eq. 2.2-13 yields the three scala
oxi op
Traditionally, the term Vp is called the gradient of the pressure and denoted ox
by "grad p"; however, the use of the symbol V is more informa ti ve and will op
be used here. oy
Very often we will find it helpful to represent the gravity vector in terms
of the negative of the gradient of a scalar, cp. This function is called the op =
gravitational potential function because it represents the gravitational oz
One-Dimensional Laminar Flow Chap. 2 Sec. 2.2 Fluid Statlcs -41

six si des of the cube gives potential energy of the fluid. The gravity vector g may be expressed

( 1p
A
uy A
uz)~ 4 .,
Forces on the g = ig., + jg!l +kg. (2.2-16)
~ x-surfaces
where the scalar components g.,, g11 , and g. are constants. If we choose 4>
Forces on the
y-surfaces (2.2-10) to be. given by
4> = -(xg., + yg11 + zg.) (2.2-17)
(kp Llx Lly) } Forces on the
z+Az z-surfaces it follows that minus the gradient of 4> is equal to g.
surfaces is such that an "x-surface" is
-Vtfo = (~ + j~ + k~)(xg., + yg11 + zg.)
and the unit outwardly directed
Eq. 2.2-10 by Llx .iy ~ and take
ax ay az
O, we obtain = ig., + jgll + kg. (2.2-18)

i( PI.,+A~: PI.,) J =g
This method of representing gis possible only because gis a constant vector.
Equation 2.2-13 may be written in terms of tf>.
~)] (2.2-11)
O= pVtf> + Vp (2.2-19)

~)]} If the density is constant, this equation may be written as

a partial derivative, and Eq. 2.2-12


O= V(ptf> + p) (2.2-20)
Since this result holds everywhere in the fluid, the term ptf> + p must be a
constant, and we may write
(2.2-12) p=C-ptf> (2.2-21)
Here we ha ve integrated the equations of fluid statics, and C is the constant
of integration. Use of the potential function to obtain the integrated form
(2.2-13) of Eq. 2.2-13 is satisfying from the mathematical point of view, but the
solution of this equation deserves a treatment more closely connected to the
student's notion of static fluid systems.
For the system shown in Fig. 2.2-2, the gravity vector is
(2.2-14)
g= -kg (2.2-22)
this result would be expressed,
where g is the magnitude of the gravity vector. Substitution of Eq. 2.2-22
op into Eq. 2.2-13 yields the three scalar equations
i-- (2.2-15)
OX
op =o (2.2-23a)
gradient of the pressure and denoted ox
V is more informative and will
op =o (2.2-23b)
represent the gravity vector in terms oy
, tfo. This function is called the op
- = -pg (2.2-23c)
it represents the gravitational oz
42 Fluid Statia and One-Dimensional Laminar Flow Chap. 2 Sec. 2.3 Barometers

These three scalar components of the original vector equation are quite vector is a continuous function. Both ti
obviously expressed by the index notation of Eq. 2.2-15, where g 1 = g 2 =O postulate discussed in Sec. 1.1. Boundat
and g1 = -g. lntegra,~on of Eq. 2.2-23c gives of the second rule-i.e., pressure is a e
p = -pgz + C(x,y) (2.2-24)
where the constant of integration may be a function of x and y; however, 2.3 Barometers
Eqs. 2.2-23a and b indicate that the pressure is neither a function of x nor y,
and we write Eq. 2.2-24
A barometer is a device for mea
p= -pgz+ e (2.2-25)
atmosphere. The one illustrated in Fi
A single boundary condition is needed in order to determine the constant e at one end and immersed in the bar
and thus specify the pressure everywhere in the fluid. lt is obtained by obtained by filling the tube with the
recognizing that the pressure at z = L is the atmospheric pressure p 0 We closing the open end (the thumb does 1
may write pool of barometer fluid. The fluid v;
pressure there is the vapor pressure
B.C. 1: p=po, z=L (2.2-26) p.,p. For mercury the vapor pressure
the application of which to Eq. 2.2-25 yields, at room temperature is approximately
3 x 10-t .a tm and may be considered
PI-L =Po= -pgL+ e (2.2-27) to be zero. The differential equation
for the pressure is
and the constant of integration is
op
e=p0 + pgL (2.2-28) oz = -pagg (2.3-1)

Substitution of e into Eq. 2.2-25 yields the final expression for the pressure. which is integrated to give

P =Po + pg(L - z) (2.2-29) p = - Pa 1gz +e (2.3-2)


The boundary condition is
This problem is such a simple one that the student can easily have missed
the significance of the various steps; therefore, a review will be helpful. B.C.1:
p =p,p ~o, z = h (2.3-3)
l. The linear momentum principie and Eq. 2.2-8 were applied to a
differential volume element to develop the equations of fluid statics. the application of which to Eq. 2.3-2
2. The equations were solved to yield Eq. 2.2-25. gives
3. A boundary condition was specified and applied to Eq. 2.2-25 to p = Pugg(h - z) (2.3-4)
obtain an expression for the pressure. Since the pressure at z = O is the
lt may not seem so to the student, but the final step in this general process atmospheric pressure, p0 , we write
is usually the most troublesome. The differential equations of motion can Po = Paggh (2.3-5)
be derived once and for all, and methods of solution (for those cases which
can be solved) are neatly tabulated in mathematics and fluid mechanics Because of the universal use of mere
texts; however, boundary conditions are specified mainly on the basis of pressure, it is often reported in terms o
physical intuition, and thus present a more diffi.cult problem. Things are pressure varies from day to day, the,
not quite as bad as they may seem, for there are two rules which guide us in
t Taking the pressure to be continuous n
specifying boundary conditions: velocity is a continuous function; the stress liquid interface.
and One-Oimenslonal Laminar Flow Chap. 2
Sec. 2.3 Barometers 43
the original vector equation are quite vector is a continuous function. Both these ideas follow from the continuum
of Eq. 2.2-15, where g1 = g 2 =O postula te discussed in Sec. 1.1. Boundary condition 1 results from application
gives
of the second rule-i.e., pressure is a continuous function for fluids at rest.t
(2.2-24)
be a function of x and y; however,
2.3 Barometers
pressure is neither a function of x nor y,

(2.2-25) A barometer is a device for measuring the absolute pressure of the


atmosphere. The one illustrated in Fig. 2.3-1 consists of a single tube closed
in order to determine the constant e at one end and immersed in the barometer fluid. Such a system may be
in the fluid. lt is obtained by obtained by filling the tube with the barometer fluid (usually mercury),
L is the atmospheric pressure p 0 We closing the open end (the thumb does nicely), and immersing the tube in the
pool of barometer fluid. The fluid vaporizes in the closed end, and the
pressure there is the vapor pressure
z=L (2.2-26)
p.,,. For mercury the vapor pressure vp

at room temperature is approximately


3 x 10-e atm and may be considered z =h
= -pgL+ e (2.2-27) to be zero. The differential equation
for the pressure is

(2.2-28) (2.3-1)

the final expression for the pressure. which is integrated to give


(2.2-29) p = - PHggz +e (2.3-2)

that the student can easily have missed The boundary condition is
therefore, a review will be helpful. B.C.l:
and Eq. 2.2-8 were applied to a p =p.,,~ o, z=h (2.3-3)
develop the equations of fluid statics. the application of which to Eq. 2.3-2 z=O
Eq. 2.2-25. gives
and applied to Eq. 2.2-25 to p = PHgg(h - z) (2.3-4)
Since the pressure at z = O is the
the final step in this general process atmospheric pressure, p 0 , we write
Fig. 2.3-1. Mercury barometer.
differential equations of motion can (2.3-5)
Po= PHggh
of solution (for those cases which
in mathematics and fluid mechanics Because of the universal use of mercury barometers to measure atmospheric
are specified mainly on the basis of pressure, it is often reported in terms of h, or inches of mercury. Atmospheric
a more difficult problem. Things are pressure varies from day to day, the average being 29.92 in. Hg. This value
there are two rules which guide us in
is a continuous function; the stress t Taking the pressure to be continuous neglects any effect of surface tension at the gas-
Iiquid interface.
Fluid Statics and One-Dimensional Laminar Flow Chap. 2 Sec. 2.-4 Manometers

is referred toas one standard atmosphere, and we may use Eq. 2.3-5 to deter- Both these conditions are derived from
mine that continuous function.t Solution of the tw
p0 (standard) = 14.696lbr/in. 2
p = -:pgz
This quantity is often written 14.696 psia (pounds per square inch absolute).
P2 = -p~Z

Application of boundary condition 1 giv


2.4 Manometers
Pllz-111 = Po = -
Manometers are devices which make use of columns of liquid to deter- or
mine pressure differences. The simplest type is the U-tube illustrated in C1 =Po+
Fig. 2.4-1. The manometer fluid must be immiscible with the fluid in the
and the pressure in fluid 1 is
/ Open tothe
/ otmosphere P1 =Po+ P~
Application of boundary condition 2 giv

P2!z=ll1 =- P2Kh1 + C2 = P1l


- -z=hz or
Pressure tonk

and the pressure in fluid 2 is

- - z=h 1 P2 = Po + p~(h1 - z
The gauge press.ure, pg, is defined as the
pheric pressure; therefore, the gauge pre
_ _ _ _ __ _ z=O
pg = (p2 -Po) = p~(h
Fl.l.4-l. U-tube manometer.
If the fluid in the tank is a gas, the
tank, and its density, p1 , must be greater than the density of the fluid in the usually be much larger than that of the
tank, p 2 , or there will be a tendency for it to replace the fluid in the tank. pressure is
The equations for the pressure in the two fluids are pg ~ pg(h
op
- = -pg (2.4-1)
oz Manometer calculation

aP2 = -p 2g (2.4-2) The fluid in the tank is a gas and a me


oz the pressure. The readings on the man01
The boundary conditions for this system are
h = 3.25
ft} or (h
B.C. 1: P1 = Po z = h3 (2.4-3) h3 =5.17ft
B.C. 2: P2 = P1 z = h (2.4-4) t Interfacial tensions are obviously being ne
and One-Dimensional Laminar Flow Chap. 2 Sec. 2.<4 Manometers 45

and we may use Eq. 2.3-5 to deter- Both these conditions are derived from the notion that the pressure is a
continuous function .t Solution of the two differential equations gives
= 14.696lb1/in.2
psia (pounds per square inch absolute).
p = -pgz +el (2.4-5)

P2 = -p~z + C2 (2.4-6)

Application of boundary condition 1 gives

use of columns of liquid to deter-


Pllz=/11 = Po = - pgha + C (2.4-7a)
or
type is the U-tube illustrated in (2.4-7b)
be immiscible with the fluid in the
and the pressure in fluid 1 is
to the
otmosphere P1 =Po+ pg(ha- z) (2.4-8)

Application of boundary condition 2 gives

or
(2.4-9b)

and the pressure in fluid 2 is


- - z = h1 (2.4-10)

The gauge press.ure, pg, is defined as the absolute pressure minus the atmos-
pheric pressure; therefore, the gauge pressure in the tank is
------z=O
manometer.
(2.4-11)

If the fluid in the tank is a gas, the density of the manometer fluid will
than the density of the fluid in the usually be much larger than that of the gas; thus, p1 :> p2, and the gauge
for it to replace the fluid in the tank. J~ressure s
two fluids are
(2.4-12)
(2.4-1)
Manometer calculation

(2.4-2) The fluid in the tank is a gas and a mercury manometer is used to measure
the pressure. The readings on the manometer are
are
z = h3 (2.4-3) h = 325 ft}
ha = 5.i 7 ft or
(h
a-
h)
1 = 1.92 tit
z = h (2.4-4) t Interfacial tensions are obviously being neglected here.
Fluid Statics and One-Dimensional Laminar Flow Chap. 2 Sec. 2.5 Forces on Submerged Plane Surfaces

The density of mercury is 13.5 gfcm3, and the gauge pressure is z_=L----------

p 11 = {(13.5 g/cm )(32.2 ft/sec~(1.92 ft)}


3
Original terms
2
{(~) (
3
X lbr sec ) (2.54 cm) (12 in.)}
453.6 g 32.2lbm ft in. ft z=t--------------
df= -npdA
Conversion factors

= (13.5)(1.92)(2.54)3(12) = 11.2lb /in.2


~ ~53~ f

To distinguish gauge pressure from absolute pressure (psia) we write


Fig. l.S-1. Surface force a
p11 = 11.2 psig (pounds per square inch gauge)
The net surface force exerted on the gat
by the liquid and by the surrounding atn
*2.5 Forces on Submerged Plane Surfaces
Fnet = - Jnp d
Now that we have learned to determine the pressure field for a fluid at .<I.

rest, we can direct our attention to the second problem noted in Sec. 2.2, i.e., The unit outwardly directed normal in
the calculation of forces on submerged surfaces. According to our previous just equal and opposite to n, and we m
definition, the force which the surroundings exert on a system is
force per unit area }
Fnet = - J(p
{acting on the system = t
<n>
= -np for a fluid at rest (2.5-1) ."'-
If the width of the gate is b, then
where n is the outwardly directed unit normal for the system. The total force dA=
F exerted by a fluid at rest on a solid surface is
where r is the distance measured alo
F = J~n) dA =- J np dA (2.5-2) force is now given by

where it must be remembered that n is the outwardly directed unit normal for
Fnet = - f
~=0
~=l/Bil

pg(

the solid-i.e., our convention is that n is the unit normal directed into the In order to evaluate the integral we ml!
phase which exerts the force. The first step in calculating the force and mom- terms of z. In the latter case we write
ent on the gate shown in Fig. 2.5-1 is to determine the pressure in the fluid
by means of the differential equation r = SI.
(Jp
-=-pg (2.5-3)
(}z dr =-
SI
and the boundary condition
and the force becomes
B.C. 1: p=po, z=L (2.5-4) t
Thus, Fnet = - ~
smO
J pg(L- z)nb
P =Po + pg(L - z) (2.5-5) o
and One-Dimensional Laminar Flow Chap. 2
Sec. 2.5 Forces on Submerged Plane Surfaces 47
, and the gauge pressure is z = L----------------t
.92 ro}
2
lbm ) ( lbr sec ) (2.54 cm) (12 in.)}
.6 g 32.2 lbm ft in. ft
3

z=t-------------------
L.
Conversion factors dF= -npdA
54 3 2
) (l ) = 11.2lbr/in. 2
.6)
absolute pressure (psia) we write
dF 0 = -n 0 podA

Fig. 2.5-1. Surface force and moment on a gate.


per square inch gauge)

The net surface force exerted on the gate, F net, consists of the force exerted
Plane Surfaces by the liquid and by the surrounding atmosphere,

the pressure field for a fluid at


, ........ llJ.......,
Fnet =- J
A
npdA- J
A
DoPo dA (2.5-6)

second problem noted in Sec. 2.2, i.e.,


The unit outwardly directed normal in the surrounding atmosphere, fto, is
surfaces. According to our previous just equal and opposite ton, and we may write Eq. 2.5-6
~-'"'' ........ 5 ., exert on a system is

= -np for a fluid at rest (2.5-1)


Fnet = - f
A
(p- Po)n dA (2.5-7)

If the width of the gate is b, then


normal for the system. The total force
surface is dA= b dr

=- J npdA (2.5-2)
where r is the distance measured along the gate from the hinge. The net
force is now given by
~=ltsin 8

the outwardly directed unit normal for


Fnet =- J pg(L- z)nb dr (2.5-8)
.,=o
n is the unit normal directed into the
step in calculating the force and mom- In order to evaluate the integral we must either put z in terms of r or r in
is to determine the pressure in the fluid terms of z. In the latter case we write
z
r=-
sin O
(2.5-3) dz
dr=-
sin O
and the force becomes
z=L (2.5-4)

pg(L- z) (2.5-5)
1
Fnet = - -.-
sm O
J' sm v
0
pg(L- z)nb dz = - . ll pgbt(L- {\
2}
(2.5-9)
o
Sec. 2.6 Forces on Submerged Curved Su
Fluid Statics and One-Dimensional Laminar Flow Chap. 2

The position vector r is best represent


The unit normal can be expressed in terms of its components
the gate, and the distance measured at
o = in., + jn11 + kn~ (2.5-10) r=
and we see from Fig. 2.5-2 that these components are given by Since A and o are orthogonal, we ma.x
n., =-sin(} (2.5-lla) r x o= rA.
n11 =O (2.5-11b) and Eq. 2.5-14 takes the form
lsil
n~ = cos (} (2.5-llc)
y= j f
o
(p

Putting the pressure difference p - Po

pgb
Y=[j-
It is often convenient to represent
resultant force. For this resultant fo
force system, it must act a point whi
distributed force system. When disc
point the center of pressure and den
the center of pressure is determined
i'x

nx =in = cos (1TI2+ 9) =-sin 8 where Y is the torque owing to the


ny=jn=O resultant force. Letting
nz =kn = cos 8 i' :

Flg. 2.5-2. Components of the unit normal. and substituting Eqs. 2.5-9 and 2.5-1
of pressure on the hinged gate.
The net force may be expressed as
kn,)(p~bt) (L- !.)
- t
Fnet = -(in.,+ (2.5-12) r= - .
sm () 2 sm
and the components are given by

F.,,net = pg( L- f)bt (2.5-13a)


*2.6 Forces on Submerged C
F,,net = - pg coto(L- f) bt (2.5-13b)
We first examine the problem of
The torque, 9", on the hinge of the gate is given by considering the force on the arbitrar
on the surface is
9" = f lslnB
r X [ -n(p - p0 )b] dr (2.5-14) F=

O j Force
where
Momentarm
and One-Dimenslonal laminar Flow Chap. 2 Sec. 2.6 Forces on Submerged Curved Surfaces

in terms of its components The position vector r is best represented in terms of a unit tangent vector to
in., + jn + kn.
11 (2.5-10) the gate, and the distance measured along the gate, r.
components are given by r = 'Ar (2.5-15)
Since 'A and o are orthogonal, we may write
(2.5-lla)
r X O = TJA X O = -jr (2.5-16)
(2.5-llb)
and Eq. 2.5-14 takes the form
(2.5-llc) lsln 8

9"' = j f (p - p0)br dr (2.5-17)


o
Putting the pressure difference p - p 0 . in terms of r and integrating, we ha ve

(2.5-18)

It is often convenient to represent a distributed force in terms of a single


resultant force. For this resultant force to describe properly the distributed
force system, it must act a point which gives rise to the same torque as the
distributed force system. When discussing hydrostatic systems, we call this
point the center of pressure and denote it by i'. The position vector locating
the center of pressure is determined by solving the equation
i' x F = 9"' (2.5-19)
+9)=-sin e where :Y is the torque owing to the distributed force system and F is the
resultant force. Letting
(2.5-20)
of the unit normal.
and substituting Eqs. 2.5-9 and 2.5-18 into Eq. 2.5-19, we locate the center
of pressure on the hinged gate.

f~
kn.)(p~bt)
Sin()
(L- !.)2 (2.5-12) - t 2L 3 (2.5-21)
r = sin () [ L _

(L-f)bt (2.5-13a)

coto( L- f)bt (2.5-13b)


*2.6 Forces on Submerged Curved Surfaces

We first examine the problem of forces on curved submerged surfaces by


considering the force on the arbitrary surface shown in Fig. 2.6-1. The force
on the surface is
(2.5-14) f opdA
F =- (2.6-1)
A

P =Po + pg(L - z)
arm
where
50 Fluid Statlcs and One-Dimensional Laminar Flow Chap. l lec. 1.6 Forces on Submered Curved

Since the projection of the surface


planes is

the surface area may be expressed i


n
odA= idA
The force acting on the solid surfa
areas to yield
F= -if .J..
pdAz

1------Y We do this because the pressure is


coordinates; therefore, it is necessa
these coordinates rather than in te
appliction of these ideas, we cons
force that the fluids exert on the s

Fl. 1.6-1. Projected areas for a curved surface.


F =- f
l
D

Atmospheric pressure =Po where n1 and n2 are the unit norma


The pressures in the two fluids are
p=po

z=L1T""""'------------.J Density =p.z


Pz =Po
Substitution of Eqs. 2.6-6 and .
Oenslty =: p,
product with the unit vector i give

zl--
Fz = i F=- f[
l

Sphere of rodius ro
imbedded In o thln woll
Upon examining Fig. 2.6-3, we sec
i 01
i Dz
Fl. 1.6-l. Force on a spherical surface.
Substitution of Eqs. 2.6-9 and 2.6
One-Dimenslonal Laminar Flow Chap. 1
S.C.l.6 Forcea on Submered Curved Surfac:ea SI

Since the projection of the surface area element dA on the three coordinate
planes is
i odA =dA, (2.6-2a)
JodA = dA 11 (2.6-2b)
kodA =dA. (2.6-2c)

, n the surface area may be expressed in terms of the projected areas as


o dA = 1 dA, + j dA" + k dA.
The force acting on the solid surface may be written in terms of the projected
areas to yield

F= -IJ pdA,-jf pdA11 - k f pdA.


(2.6-3)

(2.6-4)
~. ~. ~.

r-----y
We do this because the pressure is always given in terms of the x, y, and z
coordinates; therefore, it is necessary to perform the integration in terms of
these coordinates rather than in terms of the surface area .. To illustrate the
application of these ideas, we consider the system shown in Fig. 2.6-2. The
force that the fl.uids exert on the sphere is

: . . ..
. . F= - I
~1
Op dA - I
~.
02P2 dA (2.6-5)

. . #. where o1 and o 2 are the unit normals directed into fl.uids 1 and 2, respectively.
The pressures in the two fl.uids are given by

Pt =Po + p~(Lt - z) (2.6-6)

Pz =Po + P~K(L2 - z) (2.6-7)


Substitution of Eqs. 2.6-6 and 2.6-7 into Eq. 2.6-5 and taking the scalar
product with the unit vector 1 give

F, = i F = -J~1
[p0 + p g(L- z)]i o1 dA
1

(2.6-8)
Sphere of rodius r0
imbedded In o thln woll
-J [p0 + p~(L 2 - z)Ji o2 dA
~.

Upon examining Fig. 2.6-3, we see that


i . O dA = -dA, (2.6-9)
i o 2 dA -:- dA 11 . (2.6-10)
a spherical surface.
Substitution of Eqs. 2.6-9 and 2.6-10 into Eq. 2:6-8 and putting both terma
52 Fluid Sutics and One-Dimensional Laminar Flow Chap. l Sec. 1.7 Buoyancy Forces

lf we add and subtract p0 from the 1


we obtain
F~~e = [po + pg(L- t)]
Mean pressure

Thus, the resulting force may be wri


on the surface multiplied by the proj
of the force. We may use this resu
perpendicular! y to the gravity vector
ofz.

*2.7 Buoyancy Forces

The resultant force exerted on


buoyant force. The principie that
Archimedes (287-212 se), who sup
entering the pool of a public bat
Fl. 2.6-3. Projected areas for a spherical surface. buoyed up by a force equal to the wei
prove this by applying Eq. 2.6-1 to
under the same integral sign, this case, we consider a solid body
by a curve along which

... Fre = f {(po + pg(L- z)]- (p 0 + p~(L 2 - z)l} dAz


-.t. Thus, the normal to the solid surf
(2.6-11) The analysis can be easily extend
= f
4.
[(p1f.t - p2 LJg + (p 2 - pJgz] dA~~e dividing the volume into severa! sec
upper surface by z2 and the lower s
Since into two parts to give

the force may be written as


dA~~e = dydz (2.6-12)
F =- I D2PI.=
1
-l+ro v-+ V r!-<-ll

Fz = f f [(pL- pzLJg
1
+ (p2 - pJgz] dy dz (2.6-13)
Taking the dot product ofEq. 2.7-2
force
-l-ro v - V r!-<s-11
Integration with respect to y gives
-l+ro Since n k dA is the projection of tl:
F re= 2 f [(pf.t- P2LJg + (p2- pJgz].Jr~- (z- () 2
dz (2.6-14)
change the variables of integration 1
k n2 dA =dA,
-l-ro kn1 dA = - d
which we may integrate in tum to give
1. H. Rouse and S. Ince, History oj
F.= [(p1L1 - p2 LJg + (p2- pJg(]'"r~ (2.6-15) Inc., 1963), p. 16.
One-Dimensional Laminar Flow Chap. 2 Sec. 2.7 Buoyancy Forces 53

If we add and subtract p 0 from the right-hand side and rearrange the terms
we obtain

Mean pressure Arca Mean pressure Arca

Thus, the resulting force may be written in terms of the mean pressure acting
on the surface multiplied by tle projected area of the surface in the direction
of the force. We may use this result to quickly determine forces which act
perpendicular}y to the gravity vector provided the pressure is a linear function
ofz.

*2.7 Buoyancy Forces

The resultant force exerted on a body by a fluid at rest is called the


buoyant force. The principie that we wish to prove here is attributed to
Archimedes (287-212 BC), who supposedly discovered the phenomenon while
entering the pool of a public bath. 1 The principie states that "a body is
for a spherical surface. buoyed up by a force equal to the weight ofthe displaced fluid." We can easily
prove this by applying Eq. 2.6-1 to the solid body illustrated in Fig. 2.7-1. In
this case, we consider a solid body which may be separated inte two regions
by a curve along which
kn= O (2.7-1)
Thus, the normal to the solid surface lies in the x-y plane along this curve.
(2.6-11)
The analysis can be easily extended to more complex shapes simply by
dividing the volume into severa} sections. lf we designa te the position of the
upper surface by z 2 and the lower surface by zl> Eq. 2.6-1 may be separated
into two parts to give
(2.6-12)
F = - f D2PI.~ dA - f
.d.
Dpl-1 4A (2.7-2)

Taking the dot product ofEq. 2.7-2 with the unit vector k, we get the buoyancy
+ (p2- pJgz] dy dz (2.6-13) force
(2.7-3)

Since n k dA is the projection of the surface area on the x-y plane, we may
- pJgz]Jr~- (z- ()2 dz (2.6-14) change the variables of integration to
k n 2 dA = dA,, upper region (2.7-4a)
k D dA = -dA,, lower region (2.7-4b)
l. H. Rouse and S. Ince, History of Hydrau/ics (New York: Dover Publications,
+ (pz- pJg(]1Tr~ (2.6-15) lnc., 1963), p. 16.
S.C. 2.7 Buoyancy Forces
Fluid Sutlcs and One-Dimensional Laminar Flow Chap. l

p,

k=O The result is not difficult to extend


between two fluids. This situatio1
is left to the student.

The hydrometer

l
z

~.
A hydrometer uses the buo
densities of two fluids. In gener
Gravity
of densities, pf p 810 , is called the

Fl.l.71. Buoyancy force on a submerged solid body.

and Eq. 2.7-3 becomes A hydrometer floating first in wa

F.= - f Plz-s
.d..
dA.+ f PI-
.d..
dA.
(2.7-5) Fig. 2.7-3. Ifthe mass ofthe hyd1
the liquid is so small that the
buoyancy force, then Archimed~
Putting both terms under the same integral sign, and noting that
(2.7-6)
PI- - Pl=s = pg(zz - Z)
we obtain
F.= f
.d..
pg(z 2 - z 1) dA.
(2.7-7) where V8 0 and V0 n are the vol
1
tively. Division of Eq. 2.7-12 b:
But pg is a constant and the volume of the solid V, is 1=

V,= f
.d. a
(z 2 - z1) dA. (2.7-8)
and the specific gravity of the e
and the solid body is buoyed up by the "weight" of the displaced fluid.
F.= pgV, (2.7-9)
One-Dimenslonal Laminar Flow Chap. 2 S.C. 2.7 Buoyancy Forces
SS

Fl1. 2.7-2. Solid body floating at an interface.

The result is not difficult to extend to the case of a solid body at the interface
between two fluids. This situation is illustrated in Fig. 2.72 and the proof
is left to the student.

The hydrometer

k.
on a submerged solid body.
A hydrometer uses the buoyancy principie to determine the ratio of
densities of two fluids. In general, one of these fluids is water and the ratio
of densities, pf Pu 10, is called the specific gravity, y.

r=_P_ (2.7-10)
Pu 1 o

f
+ Plr- 1 dA. (2.7-5)
A hydrometer floating first in water and then in sorne other fluid is shown in
Fig. 2.7-3. lfthe mass ofthe hydrometer is M and the density ofthe air above
the liquid is so small that the air contributes a negligible amount to the
integral sign, and noting that buoyancy force, then Archimedes' principie yields
= pg(z2 - z1) (2.7-6) Mg = Pu 1ogVu 1o (2.7-11)
Mg = PouKVou (2.7-12)
2 - Z) dA. (2.7-7)
where Vu.o and V011 are the volumes of the displaced water and oil respec-
tively. Division of Eq. 2.7-12 by Eq. 2.711 gives
the solid V, is

(2.7-8) 1= (~)(~) (2.7-13)


Pu,o Vu 1 o

"weight" of the displaced fluid. and the specific gravity of the oil is

pgV, (2.7-9) (2.7-14)


Sec. 2.8 One-Dimensional Laminar
56 Fluid Statics and One-Dimensional Laminar Flow Chap. 2

_ _1 __
'o
u
l
Grovity
Fig. 2.8-1. Lamin

imaginary lines traced out by part


Couette viscometer described in Ch
flow illustrated in Fig. 2.8-1, v8 =
a fluid particle would move in a st
(o) (b)
We start the analysis with the r
Fig. 2.7-3. A hydrometer.

If the cross-sectional area of the graduated shaft is A then


!!...
Dt
J pvdV=
'Y,.(t)

V0 n = Vn,o +A Llh (2.7-15) which is to be applied to the ditfe


and 2.8-2. This section of fluid is a mat
1 deforming; however, the velocity
(2.7-16)
Yon = 1 + (__i_) Llh constant. Thus, the momentum is
Vn,o
We need to know both the cross-sectional area A and the volume Vn 2 0 in
f pv dV=
?",.(t)
order to determine You The area A can be meas u red easily and Vn,o may be
determined from Eq. 2.7-ll since the density of water is well known and the
mass ofthe hydrometer is measured easily. Knowing A and Vn 0 allows us to

mark off a scale on the stem of the hydrometer giving the specific gravity
directly.

2.8 One-Dimensional Laminar Flows

If only a single component of the velocity vector is nonzero, the linear


momentum principie is easy to apply, provided the flow is laminar and the
streamlines are straight. The subject of streamlines will be discussed in detail Fig. 2.8-2. Differen
in Chap. 3; for now, it will suffice to say that for steady ftow, streamlines are
Sec. 2.8 One-Dimension:al Lamin:ar Flows 57
One-Dimensional Laminar Flow Ch:ap. 2

~
1o
r .........
1

Lz
~

- _l_
7
/'
~

7
.\
7
7
j_

-
~./

l
Grovity

Fig. 2.8-1. Laminar flow in a circular tube.

imaginary Iines traced out by particles of fluid. Thus, the streamlines in the
Couette viscometer described in Chap. 1 would be concentric circles. For the
flow illustrated in Fig. 2.8-1, v8 = vr =O, and v. is only a function of r; thus
(b}
a fluid particle would move in a straight line parallel to the z-axis.
We start the analysis with the linear momentum equation
hydrometer.

shaft is A then ~t Jpv dV = Jpg dV +Jt<n> dA (2.8-1)


'f/ m(t) "f/ m(t) .r:/ m(t)

(2.7-15) which is to be applied to the differential cylindrical section shown in Fig.


2.8-2. This section offluid is a material volume, and is therefore continuously
1 deforming; however, the velocity of every fluid particle in this volume is a
(2.7-16)
A )
( VH,o
~h constant. Thus, the momentum is also constant

area A and the volume VH 20 in J pv dV = constant vector (2.8-2)


be meas u red easily and VH 2 o may be "f/ m(t)

of water is well known and the


Knowing A and VH 2 0 allows us to
rometer giving the specific gravity

-
Trrl r +11r

PJr +tu
Flows

velocity vector is nonzero, the linear


provided the flow is laminar and the
streamlines will be discussed in detail Fig. 2.8-2. Differential material volume element.
that for steady jlow, streamlines are
58 fluid Statics and One-Dimensional Laminar Flow Chap. 2
Sec. 2.8 One-Dimensional Laminar Flo

Inasmuch as the time rate of change of a constant vector is zero, Eq. 2.8-1
takes the form Equation 2.8-6 is called the stress eq
in terms of the velocity gradient if
O= f
7'",.(t)
pg dV f
+ ~n> dA
Jllf,.(t)
(2.8-3)

This result is identical to that for static ftuids, but in this case the stress vector
has both normal (pressure) and tangential (shear) components. We are only The relationship between the shea
interested in the z-component of Eq. 2.8-3, which may be applied to the flow is very similar to the relations
differential volume to yield flow. In Chap. 5, Newton's law of
the present, we must regard Eq.
Forces on the
z-surfaces relationship between the shear stres
(2.8-4) circular tu~.
Forces on the If the viscosity 1-' is constant, w
r-surfaces
to yield
A very definite sigo convention is used with the shear stress, -rr: On the op
-=p
surface having an outwardly directed normal in the positive r-direction, the oz
shear stress -rr is represented as acting in the positive z-direction. On
the surface having an outwardly directed normal in the negative r-direction, The solution of this equation req
the shear stress is represented as acting in the negative z-direction. The actual the functional dependence of v, a1
direction of the shear stress will depend on the specific problem under fluid particle must move in a straig
investigation. For example, ifthe flow in the tube were reversed, the direction this flow is constant, the volumetri
of the shear stress could certainly be reversed. In formulating problems, must remain constant (which follo
we must adhere to the sigo convention that a positive shear stress acts in the implies conservation of volume,
positive coordinate direction on positive surfaces,t and in the negative conclude that the velocity v, is not
coordinate direction on negative surfaces. This scheme is necessary to that the ftow is axisymmetric,t v, i1
insure that shear stresses on opposite sides of a differentially thin shell of
fluid act in opposite directions, which is, of course, intuitively obvious. A V

proof will be given in Chap. 4 where stress is studied in detail.


The pressure will be a function of
Dividing Eq. 2.84 by 21rt1r!1z and taking the limits !1r-+ O and !1z-+ O,
it will certainly depend on z. Thu
we get
0 = lim [-r(PI:H - PI)] p=
.,. .... o /::,.z
(2.8-5) However, the effect of gravity wil
and the pressure gradient will
dependence of opfoz may be expr
which immediately leads to the differential equation op
op + -1 -a (r-rr,) oz
O= - - (2.8-6)
oz r or Examining Eq. 2.8-8 in light of t
side is only a function of z while
t By positive surfaces we mean the surfaces having normal vectors pointing in the
positive coordinate directions.
K.ceping in mind that r and z ar
t See Birkhoff's plausible intuitive h
One-Dimensional Laminar Flow Chap. l
Sec. 2.8 One-Dimensional Laminar Flows 59
of a constant vector is zero, Eq. 2.8-1
Equation 2.8-6 is called the stress equation of motion, and it may be expressed
in terms of the velocity gradient if Newton's law of viscosity is used.
(2.8-3)
(2.8-7)
fluids, but in this case the stress vector
(shear) components. We are only The relationship between the shear stress and the velocity gradient for this
. 2.8-3, which may be applied to the flow is very similar to the relationship given in Chap. 1 for plane Couette
flow. In Chap. 5, Newton's law of viscosity will be examined in detail; for
the present, we must regard Eq. 2.8-7 as an experimentally determined
Forces on the
z-surfaces relationship between the shear stress and the velocity gradient for flow in a
(2.8-4) circular tu~.
Forces on the lf the viscosity /-' is constant, we may substitute Eq. 2.8-7 into Eq. 2.8-6
r-surfaces
to yield
with the shear stress, T rz On the
normal in the positive r-direction, the (2.8-8)
in the positive z-direction. On
normal in the negative r-direction, The solution of this equation requires sorne careful arguments regarding
in the negative z-direction. The actual the functional dependence of v, and p. Because v11 and vr are zero,t each
on the specific problem under fluid particle must move in a straight line along the tube. If the density for
in the tube were reversed, the direction this flow is constant, the volumetric flow rate across any section of the tube
reversed. In formulating problems, must remain constant (which follows from the fact that conservation of mass
that a positive shear stress acts in the implies conservation of volume, if the density is constant), and we can
surfaces,t and in the negative conclude that the velocity v. is nota function of z. If, in addition, we assume
This scheme is necessary to that the flow is axisymmetric,t v. is only a function of r.
sides of a differentially thin shell of
is, of course, intuitively obvious. A v. = v.(r) (2.8-9)
is studied in detail.
The pressure will be a function of r and O owing to gravitational effects, and
taking the limits Ar ~ O and Az ~O,
it will certainly depend on z. Thus,

p = p(r, O, z) (2.8-10)

(2.8-5) However, the effect of gravity will be the same everywhere along the tube,
and the pressure gradient will not depend on r and O. The functional
dependence of opfoz may be expressed as
equation
op =f(z) (2.8-11)
oz
(2.8-6)
Examining Eq. 2.8-8 in light of these arguments, we see that the left-hand
side is only a function of z while the right-hand side is only a function of r.
having normal vectors pointing in the
Keeping in mind that r and z are independent variables, we conclude that
t See Birkhoff's plausible intuitive hypothesis IV, Sec. 1.3.
60 Fluid Sutics and One-Dimensional Laminar Flow Chap. 2
Sec. 2.8 One-Dimensional Laminar

both sides of Eq. 2.8-8 must be constant,t and we write Eq. 2.8-18 is a very useful result,
!-' ! !!:._
r dr
(r dv,)
dr
= _ !lp
L
(2.8-12)
if the dimensions of the system, the g
The volumetric flow rate is given b

Here the partial derivatives have been replaced by total derivatives, inasmuch
as v. is only a function of r, and the pressure gradient opfoz has been replaced
by -(!lp/L), where
fr:.,
=
o o

(2.8-13) This relationship between the volu


pressure gradient !lpfL is called th
Multiplication ofEq. 2.8-12 by r dr, division by ,t, and integration give hydraulician, Hagen, and the ph

dv,
r dr = -
(!lp) r
L 2,t + e
2

1 (2.8-14)
established this result in the early p
We see by Eq. 2.8-19 that the flo
pipe radius; thus, a small error in
Dividing by r and integrating again, we have an expression for the velocity error in the flow rate or pressure
profile. fluid, specification of the pressure d

(!lp)
must be made on an economic bas
v, = - -
,2
- + e In r + e 1 2 (2.8-15) cost of the pipe is small but the cos
L 4,t constantly be confronted with prob
The boundary conditions for this flow are economic constraint be placed on tll
Other quantities of interest can
B.C. 1: v, is finite for O :::;; r :::;; r0 (2.8-16a) velocity, Vz,mu occurs at r =O and
B.C. 2: v. =o, r ='o (2.8-16b)
v.,max
Both these conditions are derived from the idea that the fluid is a continuum;
the velocity is therefore a continuous function. The first boundary condition The same conduits are often used to
is somewhat unusual in that the velocity is not specified at sorne point; it is costs be minimized and piping layout
simply restricted to a finite value. Boundary conditions of this type are often to know how long it will take before
found in the analysis of fluid motion, heat transfer, and mass transfer. conduit arrives at the exit. The ma
Because the statement is so obvious, it is sometimes overlooked by students the "breakthrough" time. The aver
unfamiliar with the solution of boundary value problems. Application of metric flow rate divided by the croSl
boundary condition 1 requires that e1 be zero, beca use In r - - oo as r - O,
and boundary condition 2 yields an expression for e2, (v.) = _Q__
7Tr~
vi-
z r-ro
= 0 =(
!lp) ,
-L- 4,t
2
~+e2 (2.8-17) The average velocity is often a use
since the extent of a chemical reacf
The velocity may now be given as in the reactor. This time is called tll

(2.8-18) average resid

Application of the Hagen-Poiseuill


indicating that the velocity profile is parabolic. therefore, we must be sure that the
t The arguments presented here are mainly qualitative and should be accepted in that t Throughout the text, angular brack
context. A quantitative treatment of this problem is given in Chap. 5.
area averages.
One-Dimensional Laminar Flow Chap. 2
Sec. 2.8 One-Dimensional Laminar Flows 61
t and we write
Eq. 2.8-18 is a very useful result, for it allows us to calculate the flow rate
if the dimensions of the system, the pressure drop, and the viscosity are given.
(2.8-12) The volumetric flow rate is given by

replaced by total derivatives, inasmuch


gradient opfoz has been replaced Q= fo 2I"o Vzr dr dO = -7TT~(/:l.p)
ro
-
8.t L
(2.8-19)

(2.8-13) This relationship between the volumetric flow rate, the radius r 0 and the
pressure gradient LlpfL is called the Hagen-Poiseuille law in honor of the
by .t, and integration give hydraulician, Hagen, and the physician, Poiseuille, who experimentally
established this result in the early part of the nineteenth century.
(2.8-14) We see by Eq. 2.8-19 that the flow rate is very sensitive to changes in the
pipe radius; thus, a small error in sizing a pipe can lead to an_appreciable
we have an expression for the velocity error in the flow rate or pressure drop. For a given flow rate and a given
fluid, specification of the pressure drop (i.e., the pump size) and the pipe size
must be made on an economic basis. If a small-diameter pipe is used, the
(2.8-15) cost of the pipe is small but the cost of the pump is large. An engineer will
constantly be confronted with problems of this type which require that an
economic constraint be placed on the derived result.
Other quantities ofinterest can be derived from Eq. 2.8-18. The ma:>fJmum
(2.8-16a) velocity, Vz,max occurs at r =O and has the value
(2.8-16b)
r~1-' L
Vs,max =
(!l.p) (2.8-20)
the idea that the fluid is a continuum; 4
!lutlcuon. The first boundary condition The same conduits are often used to carry different fluids in order that piping
is not specified at sorne point; it is costs be minimized and piping layouts simplified. In such cases it is important
conditions of this type are often to know how long it will take before the new fluid being pumped through the
heat transfer, and mass transfer. conduit arrives at the exit. The maximum velocity can be used to calculate
is sometimes overlooked by students the "breakthrough" time. The average velocity (vz)t is defined as the volu-
value problems. Application of metric flow rate divided by the cross-sectional area; therefore,
be zero, because In r ---. - oo as r ---. O,
ex~:;reS:SIOID for C2, (vz) = _Q_ = ( '~) (Llp)
7Tr~ L8.t
(2.8-17) The average velocity is often a useful quantity for sizing pipeline reactors,
since the extent of a chemical reaction depends on the time the fluid spends
in the reactor. This time is called the average residence time and is given by

average res1.dence hme


. = -
L
(2.8-18)
(v:)
Application of the Hagen-Poiseuille law is limited to steady, laminar flow;
therefore, we must be sure that the flow under consideration is not turbulent.
qualitative and should be accepted in that
is given in Chap. 5. t Throughout the text, angular brackets (()) will be used to denote both volume and
area averages.
62 Fluid Statics and One-Dimensional Laminar Flow Chap. 2
lec. 2.1 O..Oimenslonal Laminar

Calculation of the fl.ow rate should always be accompanied by a determination


If we wish the answer in terms
of the Reynolds number,
_ p(v.)D
N Re-
t
/:ip = { 1.91 X
1
10 dynefcm
Orfllnal tcrm
1
}{ e
to be sure that it is Iess than the critica! value, 2100.
Knowledge of the stress distribution for laminar fl.ow in a tube is sometimes To complete the solution, we
desired. Rewriting Eq. 2.8-6 in the form R.eynolds number is given by

~r (r-rr.) = -e~:) r (2.8-21)


and integrating, we have
T
u
= - (tlp)!:
L 2
+ C
r
(2.8-22)

Because the shear stress must always be finite, we may write,


B.C. 1: Tr is finite for O :::::;: r :::::;: r 0
Thus, the constant of integration C1 is zero, and the shear stress is a linear
function of r.
Carrying out the computation,
T
u
= -(/j.p)!:
L 2
(2.8-23)
1
Thus, the fl.ow is laminar.
The fact that T r is negative i.s in keeping with our intuition, because the shear
stress acting on the r-surface having a positive outwardly directed normal
must certainly act in the negative z-direction if the fluid is flowing in the posi- Non-Newtonlan flow betwee1
tive z-direction.
Figure 2.8-3 illustrates the fl.
Sample calculation of pressure drop parallel plates, jnclined at an
profile is indicative of a pseu
Oil having a density of 0.87 gfcm 3 and a viscosity of 1.3 poises is to be flow to be one-dimensional an
pumped through a smooth pipe lOO ft long with an inner diameter of 2.0 in. Eq. 2.8-3 to the differential vol
We wish to calculate the pressure drop for a volumetric flow rate of 1000
O= (pgr~:) !:u !:ly !:lz
ft3/hr. Rearranging Eq. 2.8-19 gives
tlp = 8p.LQ/7Tr~ (2.8-24) + {p !:ly tlzlr~: -
Substituting the appropriate values and the necessary conversion factors, we
get
!:l = {(8)(1.3 poises)(lOOft)(1000fe/hr)}
P (3.14)(1.0 in.)4 Dividing by Ax !:iy /:iz and taki
Original tcrms

dyne-~ecfcm ) (12 in.) (


2 4
X { (1 1 hr ) } O= pgs
p01se ft 3600 sec
Conversion factors

tlp = 1.91 X 106 dynefcm 2 +IiAw-+


d One-Dimensional Laminar Flow Chap. 2
S.. 2.1 On....Oimenslonal Laminar Flowa 63
be accompanied by a determination
If we wish the answer in tenns of pounds-force per square inch, we write

dp = {t.91 X 8
10 dynefcm
Ori,UW term
2
}{ c 25
:y!~-'lbt) c-~~J} = 27.6lbr/in.
Con.ualon facton
1

value, 2100.
for laminar flow in a tu beis sometimes To complete the solution, we need to verify that the flow is laminar. The
R.eynolds num.ber is given by
_ p(v.)D _ 4pQ
(2.8-21) N Re- -
1-' '"D.
3
N _ {(4X0.87 gfcmSXlOOO ft fhr)}
(2.8-22)
Re - (3.14X2.0 in.X1.3 poises)
Ori,UW terma
be finite, we may write,
1 hr ) (12 in.)
8

(2-~4 cm) (
2
O::;; r::;; r 0 X {( 1 poise ) }
zero, and the shear stress is a linear 3600 sec ft tn. g/cm-sec
Con.uaiOil facton

Carrying out the computation, we have


(2.8-23)
NRe = 1320
with our intuition, beca use the shear Thus, the flow is laminar.
a positive outwardly directed normal
if the fluid is flowing in the posi- Non-Newtonlan flow between two flat plates

Figure 2.8-3 illustrates the flow of a non-Newtonian fluid between two


parallel plates, ~nclined at an angle () from the horizontal. The velocity
profile is indicative of a pseudoplastic "power-law" fluid. Assuming the
and a viscosity of 1.3 poises is to be flow to be one-dimensional and laminar, we apply the x-component of
long with an inner diameter of 2.0 in. Eq. 2.8-3 to the differential volum.e shown in Fig. 2.8-3 to obtain
for a volumetric flow rate of 1000
O = (pg#e) dx dy dz Body force
(2.8-24) Forces on the
+ {p dy dzlz - p dy dzl~.u}
x-suaces (2.8-25)
Forces on the
+ {-Twzdxdzlw + 1'wzdxdzlw+Aw} y-suaces

Dividing by dx ay dz and taking the limit as before, we get

() - lim PI~z -
O = pgsm Plz
.U-+0 dx
(2.8-26)
64 Fluid Statics and One-Dimensional Laminar Flow Chap. 2 S.C. 2.8 OnDimenslonal Laminar Flc

wc note that the left-hand side is in


-r,. is independent of x, thus botl
constant. Integration is then straig

T
vz
=- (fl.p,
L
The boundary condition that we
integration is that the shear stress
B.C. 1:
The easiest way to demonstrate th
2.8-3 to the differential volume sl

y=O
~

Fig. 2.8-3. Flow between two flat plates.

where g., has been replaced by g sin e. Each limit represents a deriva ti ve, and
the limiting process gives us

o= pg sin e - -op
ox
+ -oT'J/X
oy
(2.8-27)

We may also apply the y-component of Eq. 2.8-3 to this flow to obtain

o= - pg cos e - .)p
-oy (2.8-28)
Flg. 2.11-4. Shear s1

stresses at y = +fl.y and y = - j


lntegration of Eq. 2.8-28 gives an expression for the pressuret
following the assumption of symrr
p(x, y) = - pgy cos e + C(x) (2.8-29)
which leads to the conclusion that opfox is independent of y. Rearranging Evaluation of the terms in Eq. 2.8
Eq. 2.8-27 into the form,
O = (pg 111)2 fl.x fl.;
op
--
. e a'TIIX
pgsm = - (2.8-30)
ox oy + 2p !l.y fl.zl.
t Note that integration of a partial derivative yields a "constant of integration," which
may be a function of the other independent variables.
One-Dimensional Laminar Flow Chap. 2 S.C.. 2.8 OnDimenslonal laminar Flows 65

we note that the left-hand side is independent of y. For one-dimensional ftow


,...., is independent of x, thus both sides of Eq. 2.8-30 must be equal to a
constant. Integration is then straightforward, and we obtain

(2.8-31)

The boundary condition that we wish to apply to evaluate the constant of


integration is that the shear stress is zero at the plane of symmetry, y= O.
B.C. 1: 1"wz = O, Y= O (2.8-32)
The easiest way to demonstrate the validity of this condition is to apply Eq.
2.8-3 to the differential volume shown in Fig. 2.8-4. Note that the shear

two flat plates.

Each limit represents a derivative, and

Jp JTyz
--+-
Jx Jy
(2.8-27)

of Eq. 2.8-3 to this flow to obtain

Fl. 1.8-4. Shear stress at a plane of sym.metry.


(2.8-28)

stresses at y= +~y and y= -~y are both actingin the same direction
following the assumption of symmetry about the plane y = O, i.e.,
(2.8-29) (2.8-33)
Jx is independent of y. Rearranging Evaluation of the terms in Eq. 2.8-3 gives
O = (pgz)2 ~X ~y ~z Body force
(2.8-30) Forces on the
+ 2p ~y ~zlz - 2p ~y ~zlz+4z x-suaces (2.8-34)
Forces on the
y-surfaces
66 Sec. 2.8 One-Dimensional Laminar Flows
Fluid Statics and One-Dimensional Laminar Flow Chap. 2

Ifwe Jet dx and dz remain finite and allow dy to go to zero, Eq. 2.8-34 reduces The volumetric flow rate per unit wid
to Eq. 2.8-43.
2rul.=o dx dz = O (2.8-35) dp -H
and it follows that the shear stress at the plane of symmetry must be zero. q= (~) ( 1:_
Thus, C1 is zero and the stress distribution is 2n + 1
The velocity profile is somewhat
Tu= -(d: + pgsin o)y (2.8-36) dimensionless by dividing by the avell

The shear stress for a power-Iaw fluid in one-dimensional flow is given by


v.,
U.,= (v.,)
(2n
= -;;-
(see Eq. 1.5-9)
where Y= 2yfh, a dimensionless di
(2.8-37) puted from Eq. 2.8-45, are shown in

Since the flow is symmetric about y = O, we need only treat the region
O L. y L. h/ 2. Noting that the velocity gradient will always be negative in this
region, we can write

~;"' = -1 ~;"' , (2.8-38)

Combination of Eqs. 2.8-36, 2.8-37, and 2.8-38 yields

Tu= -m 1~;"'In= -(d: + pgsin o)y (2.8-39)

Dividing by -m and raising both sides to the 1/n power,

dp + pg sin 0) 1
/n
dv., = ( L ylfn
(2.8-40)
1dy 1 m
Using Eq. 2.8-38 yields

dp + pg sin 0) 1
/n

e;.,) = - (L m i'" (2.8-41)

This equation may be integrated, subject to the boundary condition

B.C.l: h
y=- (2.8-42)
2 Fl1. 2.8-5. Velocity profiles for ~
to obtain the velocity distribution flat plates.
t The total volumetric flow rate (in cub
Q, while the volumetric flow rate per unit 1
represented by q.
(2.8-43) ~ Whenever possible, capital letters w
and One-Dimensional Laminar Flow Chap. 2 Sec. 2.8 One-Dimensional Laminar Flows 67

allow dy to go to zero, Eq. 2.8-34 reduces The volumetric flow rate per unit width qt is readily obtained by integrating
Eq. 2.8-43.
dz =O (2.8-35)
at the plane of symmetry must be zero. (2.8-44)
"bution is
The velocity profile is somewhat easier to examine if the velocity is made
(2.8-36) dimensionless by dividing by the average velocity.!

in one-dimensional flow is given by U :e=~= (2n + 1)(1- y<n+llfn) (2.8-45)


(v:e) n+1
where Y= 2yfh, a dimensionless distance. Severa! velocity profiles, com-
(2.8-37) puted from Eq. 2.8-45, are shown in Fig. 2.8-5.

Y = O, we need only treat the region


gradient will always be negative in this

. (2.8-38)
Pseudoplastic
2.8-38 yields

-( + pg sin o)y (2.8-39) 1


Ux=l.5
to the 1/n power,

pg sin
i---
0) 1111

ylfn (2.8-40)
m

pg sin 0) 1
/n
1-----
m
l'" (2.8-41) Dilatan!

to the boundary condition


1
h 1
Ux=l .5
y=- (2.8-42)
2
Fl1. 2.8-5. Velocity profiles for a power-Iaw fluid ftowing between two
ftat plates.
t The total volurnetric ftow rate (in cubic feet per second) will always be represented by
h)(l+n)/n[ (2y)(l+n)/n] Q, while the volurnetric ftow rate per unit width (in cubic feet per second per foot) will be
rcpresented by q.
(2 1 - h (2.8-43) t Whenever possible, capital letters will be used to denote dimensionless quantitics.
Problema
68 Fluid Statics and One-Dimensional Laminar Flow Chap. 2

~Atmospheric p
If we compare Eq. 2.8-37 with Newton's law of viscosity, we see that the
apparent viscosity is given by

f-lapp =m
1
d:., n-1
d
(2.8-46)

Thus, if n < 1 the apparent viscosity decreases with increasing shear rate,
and the fluid is called pseudoplastic. Since the shear rate is zero at the
centerline and a maximum at the wall, the apparent viscosity becomes
increasingly smaller near the wall, and the flow resembles a case where there
might be "slip" at the wall. Note in Eq. 2.8-44, that as n becomes small, the
volumetric flow rate becomes more sensitive to the channel depth, h. In
working with highly pseudoplastic fluids, we must be extremely careful when
specifying pipe diameters or channel depths, for a small error in geometry can
lead to a large error in the flow rate at a given pressure drop.
lf n > 1, the apparent viscosity increases with increasing shear rate; thus,
the fluid appears to be more viscous near the walls, and a larger proportion
of the flow takes place in the central region of the channel. Fluids which Fla.
behave in this manner are called dilatant, and are not nearly as common as
pseudoplastic fluids . Atmospherlc
pressure
If n = 1, the fluid is Newtonian and we replace m with p,. The velocity
profile is parabolic, and the volumetric flow rate per unit width is given by

q = -h3 (!!:.p
- + pg sin () ) (2.8-47)
12t-t L
This type of flow is sometimes referred to as "plane" Poiseuille flow.
The methods presented in this section have allowed us to solve sorne
rather important practica! problems without recourse to detailed analysis;
however, this approach is only useful when the streamlines are straight. Much
work remains to be done before we can handle more complex flows.

PROBLEMS

o
2.1 . If the density of fluid 1 is 62.4lb m/ft 3 and the density of fluid 2 is 136.8 lb m/ft,
determine the gas pressure in the tank shown in Fig. 2-1. Assume that the Fla. 2-2
density of the gas in the tank is negligible compared to the two manometer
fluids. 2-3. For ftuids with a density close to,
gravity is best determined in the
Ans.: 20.4 psia. pression for y in terms of z1, z1, z
2-2. A simple U-tube manometer can be used to determine the specific gravity y Pt (z4 - zs) - (za -
of fluids which are more dense than water by the arrangement illustrated in A.ns.: y = PI = (z - zs)
Fig. 2-2. Derive an expression for y in terms of z1, z2, and z3
and One-Dimensional Laminar Flow Chap. 2 Problema 69

's law of viscosity, we see that the ~Atmospheric pressure

(2.8-46)

decreases with increasing shear rate,


Since the shear rate is zero at the
wall, the apparent viscosity becomes
the flow resembles a case where there
Eq. 2.8-44, that as n becomes small, the
sensitive to the channel depth, h. In
we must be extremely careful when
for a smal! error in geometry can
at a given pressure drop.
tcn~as,es with increasing shear rate; thus,
near the walls, and a larger proportion
region of the channel. Fluids which 2
and are not nearly as common as Fl. 2-1

we replace m with p. The velocity


flow rate per unit width is given by Atmospherlc
/pressure~

+ pgsin o) (2.8-47)
Less dense
to as "plane" Poiseuille flow. fluid
have allowed us to solve sorne
recourse to detailed analysis;
the streamlines are straight. Much
handle more complex flows. Woter

LEMS

and the density offluid 2 is l36.8Ibm/ft,


shown in Fig. 2-l. Assume that the Fl.l-l
compared to the two manometer
2-3. For fluids with a density close to, but less than, _that_ of water, t~e specific
gravity is best determined in the system shown m Flg. 2-3. Denve an ex-
pression for y in terms of Z, z1, z8 and z,.
used to determine the specific gravity y
water by the arrangement illustrated in P1 (z4 - za) - (z8 - zJ
in terms of z1, z2, and z3 Ans.: Y = PI = (zl - za)
Problema
70 Fluid Statics and One-Dimensional Laminar Flow Chap. 2

2-4. If the air can be treated as an ideal gas, and the temperature of the atmos-
phere vares linearly with height above the earth,
T = T0 - a.z
derive an expression for the pressure as a function of z neglecting fluid motion
in the atmosphere and the rotation of the earth.
" 2-5. Two fluids are confined by a hinged gate as shown in Fig. 2-5. If the lower
fluid is water and the upper fluid has a specific gravity of 0.8, determine the
moment per unit width about point "A."
Ans.: Moment per unit width is 69.4 x 104 lbr-ft/ft.

Oil y=0.8

L.
Fla.l-s

2-6. A wooden sphere (p = 58 lbm/ft3) is floating atan air-water interface. What


fraction of the sphere is submerged? Solve by first deriving Archimedes'
principie for an arbitrary body located at a fluid-fluid interface.
2-7. A wooden plank of density 46lbm/ft3 is anchored in a submerged surface as
illustrated in Fig. 2-7. If the plank is 1 in. thick and 10 in. wide, what is thc
moment about point "A"?
Ans.: Moment is 571br-ft.
2-8. Determine both the horizontal and vertical components of the force per unit
width exerted by the fluid on the curved gate shown in Fig. 2-8. Use Eq.
2.5-2 to determine these forces, and deduce from the result that the y-com-
ponent is simply the weight of the fluid above the gate, while the x-component
is the average pressure exerted on the gate times the projected area. In ordcr 2-10. The spherical tank shown in Fig. 2-
that the units remain consistent, take {J = 1 ft 112 indicated by the mercury manometer
is 2000 lbm. If30 bolts are used to bol
2-9. Work Prob. 2-8 for the case where the density is a linear function of y.
perbolt?
P = Po - 11p (L)
6ft
2-11. A weather balloon 1Oft in diamete
rcsult obtained in Prob. 2-4 with o
where Ap =10 lbm/ft3
the balloon will risc. Assume tha
Po =62.4lbm/ft3 is constant at 10ft, and the mass
y= ft Given: Ralr = 0.0253 atm ft3/lbm
Ans.:f., = 896lbr/ft; [ 11 = -2960 lbr/ft.
and One-Dimensional Laminar Flow Problema 71
Chap. 2

gas, and the temperature of the atmos- y

L.
the earth,

as a function of z neglecting fluid motion


of the earth.
gate as shown in Fig. 2-5. If the lower
has a specific gravity of 0.8, determine the
"A."
x 104 lbr-ft/ft.

=0.8

L. Fl.l-7

is floating at an air-water interface. What


? Solve by first deriving Archimedes'
at a fluid-fluid interface.
L.
is anchored in a submerged surface as
is 1 in. thick and 10 in. wide, what is the

vertical components of the force per unit


gate shown in Fig. 2-8. Use Eq. Fl.l-1
deduce from the result that the y-com-
above the gate, while the x-component
gate times the projected area. In order 2-10. The spherical tank shown in Fig. 2-10 contains both water and oil ata pressure
p= 1 ftl/2,
indicated by the mercury manometer. The mass of each half of the spherical shell
the density is a linear function of y. is 2000 lbm. If 30 bolts are used to hold the two sections together, what is the force
perbolt?
- llp (L) 6ft 2-11. Aweatherballoon lOftindiameterisfilled with 7.0 lbm ofhelium. Using the
result obtained in Prob. 2-4 with a: = 0.005K/ft, determine to what altitude
the balloon will rise. Assume that T0 = 298K, the diameter of the balloon
is constant at 1O ft, and the mass of the balloon is 1O lb m:.
Given: Ralr = 0.0253 atm ft 3/1bm 0 R, RHe = 0.182 atm ftll/lbm 0 R.
Ibr/ft.
72 Fluid Statics and One-Dimensional Laminar Flow Chap. 2
Problems

Fig. 2-10
Fluid 2
2-12. Determine the thickness t of the concrete dam shown in Fig. 2-12, which is Fig. j
required to produce a zero moment about point B. Assume that the hydro-
static head on the bottom of the dam varies linear! y from 80 ft atA to zero
at B. 2-14. The center of stress (often referred t
Ans.: t = 34.1 ft. systems) is defined as the positio?
still give the same torque as the d1st1

1'1 position vector locating the center ol


above definition is
, .. . ,.
.
.. ~..:; ~~:
. ..
..
: . i' X F =
Water ;.:: :. . ' .. 9
.. Coherete :.
80ft ~ : y=25 .'
. . ... where F'
.. 6
. . . . 4 ..
. ... . .

Use these equations to loca te the ce1


/ when the density is constant.
Fig. 2-12 2-15. Prove the following:

(a) A (B x C) = B (C X A) =
2-13. The micromanometer illustrated in Fig. 2-13 is a useful device for accurately
measuring small pressure differences. If the densities of the two manometer (b) A X (B X C) = -(B X C) X
fluids are nearly the same ( p1 ""' p2), measurable val ues of the distance d can be
Given that
obtained for very small values of the pressure difference, PB - PA Letting
A 1 be the cross-sectiona1 area of the reservoirs, and A 2 be the cross-sectional ixj=k=-
area of the connecting tu be, derive an expression for PB - pA in terms of jxk=i=-
P1 p2 , g, A 1 , A 2 , and d.
kxi=i=-
and One-Dimensional Laminar Flow Chap. 2
Problems 73

Fluid 2
concrete dam shown in Fig. 2-12, which is
about point B. Assume that the hydro- Fig. 2-13
dam varies linearly from 80ft atA to zero

2-14. The center of stress (often referred toas the center of pressure for hydrostatic
systems) is defined as the position at which total force could be applied and
still give the same torque as the distributed force system. Ifwe define i' as the
position vector locating the center of stress, the mathematical equivalen! of the
. , . d ... f above definition is

f
. A .

: ~ .~;;, :_: ~::_


i' X F = r X ten> dA
.-.: :.. ... 9
Concrete A
~ r = 2.s -~.
~ where
;_:
..
.
.
J
. ,4.
F = Jtcn>dA
<111
A

Use these equations to loca te the cen!er of stress for the gate shown in Fig. 2-8
/
when the density is constant.
2-15. Prove the following :

Fig. 2-13 is a useful device for accurately (a) A (B x C) = B (C x A) = C (A x B)


If the densities of the two manometer (b) A X (B X C) = -(B x C) x A = B(A C) - C(A B)
measurable val ues of the distan ce d can be
pressure ditference, PB - PA Letting Given that
reservoirs, and A 2 be the cross-sectional i X j = k = -j X i, i X i = 0
an expression for PB - pA in terms of
j X k = i = -k X j, jxj=O
k X i = j = -i X k, kxk=O
74 Fluid Statlca and One-Dimenalonal Laminar Flow Chap. l

2-16. Demonstrate that the cross product, A x B, can be written in term~ of the
determinant
j k
A X8 = A.., A" A.
B.., B" B,
2-17. Show that A (B x A) is zero.
Kinematics
2-18. Show that the cross product rules may be expressed in index notation as
e< 1> x e11 , ~ O, i = j
e< 1> x e(i, = e(kl i "- j ,. k ,. i
2-19. Prove that
V X (Vp) = 0
thus, the curl of the gradient of a scalar is zero. To do this, expand the term
V x (Vp) and regroup the scalar components.
2-20. Determine the velocity profile for laminar flow of a Newtonian fluid in the
annular region illustrated in Fig. 2-20.

In the previous chapter, we anal


pressure field and surfa.ce forces .ror st2
profiles for one-dimens10nallammar fi
tum principie was applied to a diffe1
appropriate differential equations. Th
more general differential equations e
possible only because the acceleration 1
was zero. To analyze the complex fi01
we must be capable of describing tt
2-21. Derive the flow-rate, pressure-drop relationship for the flow of a power-law
Kinematics is the description of moti
fluid through a circular tu be. The relationship between shear stress and shear
how this motion is brought about. 1
rate is given by

Tr
= 1dv, ,..-
m dr
1

dr
(dv,) represents a key step toward our obje
tions of motion. In addition, sorne of
formulation of the macroscopic balan
for one-dimensional laminar flow in a tube. The student is encouraged not to e
as "mathematical gymnastics" unrela
final result, for the derived differential
balances (Chap. 7) cannot be appliec
fidence unless the derivation is unde
understood unless the mathematical
mastered.
One-Dimenslonal Laminar Flow Chap. 2

A x B, can be written in term~ of tbe


k

Kinematics 3
be expressed in index notation as
j =j
i#j#k-:Fi

is zero. To do this, expand the term

In the previous chapter, we analyzed the problem of determining the


pressure field and surface forces for static fluids, and we determined velocity
profiles for one-dimensionallaminar flows. In both cases, the linear momen-
tum principie was applied to a differential volume element to derive the
appropriate differential equations. These were, in fact, special forros of the
more general differential equations of motion, and the derivations were
possible only because the acceleration term in the linear momentum equation
was zero. To analyze the complex flows commonly encountered in practice,
we must be capable of describing the fluid motion in a precise manner.
~"""u'''""'~' for the flow of a power-Jaw
, .... vuo'"'~' between shear stress and sbear Kinematics is the description of motion per se, and it takes no account of
how this motion is brought about. The material presented in this chapter
represents a key step toward our objective of deriving the differential equa-
tions of motion. In addition, sorne of the derived results must be used in the
formulation of the macroscopic balances treated in Chap. 7.
The student is encouraged not to dismiss these preliminary developments
as "mathematical gymnastics" unrelated to the practical application of the
final result, for the derived differential equations (Chap. 5) or the macroscopic
balances (Chap. 7) cannot be applied to practical problems with any con-
fidence unless the derivation is understood. And the derivation cannot be
understood unless the mathematical tools used in the development are
mastered.
75
Sec. 3.2 Time Oerlvatlves
76 Kinematics Chap. 3

material coordinates do not represent a


deforms with the fluid; they simply
*3.1 Material and Spatial Coordinates i coordinates-namely, those which the fl
-time, t = O. Since no two fluid particles
Our objective in this section is to develop an understanding of material at the same time, the material posit!
coordinares and their relationship to spatial coordinates, so that we can particle. Equation 3.1-2 may be writte '
adequately describe the motion of a fluid. The term "spatial coordinates"
refers to a fixed rectangular coordinate system in which all points may be r = r(l
located. The existence of such a coordinate system follows from the assump-
tion that the space is Euclidean. There are two possible methods of locating which simply expresses the fact that the
or identifying a "particle" of fluid, which may be defined as a differential are a function of time and the materia
material volume element, d"Ym(t), i.e. it is a vanishingly small volume of The time rate of change of the spatil
fluid which contains the same material at all times. At sorne time, t, we may particle is the velocity of that particle.
designate the position of a fluid particle in terms of its spatial coordinates-x, with the material coordinates held con
y, and z. In this manner, all fluid particles may be located in terms of the In keeping with the nomenclature to 1:)
spatial coordinates and time. At sorne reference time, chosen as t = O for
convenience, the position of any fluid particle can be specified as

x=X, y= Y, z=Z, at t =O
At sorne other time, t > O, the position of this particle may be expressed *3.2 Time Derivatives
as
t

+ f (~:) dt
The time derivative of an arbitral!
X = X (3.1-1a)
given by
o
dS = lim [S(t
dt AC-oO

(3.1-lb)
where the spatial dependence of S is \l
S is a function of time only, this defin
However, if Sisa function ofthe spati
(3.1-1c) by Eq. 3.2-1 is not well defined until s'
in space at which S is measured for ti
As our first example, we shall e
We may put this result in more compact vector form by multiplying these through space, letting x, represent tl
three equations by i, j, and k, respectively, and then adding to obtain illustrated in Fig. 3.2-1. According to
t

r =R + J(~:)
o
dt (3.1-2)
dx, = lim [x,(
dt At-oO

where r = ix + jy + kz = Vz, theve


x-directi<
R = iX +iY + kZ

We shall call r the spatia/ position vector, because it locates the fluid particle
t Describing the fluid motion in terms e
referred to as the Lagrangian method, while tb
inspace, while R will be called the material position vector, because it represents to as tbe Eulerian method.
the coordinates used to "tag" or identify a given particle. Notice that the
Kinematics Chap. 3 Sec. 3.2 Time Derlvatlves n
material coordinates do not represent a coordinate system which moves and
deforms with the fluid; they simply represent a specific set of spatial
/coordinates-namely, those which the fluid particle occupied at the reference
develop an understanding of material time, t =O. Since no two fluid particles can occupy the same spatial position
at the same time, the material position vector uniquely defines a fluid
spatia/ coordinates, so that we can
fluid. The term "spatial coordinates" particle. Equation 3.1-2 may be written as
system in which all points may be r = r(R, t) (3.1-3)
system follows from the assump-
are two possible methods of locating which simply expresses the fact that the spatial coordinates of a fluid particle
may be defined as a differential are a function of time and the material coordinates.t
it is a vanishingly small volume of The time rate of change ofthe spatial position vector for a particular fluid
at all times. At sorne time, t, we may particle is the velocity of that particle. Since this time derivative is evaluated
in terms ofits spatial coordinates-x, with the material coordinates held constant, it is called a material derivative.
may be located in terms of the In keeping with the nomenclature to be established in Sec. 3.2, we write
reference time, chosen as t = O for

(~:)a=~:=
particle can be specified as (3.1-4)
T
at t =O
of this particle may be expressed
3.2 Time Derivatives

(3.1-1a) The time derivative of an arbitrary scalar function, S= S(x, y, z, t), is


given by
dS = 1im [S(t + dt)- S(t)J (3.2-1)
dt At-oO dt
(3.1-1b)
where the spatial dependence of S is understood. For the special case where
Sisa function of time only, this definition of the derivative is unambiguous.
However, if Sisa function ofthe spatial coordinates, the derivative expressed
(3.1-1c) by Eq. 3.2-1 is not well defined until sorne statement is made about the point
in space at which S is measured for the two times, t and t dt. +
vector form by multiplying these As our first example, we shall consider a system of particles moving
, and then adding to obtain through space, letting x 11 represent the x coordinate of the pth particle, as
illustrated in Fig. 3.2-1. According to Eq. 3.2-1, we write

(3.1-2) dx 11 = 1im [X (t + dt) -


11 X 11(t)J
(3.2-2)
dt At-oO dt
= v.,, the velocity of the pth particle in the
x-direction

because it locates the fluid particle t Describing the ftuid motion in terms of the material coordinates and time is often
position vector, beca use it represents referred to as the Lagrangzn method, while the use of spatial coordinates and time is referred
fy a given particle. Notice that the to as the Eulerzn method.
78 Kinematics Chap. 3

z il meaningless until we specify where, in 1


aample, we measured the temperature
ever, the temperature of a fluid may be t
a thermocouple or sorne other device.
constant for the limiting process, we Wli

ar _ (d!'\ _lim [~
1 y Ot dt }r At_,O
/
/

.........1 / where oTfot is called the partial deriva


1 to measure the temperature of a fluid
1
___ ,-:::.
1 /-- the fluid. Under such conditions the 1
.............. constant, and the material derivative lJ
/

........ /
held constant, the temperature of a sin~
~~----~~----~~~-------------+X the temperature of a succession of fiuJ
xP(t) xP(t + fl f)
constant.
Fi. 3.1-1. Motion of a particle in space. Thus far we have indicated that the f
takes on special meaning when either t
held constant. lf neither is held cons
The derivative given by Eq. 3.2-2 is the time rate of change of a scalar, which
if either the function depends only 01
is measured as we move with the particle-i.e., we might imagine an observer
at which the function is measured has 1l
riding on the particle and continuously observing his position along the
Before we present specific equations
x-coordinate. This type of derivative, encountered previously in the study of
it will be helpful to give examples of ti
particle dynamics, is called a material derivative and denoted by
sidering a skin diver measuring the wat
The temperature may be a function of tij
Dxv = (dxv) = lim [xv(t + Llt) - xv(t)J (3.2_3)
time t. If the diver anchors himself at se
Dt dt R M .... O Llt
Material coordinates
held constant
of change of temperature that he meast

In this particular example, there is no question about the type of derivative


we obtain by the limiting process indicated by Eq. 3.2-l. The time derivative (~n.
of any quantity associated with a particle is necessarily a material derivative, lf he allows himself to drift with the cw
since the quantity can only be measured by an observer or a device which of change of temperature, the materia] ,
moves with the particle. For example, if we evaluate the time rate of change
of the temperature of the pth particle,

dTv = lim [Tv(t + Llt) - Tv(t)J (3.2-4) If the skin diver s energetic and moves ;
dt At_,O Llt
the total derivative, dTfdt, which depend
it is obviously a material derivative. velocity w, and the partial derivative wit
Consider now the difficulties encountered in specifying the time rate of We would now like to formulate thel
change of the temperature of a fluid. The deriva ti ve temperature is a function of the spatial e
this functional dependence as
dT = lim [T(t + Llt) - T(t)J (3.2-5) T= T(1
dt At_,O Llt
Kinematics Chap. 3 lec. 3.2 nme O.rlvatlvea 79

ia meaningless until we specify where, in space, T is measured. In the previous


eumple, we measured the temperature while moving with a particle; how-
ever, the temperature of a fluid may be measured at a point fixed in space by
a thermocouple or sorne other device. lf the spatial coordinates are held
constant for the limiting process, we write

ar = (d!'\ = lim [T(t + At)- T{t)J (3.2_6)


at dt }r At-+0 !it
Spatial coordillateo
/ beld C:ODIIaDI

........ /
1
where arat is called the parta/ derivative. It is possible, but not practical,
1
1
to measure the temperature of a fluid with sorne device which moves with
___
,_..,
____
)~~- the fluid. Under such conditions the material coordinates would be held
constant, and the material derivative DT/ Dt would be detemiined. lf R is
/

..,.. /
/
held constant, the temperature of a single fluid particle is measured, whereas
H.f)
the temperature of a succession of fluid particles is measured if r is held
constant.
of a particle in space. Thus far we have indicated that the time rate of change of a scalar function
takes on special meaning when either the spatial or material coordinates are
time rate of change of a scalar, which held constant. lf neither is held constant, the derivative is meaningful only
.e., we might imagine an observer if either the function depends only on time or the velocity of tlie point
observing his position along the at which the function is measured has been specified.
encountered previously in the study of Before we present specific equations for the material and total derivatives,
derivative and denoted by it will be helpful to give examples of the three types of derivatives by con-
sidering a skin diver measuring the water temperature in the Big Sur River.
xv(t + Llt) - Xv(t)J (3 _2 _ )
3
The temperature may be a function of the spatial coordinates, x, y, and z, and
o[ Llt time t. lf the diver anchors himself at sorne point in the river, the time rate
Material coordinatcs of change of temperature that he measures is given by
held constant

question about the type of derivative (3.2-7)


by Eq. 3.2-l. The time derivative
is necessarily a material derivative,
If he allows himself to drift with the current while he measures the time rate
by an observer or a device which
of change of temperature, the material derivative is determined.
if we evaluate the time rate of change
(3.2-8)
(3.2-4) If the skin diver is energetic and moves about with a velocity w he measures
the total derivative, dTfdt, which depends on the spatial variations of T, the
velocity w, and the partial derivative with respect to time.
in specifying the time rate of We would now like to formulate these ideas in mathematical terms. The
The derivative temperature is a function of the spatial coordinates and time, and we express
this functional dependence as
+ (3.2-5) T= T(r, t) (3.2-9)
Sec. 3.2 Time Derivativa

lf the material coordinates are to be held constant, we express the spatial Returning now to Eq. 3.2-ll,
coordinates in terms of R and t by means of Eq. 3.1-3. coordinates holding R constant a
T = T(r, t) = T[r(R, t), t] (3.2-10) vector v, and the derivative of the ten
deriva tive.
= T[x(R, t), y(R, t), z(R, t), t]

Holding R constant, we differentiate with respect to time and get


DT = (o!'\ + v.,(~
Dt ot J a

(:~)a- ~~- (~~ (::)a + (~~ (:na + (~D (::)a + (~n.


Using vector notation, we can expr

DT =~
(3.2-11) Dt
where the chain rule has been used to obtain the first three terms. This step Noticing the repetition of x, y, an
always presents sorne difficulties, and it may be helpful to consider it further. quite naturally leads us to the use
If the student has had a course in thermodynamics, he is familiar with convention
functions of several variables. For example, consider the Gibbs free energy,
which may be expressed as a function of the temperature, volume, and
number of moles, n.
G = G(T, V,n) (3.2-12) If the location of the point at
The total derivative is with time owing to the motion of t
no longer held constant. Instead of
(oG) dT + (oG)
dG =
oT "" oV T,n
dV + (oG)
on
There is sorne curious notation associated with thermodynamics which is
rarely encountered elsewhere-i.e., the use of subscripts on the partial
T,v
dn (3.2-13)
(~n = (:n + (~~ (~t
where dxfdt, dyfdt, and dzfdt are
derivatives to indicate which variables are being held constant. In sorne which describes the diver's motion.
respects, this system is superfl.uous, because a partial derivative implies that
dT
"all other independent variables are held constant" during the limiting process.
However, in thermodynamics there is a wide choice of independent variables, dt
and the subscripts are used as a reminder that the independent variables are (in This result is the most general t
this example) T, V, and n. If T, V, and n are functions of time, Eq. 3.2-13 Equation 3.2-17 is justa special case
may be divided by dt to obtain derivative (i.e., w = 0).

dG _
dt
(oG)
oT V,n
(d!'\
dt }
+ (oG)
oV V,n
(dV)
dt .
+ (oG)
on T,v
(dn)
dt
( 3.2_14) The sky-dlver

Under these conditions, the functional dependence of G could have been As an illustration of the applica
expressed as
diver who has mistakenly strapped a
G = G[T(t), V(t), r(t)] (3.2-15) onto bis wrist. His error makes th
which is very similar to the functional dependence of T in Eq. 3.2-10 when R to gather sorne interesting meteorolo
is held constant. The only difference is that the temperature, T, depends ture is independent of time and dec
explicitly upon time in addition to the implicit time dependence via x(t), an altitude of 10,000 ft. Thus, we
y(t), and z(t). T=
Klnematlcs Chap. l
S.c. l.l Time Derivativa 81
be held constant, we express the spatial
Returning now to Eq. 3.2-11, we note that the derivatives of the spatial
means of Eq. 3.1-3.
coordinates holding R constant are the components of the fluid velocity
11r(R, t), t] (3.2-10} vector v, and the derivative ofthe temperature holding r constant is the partial
derivative.
t),y(R, t), z(R, t), t]
with respect to time and get DT =
Dt
(o!'\J + v:l)(o!'\
ot ox}
+ vv(o!'\ + v.(o!\
oy} oz}
(3.2-16)

(~D (~;)R + (~D (~;)R + (~n.


Using vector notation, we can express Eq. 3.2-16 as

(3.2-11) DT = oT + vVT (3.2-17)


Dt ot
to obtain the first three terms. This step
Noticing the repetition of x, y, and z in the last three terms of Eq. 3.2-16
it may be helpful to consider it further.
quite naturally 'leads us to the use of index notation and the summation
in thermodynamics, he is familiar with
convention
example, consider the Gibbs free energy,
of the temperature, volume, and
~~ = (~D + vi (~~ (3.2-18)

(3.2-12) If the location of the point at which the temperature is measured varies
with time owing to the motion of the skin diver, the material coordinates are
no longer held constant. Instead of Eq. 3.2-11, we write
dV + (aG) dn
,n

P-'V"'"'""
On T,v

with thermodynamics which is


(3.2-13)
(~D = (~D + (~~ e:) + (~~ (~n + (~v (~:) (3.2-19}
, the use of subscripts on the partial where dxfdt, dy/dt, and dzfdt are the components of the velocity vector w
bies are being held constant. In sorne which describes the diver's motion. In vector form, Eq. 3.2-19 becomes
because a partial derivative implies that
dT oT
constant" during the limiting process. -=-+wVT (3.2-20)
a wide choice of independent variables, dt ot
that the independent variables are (in This result is the most general type of time derivative we shall encounter.
and n are functions of time, Eq. 3.2-13 Equation 3.2-17 isjust a special case (i.e., w = v) ofEq. 3.2-20, as is the partial
derivative (i.e., w = 0).

oG) (dv) + (aG) (dn) (3.2_14)


OV .
dt . On
V n dt .
T" The sky...c:tlver
dependence of G could have been
As an illustration of the application of Eq. 3.2-20, let us consider a sky
diver who has mistakenly strapped a dial thermometer instead ofhis altimeter
V(t), r(t)] (3.2-15}
onto his wrist. His error makes the jump more precarious but allows him
dependence of Tin Eq. 3.2-10 when R to gather sorne interesting meteorological data. Assume that the air tempera-
is that the temperature, T, depends ture is independent of time and decreases linearly between ground level and
the implicit time dependence via x(t), an altitude of 10,000 ft. Thus, we may represent the temperature as
T= T0 - ru
82 Kinematia Chap. J S.C. J.l Time Derivativa

where tX = 5 X 10-3F/ft, and Eq. 3.2-20 reduces to Here we see that the acceleration C<l
of change ofvelocity ata fixed pointI
dT The second is called the convective a
- = -WtX (3.2-21)
dt magnitude of the velocity and the
An observer riding with a fluid p
In the free-fall stage, the sky diver's velocity is about 200 mph. Hence,
terms of a single vector, a; the fixe
local time rate of change of velocity,
dT = -{(-200mi)(5 x 10_
3
oF)}{(528~ft)( hr )} =1. 47 oF/ sec quantities he would deduce the accel
dt hr ft mt 3600 sec
i.e., v is not a function of time-Eq
and he experiences a rapid rise in temperature. After the parachute is opened, not necessarily zero.
the velocity decreases to about 5 mph, and the rate of temperature rise is As an example of a steady, accele
reduced to only 0.037F /sec. entrance region of the Reynolds ap
duced in Fig. 3.2-2. As a fluid par
The veloclty field

In the study of fluid motion, we will be concerned with fluid acceleration


and the forces which give rise t~ it. The time rate of change of velocity of a
fluid particle is the acceleration a, given by

a= (dv) = Dv (3.2-22)
dt R Dt
The idea of acceleration is most easily grasped if we picture ourselves as
moving with a system and continuously noting the rate at which our velocity
is changing. For example, acceleration is vividly experienced as we rapidly
pull away from a stoplight in a high-powered, low-slung, sports car. Under
these circumstances, we can both "see" the acceleration (by observing our
rapidly changing position) and "feel" the acceleration (in terms of the forces
required to produce it).
Consider now a person fixed in space observing our sports car as it moves
Flr. 3.2-2. Flow in the en
away from the stoplight. It is much more difficult for him to get the feel of
the acceleration, for he can only note the velocity and the rate at which the
velocity is changing with distance.
centerline, the velocity changes from
We have a similar difficulty in describing fluid acceleration from the point
of view of a fixed observer, for the acceleration is not expressed as "the time
rate of change of the velocity." Let us expand the right-hand side of Eq.
3.2-22 to express the acceleration as The time required for the particle
At and expressed as
ov vVv
a=-+ (3.2-23)
ot f
t
Local
Convective
ac:cderation
acoeleration Here " representa some average
Klnematlcs Chap. 3 S.C. 3.1 Time Derivativa 83

Here we' see that the acceleration consists of two terms. The first is the rate
of change ofvelocity ata fixed point in space and is called the local acceleration.
Thc second is called the convective acceleration, and it depends upon both the
(3.2-21)
magnitude of the velocity and the velocity gradients.
An observer riding with a fluid particle would describe bis acceleration in
is about 200 mph. Hence, terms of a single vector, a; the fixed observer would note the velocity, the
local time rate of change ofvelocity, and the velocity gradients, and from these
quantities he would deduce the acceleration. Note that if the flow is steady-
i.c., v is not a function of time-Eq. 3.2-23 indicates that the acceleration is
After the parachute is opened, not necessarily zero.
and the rate of temperature rise is As an example of a steady, accelerating flow, let us consider the flow in the
cntrance region of the Reynolds apparatus discussed in Sec. 1.2 and repro-
+
duced in Fig. 3.2-2. As a fluid particle moves from z to z !:u along thc

be concerned with fluid acceleration


time rate of change of velocity of a
by

(3.2-22)
'Zz 'Ziz+.U

grasped if we picture ourselves as 4z


noting the rate at which our velocity
is vividly experienced as we rapidly
low-slung, sports car. Under
the acceleration (by observing our
acceleration (in terms of the forces

observing our sports car as it moves


F11. 3.1-l. Flow in the entrance region of the Reynolds apparatus.
difficult for him to get the feel of
velocity and the rate at which the
ccnterline, the velocity changes from
fluid acceleration from the point
is not expressed as "the time
us expand the right-hand side of Eq.
The time required for the particle to move this distance may be denoted by
lit and expressed as.
+vVv (3.2-23) !:u
At=- (3.2-24)
t
Convectlve
aca:leratlon
v.
Here " reprcsents sorne average value of v The acceleration in the
8-4 Kinematics Chap. l
Sec. l.l The Divergence Theorem

z-direction is the rate at which v. changes with time and is therefore given by The proof of this result will be accc
a, = lim
.l.t-+0
(vlz+.l.z - v.l.)
At
= lim
.l.z-+0
[zJ.(v.lz+.l.z - v.I.)J
Az J(aa.)
-roz
dV
(3.2-25)
=v
%
(av.)
oz The volume "Y under consideratio
sume that the surface may be divide
Now let us compare this result with that given by Eq. 3.2-23. Forming n
the scalar product of Eq. 3.2-23 with k, we havet
z
ov,
a =-+v v
V
at (3.2-26)
n

= av. + v,(av.) + ~(av.) + v.(av.)


ot or r ao oz
For steady [(ov,fot) = 0], axisymmetric (v = O and v, independent of 0),
9
flow along the centerline (v, = 0), Eq. 3.2-26 readily reduces to
_J_]
a = v
%
(av.)
oz (3.2-27) 1
1
1
1
which is the previously derived expression for the acceleration along the 1
centerline. 1
1
1
l
3.3 The Divergence Theorem

The divergence theorem,t

f-r
V GdV=fG odA
.!#
(3.3-1)

is a necessary mathematical tool for developing the differential equations of


mass, momentum, and energy. In addition, it is very useful in formulating
the macroscopic balances and in solving certain problems. So me of the ideas
in this development have already been encountered in the discussion of surface where n is the outwardly directed
forces and buoyancy effects, presented in Chap. 2. Expansion of Eq. 3.3-1 called "x-y simple" or "x-y proj
gives of the surface is designated as z1(
With this in mind, we write the v

f (ax
-r
00
.: + oG" +
ay
00
az
). dV = f (Gzi. n + G11j n + G,k n) dA (3.3-2) Eq. 3.3-3 as

(~~) av ~J{b a, a,
.!#

t Here we have made use of an underived expression for v V in cylindrical coordinates.


t The
divergence theorem is also known as Green's theorem, Gauss' theorem, and
Ostrogradsky's theorem. -r z 11 1 (z .wl
Kinematics Chap. 3
Sec. 3.3 The Divergence Theorem 85
with time and is therefore given by
The proof of this result will be accomplished by first proving that

(3.2-25)
J(aG.)
-yaz
dV = J
g
G.k ndA (3.3-3)

The volume "f/' under consideration is illustrated in Fig. 3.3-1, and we pre-
sume that the surface may be divided into two regions by a curve along which
that given by Eq. 3.2-23. Forming
k, we havet nk =O (3.3-4)
z
Element of surface
orea, dA
n
(3.2-26)
+ ;(:;) + v.(::)
(v 6 =O and v. independent of 0),
3.2-26 readily reduces to

(3.2-27)

for the acceleration along the y

(3.3-1)

the differential equations of


it is very useful in formulating Fig. 3.3-1
certain problems. Sorne of the ideas
~~~'""+ 4 r4 ..:~ in the discussion of surface
where n is the outwardly directed unit normal. Volumes of this type are
in Chap. 2. Expansion of Eq. 3.3-1 called "x-y simple" or "x-y projectable." In the lower region, the position
of the surface is designated as z1(x, y), and in the upper region, z2(x, y).
With this in mind, we write the volume integral on the left-hand side of
n + Gflj n + G,k n) dA (3.3-2) Eq. 3.3-3 as
sC:~:,fl)

xnressi4JO for,. V in cylindrical coordinatcs.


as Green's theorem, Gauss' theorem, and f(:~) fJJ(~~)
dV = dz dy dx =f J[G.IzsC:~:.fll- G.l.1 c.,,fl)] dy dx
-y ., fl t(:l:,fl) ., fl (3.3-5)
Kinematics Chap. 3 S.C. 3.3 The Dlverence Theorem
86

Our limits of integration are now in terms of x and y; however, we wish to The divergence of a scalar times a
change these limits, representing them in terms of the surface area. In the V (Sb) =
lower region, the projected area dx dy is given by
In Gibbs notation, Eq. 3.3-13 ma
dx dy = -k n dA, Lower region (3.3-6a) show that they they are equal; how
of straightforward product differen
while in the upper region,
dx dy =k n dA, Upper region (3.3-6b) L(Sb) =
OX
These relationships are precisely those used in the proof of Archimedes'
principie; when applied to Eq. 3.3-5, they yield Since b is a constant vector, V b
(
f (~~z) dV = f Gzlzs<:t.vl k n dA - f Gzlz <.,.v (-k
1 1 n) dA (3.3-7)
f
~
b. vs d
~ As A1
Putting both integrals on the left-h
where A1 and A2 represent the lower and upper surfaces, respectively. Since
from the integral signs, we get
the sum of A1 and A2 is the total surface area, we may write

f (oGz)
~ oz
f Gzk dV =
d
odA (3.3-8) b. [J vs dl
Since bis an arbitrary, constant v~
This relationship holds for any scalar function G., provided G. and oG.foz also a vector) must be zero, and
are continuous in the volume "!'. The analysis can easily be extended to
obtained.
yield the two additional equations,
f VSd
(3.3-9) ~

The value of this result may be

f (oGay
~
11
) dV = fd
G11j n dA (3.3-10)
momentum equation for a static fl1

provided the volume is y-z simple and z-x simple. Addition of Eqs. 3.3-8,
3.3-9, and 3.3-10 gives Eq. 3.3-2, which completes the proof. The result is The stress vector for a static fluid
easily extended to more complex volumes if they can be split up into a finite
number of auxiliary volumes that meet the requirements of this proof.
. fto'-nother useful form of the divergence theorem may be obtained by and Eq. 3.3-18 becomes
wntmg
G=Sb (3.3-11) O=f PI
~

where Sisan arbitrary scalar, and bis an arbitrary, constan! vector. By this Application of the divergence ti
we mean that the scalar components of b can take on any value, but they
the area integral to a volume integr
must be independent of x, y, and z. Substitution of Eq. 3.3-11 into Eq. 3.3-1
same integral sign
yields
f V {Sb) dV = f Sb n dA (3.3-12)
~ d
Kinematics Chap. 3 S.C. 3.3 The Dlverence Theorem 87

of x and y; however, we wish to The divergence of a scalar times a vector is given by


in terms of the surface area. In the
is given by
V (Sb) = b VS + SV b (3.3-13)
In Gibbs notation, Eq. 3.3-13 may be proved by expanding both sides to
Lower region (3.3-6a)
show that they they are equal; however, in index notation the proof consists
of straightforward product differentiation.
(3.3-6b)
(3.3-14)
used in the proof of Archimedes'
yield Since bis a constant vector, V bis zero and Eq. 3.3-12 takes the form

- f G.i.,<.,., (-k n) dA
.A,
(3.3-7) f
-r
b VS dV = f
.!11
Sb n dA (3.3-15)

upper surfaces, respectively. Since Putting both integrals on the left-hand side of the equation and removing b
area, we may write from the integral signs, we get

(3.3-8) b [f VS dV - I Sn dA J = O (3.3-16)

function G., provided G, and oG.foz Since bis an arbitrary, constant vector, the term inside the brackets (which is
analysis can easily be extended to also a vector) must be zero, and the divergence theorem for a scalar is
obtained.

(3.3-9)
f
-r
VS dV =f SodA
.!11
(3.3.17)

The value of this result may be demonstrated by applying it to the linear


momentum equation for a static fluid, which, in Chap. 2, was
(3.3-10)

z-x simple. Addition of Eqs. 3.3-8,


O= f pg dV +f ~> dA (3.3-18)
-r .!11
completes the proof. The result is
if they can be split up into a finite The stress vector for a static fluid is
the requirements of this proof. t(a) = -np
theorem may be obtained by and Eq. 3.3-18 becomes

(3.3-11) O= f
-r
pg dV - f
.!11
pn dA (3.3-19)
an arbitrary, constant vector. By this
b can take on any value, but they Application of the divergence theorem for a scalar allows us to transform
ofEq. 3.3-ll into Eq. 3.3-1 the area integral to a volume integral; hence, both terms can be put under the
same integral sign

= f
.!11
Sb odA (3.3-12) O =f.,. (pg- Vp) dV (3.3-20)
88 Kinematia Chap. 3 Sec. 3.4 The Transport Theorem

Since the limits of integration are arbitrary, the integrand must be zero, and
we obtain the previously derived equation for the pressure.
O= -Vp + pg (3.3-21)
The technique of extracting a differential equation from an arbitrary
volume integral is an important one and deserves clarification. For the
volume integral to be zero, the integrand must either be identically zero or
take on both plus and minus values. However, if there is sorne region where
the integrand is either positive or negative, we may form the integral over
that region and thus violate Eq. 3.3-20. It follows, then, that if the limits of
integration are arbitrary and the integral is equal to zero, the integrand must
be identically zero. This result requires, of course, that the pressure and
density be continuous functions.

*3.4 The Transport Theorem dY=-nw MdA 1

The objective of this section is to develop a general equation for the time
derivative of a volume integral, under conditions such that points on the
swept out by the moving surface shall
surface of the volume move with a velocity, w. By our previously discussed
volume left behind by V1 (~t). The s
nomenclature, this volume is designated as fa{t). The velocity w is a continu-
da(t), and may be split into two ar
ous function of the spatial coordinates and time, and may be set equal to the
surface along which n w = O. In a
fluid velocity vas a special case. The volume then moves with the fluid and is
divergence theorem, we will limit ou
a material volume designated by f m{t). Under these conditions, we refer to
separates the volume into two parts.
the derived result as the Reynolds transport theorem.
this simple situation by appropriate
Considering the arbitrary volume fa(t) illustrated in Fig. 3.4-1, we wish
Our first step in the derivation of
to determine the time derivative of the volume integral of sorne scalar
express the volume, "Ya(t + ~t), as
function, S. By definition,

J + ~t)
S(t dV - f S(t) dV]
r_Js dV = lim [
,-.<tHt>
~t
,-.(t) (3.4-1) thus allowing us to write the integral
dt
f.(t)
M-+0

To visualize the process under consideration, we must think of a volume, f S(t + ~t) dV =
7'"0 (t+Atl
J S(t
f(t)
+ Llt) dV
such as a sphere, moving through space so that the velocity of each point on
the surface of the volume is given by w. The velocity w may be a function of Substitution of Eq. 3.4-3 into Eq. 3.
of the spatial coordina tes (if the volume is deforming) and time (if the volume
is accelerating or decelerating). At each instant of time, we assume that sorne r!_f S dV
scalar quantity-such as the density, or the temperature, or a scalar com- dt
?'".(t)
ponent of the velocity vector-is evaluated and the integral of this quantity

~~ + ~~:)
over the volume is obtained. The time rate of change of the integral of this
scalar, designated by S, is given by Eq. 3.4-1. = lim[! bt) dV - [ dV
As in Sec. 3.2, the time derivative has no meaning unless the position in
space at which the function is measured is specified. In this case the infor- At-+0

mation is provided in terms ofthe velocity w. In a time ~t, the "new" volume
Kinematics Chap. 3 Sec. 3.4 The Transport Theorem 89
, the integrand must be zero, and n
on for the pressure.
(3.3-21)
equation from an arbitrary
and deserves clarification. For the
must either be identically zero or
if there is sorne region where
we may form the integral over
1t follows, then, that if the limits of
is equal to zero, the integrand must
of course, that the pressure and
n

dY=-nw MdA 1
Fig. 3.4-1. The moving volumeof';.(t).
a general equation for the time
conditions such that points on the
, w. By our previously discussed swept out by the moving surface shall be designated by Vu(D.t), and the "old"
as "Ya(t). The velocity w is a continu- volume left behind by V1(D.t). The surface area at time t is represented by
and time, and may be set equal to the ~a(t), and may be split into two areas-A 1(t) and Au(t)-by a curve on the
then moves with the fluid and is surface along which o w = O. In a manner similar to our derivation of the
. Under these conditions, we refer to divergence theorem, we will limit ourselves to the case where a single curve
theorem. separates the volume into two parts. More complex cases can be reduced to
a(t) illustrated in Fig. 3.4-1, we wish this simple situation by appropriate formation of auxiliary volumes.
the volume integral of sorne scalar Our first step in the derivation of the transport theorem requires that we
express the volume, "Ya(t + D.t), as

(3.4-2)
(3.4-1) thus allowing us to write the integral in Eq. 3.4-1 as

we must think of a volume, J S(t + D.t) dV = J S(t + D.t) dV + J S(t + D.t) dVII - J S(t + D.t) dV 1
so that the velocity of each point on i'"a<t+dt) i'"a(t) VII(dt) Vr(dl)
The velocity w may be a function of (3.4-3)
is deforming) and time (if the volume Substitution of Eq. 3.4-3 into Eq. 3.4-1 yields
instant of time, we assume that sorne
the temperature, or a scalar com- c!...Jsdv
dt
and the integral of this quantity i'".(t)

rate of change of the integral of this


3.4-1.
no meaning unless the position in
= lim[I.~~t + M dV - [ ~~:) dV +L~~"+ L'>t) dVn - t~ + L'>t) dV, J
is specified. In this case the infor- dt-+0 D.t
w. In a time D.t, the "new" volume (3.4-4)
Klnematlcs Chap. 3 S.C. 3.4 The Transport Theorem
90

Since the limits of integration on the first two integrals are the same, they can cylinder swept out by a moving
be combined to give L=

~Js dV = lim
dt 4t-+O
J [S(t + ~t)- S(t)J dV
~t
where w is the magnitude of the veloc
in terms of a unit vector l (which ~
7'"a(t) 7'"a(t) w
The volume dV is given .by the leng
I S(t + ~t) dVn - J S(t + ~t) dV1] dV=
+ lim [ Vu(4tl v 1 <4tl (3.4-5) However, we seek a relationship betv
4t-+0 ~t
dA. The relationship between the cr
We may change the order of the integrating and limiting processes in the and its surface area is
first term of Eq. 3.4-5 to obtain dAca=

J
~ S dV
. =J (as) dV + Iim
[f S(t
vu<At >
+ ~t) dVn - f
V<M
S(t + ~t) dV 1]

(3.4-6)
where the cosine of Ois given by the
cos 6
The use f either the plus or minus .
dt ot At--o ~t is greater or less than 1rj2. Equatio
7'"a(t) 7'"a(t)
to give
Our next step in this analysis is to change the volume integrals to surface dV=
integrals. Examining Fig. 3.4-2, we see that the length L of an oblique =
Applying this result to the differenti
dV1=
Surfoce al lime t dVn =
Using these results allows us to write
Surfoce oreo elemenl

+ ~t) dVII = +
dA al lime f+flf
I S(t
Vu(41)

I S(t
J7(4t)
+ At) dV = -

n
Substitution of Eqs. 3.4-14 into Eq.

Cross seclionol
~tI S dV = (~~) dV
1".(1)
f
1". (t)

[
oreo elemenl dAc,
~tf S(t +
+ lim ---=..4:!!nc:.::<t:....>-~
41-+0

Flg. 3.4-l. Differential volume swept out by a moving surface. Taking the liinit as indicated, and
Kinematlcs Chap. 3 S.C. 3.4 The Transport Theorem 91

two integrals are the same, they can cylinder swept out by a moving surface in a time !l.t is
L=w!l.t (3.4-7)
where w is the magnitude ofthe velocity vector w. This vector may be written
in terms of a unit vector A (which gives the direction) and the magnitude w.
" = AW (3.4-8)
Tbe volume dV is given .by the length times the cross-sectional area, dAca
dVu - JS(t + !l.t) dV 1]
dV = L dAca (3.4-9)
(3.4-5)
6.t However, we seek a relationship between the volume dV and the surface area
dA. The relationship between the cross-sectional area of the oblique cylinder
and Iiiniting processes in the and its surface area is
dAca = cos OdA (3.4-10)
where the cosine of Ois given by the scalar product between A and n.
+ At) dVu - J S(t + !l.t) dV1] cosO =A n (3.4-11)
fl.t Vr!At) (3.4-6) The use f either the plus or minus sign in Eq. 3.4-10 depends on whether O
is greater or less than 1rj2. Equations 3.4-7 through 3.4-11 can be combined
to give
the volume integrals to surface dV= wAtA ndA
(3.4-12)
see that the length L of an oblique = w n!l.tdA
Applying this result to the differential volumes, dV1 and dV11 , we write
dV1 = -w n !l.t dA 1 (3.4-13a)
dV11 = +w n !l.t dA 11 (3.4-13b)
Surfoce oreo element Using these results allows us to write the two volume integrals in Eq. 3.4-6 as
dA ot time f+M
JS(t + !l.t) dVu =
Yn!Ail
+!l.tJ S(t
.4:nW
+ At)w n dA 11 (3.4-14a)

J S(t
V(At)
+ At) dV1 = -!l.t JS(t +
..4(1)
!l.t)w n dA 1 (3.4-14b)
n
Substitution of Eqs. 3.4-14 into Eq. 3.4-6 gives

Cross sectionol
~~ I S dV =
1'".w
I e~)
1'".w
dV
oreo element dAcs

+ lim [
f
!l.t S(t
..4u<tl
+ At)w n dA 11 + !l.t f S(t
..4r!tl
+ !l.t)w n dA 1]
~~o M
(3.4-15)
out by a moving surface. Taking the limit as indicated, and noting that A1(t) + A 11(t) = d 4 (t), we
92 Kinematics Chap. 3
Conservation of Mass
Sec. 3.5

finally obtain the general transport theorem mass balance. The mass M containe
~~ f S dV = f (~~) dV + f Sw n dA General transport theorem (3.4-16) M=
7'"o(t) 7'"o(t) do(t)

An obvious special case of this result is a volume fixed in space. Then, w = O Since conservation of mass requires
and
:t f f e~) change of M must be zero.
(3.4-17)
(dM)
S dV = dV DM
7'" 7'"
;;- R= Dt
where the limits ofintegration have been changed from "Ya(t) to"Y to indicate
that the volume is fixed in space. The most important use of Eq. 3.4-16 is for Application of the Reynolds transpo
material volumes. Then w is equal to the fluid velocity v and we write

!!_f S dV. =fasat dV +f Sv n dA Reynolds transport theorem (3.4-18)


~~ f p dV =
7'" (t)
f(::)
7'".. (t)
Dt
7'",.(t) 7'",.(t) ..,,.(t) The divergence theorem may now be
This result is referred to as the Reynolds transport theorem. The nomen- a volume integral, allowing both tern
clature indicates that the volume is a material volume, "Y m(t), the surface is sign.
a material surface, d m(t), and the time derivative D/Dt is taken with the
material coordinates held constant.
~~f pdV = f[~
7'",.(t) 7'".,(t)
This derivation was presented in terms of a scalar Sto focus the students'
attention on the development of the transport theorem. The extension to Since the limits of integration are a
integrals of vectors is straightforward. For example, we need only apply the zero and the continuity equation res
transport theorem to the three scalars, v.,, v11 , and v.; then we multiply by
i, j, and k, respectively, and form the sum to obtain

~ f V dV =f av dV +f v(w. n) dA (3.4-19) This equation applies to every poi


dt at restriction that the material under
7'"o(t) 7'"0 (t) d (t)
0 nuclear reaction. Making use of Eq
In concluding this treatment of the transport theorem, a warning is necessary: scalar function, and noting that,
Much more satisfactory methods of deriving the transport theorem1- 2 are
V (pv) = v
available through the use of a coordinate transformation from spatial to
material coordinates. The development presented here is intuitive in nature we may write the continuity equatio
inasmuch as the geometry of curved, moving surfaces is more complex than Dp +
we have intimated. Dt
Equations 3.5-7 and 3.5-5 are entirely
3.5 Conservation of Mass to familiarize the student with a
equation. An especially important
We will now make use of the previous developments to formulate the incompressible flow, for which both
continuity equation, which may be descriptively referred to as the differentia/ V
l. R. Aris, Vectors, Tensors, and the Basic Equations of Fluid Mechanics (Englewood The continuity equation can be
Cliffs, N.J.: Prentice-Hall, Inc., 1962), p. 78.
2. C. Truesdell and R. Toupin, The C/assical Field Theories, Handbuch der Physik,
of the Reynolds transport theorem.
S. Flgge, ed. (Rerlin: Springer-Verlag, 1960), p. 347. S
Kinematics Chap. 3
Sec. 3.5 Conservation of Mass 93

.mass balance. The mass M contained in a material volume is given by


General transport theorem (3.4-16)
M= f pdV
j",.(t)
(3.5-1)

a volume fixed in space. Then, w = O


Since conservation of mass requires that M be a constant, the time rate of
change of M must be zero.

changed from "Ya(t) to "Y to indicate


(3.4-17)
(dM)
dt
=
Dt
= p__ p dV =
R
DM
Dt 7",.(t)
f O (3.5-2)

mostimportant use ofEq. 3.4-16 is for Application of the Reynolds transport theorem yields
the fluid velocity v and we write

Reynolds transport theorem (3.4-18)


~~ JP dV = J
7",A(j)
G:)
j" m(l)
dV + Jpv n dA =
d ,.(t)
O (3.5-3)

The divergence theorem may now be used to transform the area integral into
transport theorem. The nomen- a volume integral, allowing both terms to be written under the same integral
volume, "Y m(t), the surface is sign.
derivative Df Dt is taken with the
~~ f p dV =
7" ,.(t)
f[!: +V
7",.(t)
(pv)J dV =0 . (3.5-4)
of a scalar Sto focus the students'
transport theorem. The extension to Since the limits of integration are arbitrary, the term in brackets must be
For example, we need only apply the zero and the continuity equation results
v.,, V11 , and v.; then we multiply by
to obtain op +V (pv) =O (3.5-5)
ot
+ J v(w n) dA (3.4-19) This equation applies to every point in space, and is subject only to the
restriction that the material under consideratioli is a continuum free from
d.(t)
nuclear reaction. Making use of Eq. 3.2-17 for the material derivative .of a
theorem, a warning is necessary : scalar function, and noting that,
the transport theorem1- 2 are
nate transformation from spatial to V (pv) = v V p + pV v (3.5-6)
presented here is intuitive in nature we may write the continuity equation as
surfaces is more complex than
Dp + pV V = 0 (3.5-7)
Dt
Equations 3.5-7 and 3.5-5 are entirely equivalent, and the latter is given only
to familiarize the student with a commonly used form of the continuity
equation. An especially important form of this equation is that for an
developments to formulate the incompressible How, for which both Eqs. 3.5-7 and 3.5-5 reduce to
referred to as the differentia/
t-rn"t; ..,,.Jv
V .V= o (3.5-8)
The continuity equation can be used to derive an interesting special form
Field Theories, Handbuch der Physik, of the Reynolds transport theorem. If the scalar S is given by,
p. 347. S = pS (3.5-9)
S.C. J.S Conservatlon of Mus
94 Kinematics Chap. 3

z
substitution in Eq. 3.4-18 yields

!!_
Dt
f pS dV
.,.,.(t)
=f ~ ot
.,.,.(t)
(pS) dV +f pSv n dA
.14,.(t)

= f [~t (pS) + V (pSv)J dV (3.5-10)


.,.,.(t)

=J[P as+ s op + sv. (pv) + pv. vs] dV


ot ot
"f",.(t)
Here, the divergence theorem has been used to change the area integral to a
volume integral, and the products have been differentiated according to the
usual rules. Remembering the formula for the material derivative allows us
to express this result as

~~ J pS dV = J[P
.,.,.(t) .,.,.(t)
~~ + sG: +V. J
pv) dV (3.5-11)

The continuity equation may be used to show that the term in parentheses is
zero, and we obtain

-D
Dt
f
.,... w
pSdV
.,.,.(t )
= f DS
p-dV
Dt
Special form of the
Reynolds transport theorem (3.5-12) Flg. 3.5-1. Components
cylindrical coordinates.

This form of the Reynolds transport theorem will be especially useful in The derivation presented in this
deriving the differential momentum and energy equations. coordinates; for simplicity, all sub!
The expanded forms of the continuity equation are presented in Table restricted. This procedure allows us
3.5-1 for rectangular, cylindrical, and spherical coordinates. Alphabetical the same way as scalars, without r
subscripts are used to denote the physical components, which are the pro- matical techniques. The fundament
jections of the velocity vector on the tangents to the coordina te curves. These present a sufficient challenge to the
projections are illustrated for cylindrical coordinates in Fig. 3.5-1. linear coordinates need not be add
simpler to analyze in either cylind
Table 3.5-1
THE CoNTINUITY EQUATION
reason, derived results will always
systems. Satisfactory discussions of
Rectangular Coordinates (x, y, z) equations to the different coordinate
op o o o
at + ox (pv.) + oy (pv.) + oz (pv,) = o (a) An alternate derlvatlon of t he con
Cylindrical Coordinates (r, (J, z) equatlon
op 1 o 1 o o Another method of deriving the
-ot + -r -or (prv,) + -r u"(J (pv8) + ;-
uz
(pv,) = O (b)
require the use of the divergence th
Spherical Coordinates (r, (J, </>)
3. A. J. McConnell, Applications of Te
op 1 o 1 o . 1 o Inc., 1957).
~
ut
+ 1r ;- (prv,)
r~r
+ r-. - "(J (pv8 sm 8) + -.
sm 8 "
- ".L (pv~) = O
r sm 8 "'~'
(e)
4. G. E. Hay, Vector and Tensor Analys
Kinematics Chap. 3 S.C. J.S Conservatlon of Mass 95
z

JpSv n dA
V

d,.(t)

(pSv)J dV (3.5-10)

+ SV (pv) + pv VSJ dV

used to change the area integral to a


been differentiated according to the
for the material derivative allows us

s(~~ +V pv)J dV (3.5-11)

show that the term in parentheses is

Special form of the


Reynolds transport theorem (3.5-12) Fig. 3.5-1. Components of the velocity vector in
cylindrical coordinates.
theorem will be especially useful in The derivation presented in this section has been done in rectangular
energy equations. coordinates; for simplicity, all subsequent developments will be similarly
equation are presented in Table restricted. This procedure allows us to integrate and differentiate vectors in
spherical coordinates. Alphabetical the same way as scalars, without recourse to sorne more advanced mathe-
components, which are the pro- matical techniques. The fundamental concepts involved in fluid mechanics
to the coordina te curves. These present a sufficient challenge to the student so that the complexity of curvi-
coordinates in Fig. 3.5-1. linear coordinates need not be added. However, many problems are much
simpler to analyze in either cylindrical or spherical coordina tes; for this
reason, derived results will always be tabulated in these other coordinate
systems. Satisfactory discussions of the problem of transforming vector
equations to the different coordinate systems are given elsewhere.1.3-'

+a;iJ (pv,) =O (a)


An alternate derlvatlon of the continuity
equatlon
iJ
(pv6) + a; (pv,) = O (b) Another method of deriving the continuity equation exists that does not
require the use of the divergence theorem or the transport theorem; it does,
3. A. J. McConnell, Applications ofTensor Analysis (New York: Dover Publications
1 iJ Inc., 1957).
sin ll) + - . -6 ;- (pv) = O (e)
4. G. E. Hay, Vector and Tensor Analysis (New York: Dover Publications, lnc., 1953).
rsm ""'
Kinematics Chap. 3 Sec. 3.6 Streamlines, Path Lines, and St
96

however, involve a somewhat more lengthy algebraic effort. Nevertheless it Dividing Eq. 3.5-14 by l:lx Ay Az,
is instructive to examine this derivation, for it will give the student sorne l1z-+ O, we have
practice in handling "mass flux" terms. For a volume fixed in space, we
state the principie of conservation of mass as op + lim [pt
Ot Az -+0

time rate of change} {mass flux } tmass flux }


{
of the mass of the = into the - out of the (3.5-13) + lim
Av-+O
[eJ
volume control volume control volume control

The term "flux" is used repeatedly in reference to mass, momentum, and +lim
Az-+0
[~
energy transport and it means "flow per unit time." Thus, the mass flux into
the system is the rate at which mass flows into the system (lbm/sec). Each limiting process yields a part.
z pvz op o
Z+Z
ot + OX (pv.,) +
which is the previously derived Eq.
coordinates. The statement given
given by Eq. 3.5-2, which states that
It obviously turns out that both are
conservation of mass, the latter fol
easy intuitive deduction results fro
of mass is relatively simple; howev
equations, we will encounter m
matical tools presented in this cha~
derivation of the differential equat

*3.6 Streamlines, Path Lin

It is of sorne help in visualizing


sentation of the velocity fiel d. A st
Fig. 3.5-2. Mass balance on a differential cube fixed in space.
the velocity vector is everywhere
streamline and the stream function (
able importance and will be discu
Applying Eq. 3.5-13 to the cube illustrated in Fig. 3.5-2, we get A path line is a curve in spa
follow. The locus of points on a
o Mass flow rate
ot (p Ax Ay Az) = [pv.,J., - pv.,lx+Ax] Ay Az across x-surfaces r = r(R, t),
Mass flow rate Thus, the path Jine represents the
+ [pvulv - pvliJY+Au] Ax Az across y-surfaces (3.5-1 4)
motion.
Mass flow rate The streak fine represents the ft
across z-surfaces see easily, for it is a curve traced
Kinematics Chap. 3 Sec. 3.6 Streamlines, Path Lines, and Streak Lines 97

y algebraic effort. Nevertheless it Dividing Eq. 3.5-14 by !:u Ay Az, and taking the limits !:u-+ O, Ay-+ O,
for it will give the student sorne Az-+ O, we have
For a volume fixed in space, we
op + lim [pv.,/.,+11"'- pv.,i.,J
ot 11.,-+o Ax
mass flux }
- out ofthe
{
(3.5-13) + Iim [pv11/11+l1u - pv11/v] (3.5-15)
volume control 1111 .... 0 Ay
reference to mass, momentum, and + lim [pv./z+l1 - pv./.J = O
unit time." Thus, the mass flux into 11-+o Az
into the system (lbm/sec).
Each limiting process yields a partial derivative, and the result is
pvz
otop + ~
ax (pv.,) + ~ ~ (pv.) =o
ay (pvl) + az
Z+tlz (3.5-16)

which is the previously derived Eq. 3.5-5, Iisted in Table 3.5-1 for rectangular
1 .,. . . . .- pvy /r+tly
coordinates. The statement given by Eq. 3.5-13 is very different from that
given by Eq. 3.5-2, which states that the mass of a material volume is constant.

l., ...
!u pvxX+tlx
It obviously turns out that both are equivalent statements of the principie of
conservation of mass, the latter following intuitively from the former. This
easy intuitive deduction results from the fact that the principie of conservation
ofmass is relatively simple; however, in attacking the momentum and energy
equations, we will encounter more complex concepts, and the mathe-
y matical tools presented in this chapter will be valuable in obtaining a rational
derivation of the differential equations.

*3.6 Streamlines, Path Lines, and Streak Lines

It is of sorne help in visualizing various flows to have a geometrical repre-


sentation of the velocity fiel d. A streamline is a curve in space drawn so that
cube fixed in space. the velocity vector is everywhere tangent to the curve. The concept of the
streamline and the stream function (to be defined subsequently) is of consider-
able importance and will be discussed in detail in this section.
in Fig. 3.5-2, we get
A path fine is a curve in space that an individual fluid particle would
Mass flow rate follow. The locus of points on a path line may be expressed by Eq. 3.1-3 as,
x+Axl tly LlZ across x-surfaces
r = r(R, t), R = constant vector
Mass flow rate
tlx tlz across y-surfaces (3.5-14) Thus, the path line represents the Lagrangian method of describing the fluid
motion.
Mass flow rate The streak fine represents the fluid motion in a way that an observer can
tlx tly across z-surfaces
see easily, for it is a curve traced out by all particles passing through sorne
98 Kinematlcs Chap. l Sec. 3.6 Streamlines, Path Lines, and

z
fixed point, r 0 The plume of srnoke issuing frorn a burning cigarette repre-
sents a streak line, provided we neglect the lateral diffusion of the srnoke
particles.
When the flow is steady, iJvfot =O, the strearnline, path line, and streak
line coincide; however, they generally differ for unsteady flows. The rnost
lucid discussion of these ideas is contained in the rnovie entitled "Flow
Visualization. " 5
Returning to our analysis of streamlines, such as those shown in Fig. 3.6-1,
we note that if the flow is steady, they can be visualized by continuously

z V

Flg. 3.6-l. Unit

and the pararnetric representation


x = x(s),

~------------------------~x
Equation 3.62 rneans that if the .
of that point on the curve are d(
Flg. 3.6-1. Streamlines.
that if the spatial coordinates are g
frorn sorne arbitrary reference) is'
injecting a dye into the fluid at sorne fixed point. Inasrnuch as the streamline is 3.6-2 in vector notation as
defined in terrns of a tangent vector, we need to define the unit tangent
vector to a curve.
The unit tangent vector A is given
Unit tangent vector

It is easy to prove that A is a unit


The tangent vector to a curve is defined as the limiting position of the
secant of a curve, such as that illustrated in Fig. 3.6-2 The arclength s along
the curve is a function of the spatial coordinates A A = (~:) (~:)
s = s(x,y, z) (3.6-1)
dx +. dy
= (-
ds J-
ds 1

S. Prepared by Professor S. J. Kline of Stanford University and distributed by Educa


tional Services, Inc., 47 Galen St., Watertown, Mass. 02172.
= (::J + (~:J
Klnematlcs Chap. 3
Sec. 3.6 Streamlines, Path Lines, and Streak Lines 99
issuing from a burning cigarette repre- z
the lateral diffusion of the smokc

O, the streamline, path line, and streak


differ for unsteady fiows. The most
;uuJLi:IIJJJcu in the movie entitled "Flow

such as those shown in Fig. 3.6-1,


can be visualized by continuously

Flg. 3.6-2. Unit tangent vector to a curve.

and the parametric representation of the curve is


x = x(s), y = y(s), z = z(s) (3.6-2)
Equation 3.62 means that if the arclength is given, the spatial coordinates
Streamlines. of that point on the curve are determined; conversely, Eq. 3.6-l indicates
that ifthe spatial coordinates are given, the arclength to that point (measured
from sorne arbitrary reference) is determined. We may write Eqs. 3.6-1 and
point. Inasmuch as the streamline is 3.6-2 in vector notation as
we need to define the unit tangent s = s(r) (3.6-3a)
r = r(s) (3.6-3b)
The unit tangent vector A is given by
. ~r dr
A=hm- = - (3.6-4)
As-+o ~s ds
as the limiting position of the It is easy to prove that A is a unit vector by forming the scalar product
in Fig. 3.6-2 The arclength s along
A A = (~:) e:)
(3.6-1)
= ( -
. dx + J-
. dy + k -dz) ( dx + . dy
. ds- Jds -
+ k -dz) (3.6-5)
ds ds ds ds
University and distributed by Educa
Mass. 02172.
= (~:)2 + (~:)2 + (~:)2
Sec. 3.6 Streamlines, Path Lines, and Stre
lOO Kinematics Chap. 3

Figure 3.6-3 illustrates the streamlin


Since the square of the differential arclength is given by indicates an arbitrary closed curve
performed to define the stream funcf
ds 2 = dx 2 + dy + dz2 2 (3.6-6)
The mass flow rate, per unit lengt
we see that
A -. B is given by
(3.6-7) B

We may now define the streamline more explicitly by noting that the f pv nds =
velocity vector v divided by its magnitude v is a unit tangent vector to the A
streamline, and is therefore equal toA.

dr v y
A=-=- (3.6-8)
ds v
The three scalar components of Eq. 3.6-8 are

dx v.,
-=- (3.6-9a)
ds v
dy = V11 (3.6-9b)
ds v
dz =~ (3.6-9c)
ds v
where v.,, v11 , and v. are the three components ofthe velocity vector. We may
rearrange these three equations to give

dx dy dz
-=-=- (3.6-10)
v.,
These equations define the streamline, and we shall find that they are a
key part of the derivation of Bernoulli's equationt presented in Chap. 7.
In addition to being useful in mathematical analysis, streamlines are an aid
to obtaining a qualitative understanding of fluid motion. To illustrate this
point, we shall define the stream function 1p and indicate how a qualitative Fl. 3.6-3. Streamlines f
sketch of the streamline helps us visualize the flow field.
To keep matters simple,t we must restrict our discussion to two-dimen- where s is the arclength measured com
sional flows so that Eqs. 3.6-10 take the form
the mass flow rate across the surface
dx dy A
-=- (3.6-11)

t Bernoulli's equation is a sea lar component of the equations of motion in the direction
f
B
pv n ds =

of a tangent vector to the streamline when viscous effects are neglected.


:j: Fully three-dimensional flows give rise to two stream functions. See C. S. Yih, Since the mass in the region bounde<
Houille blanche, 1957, 12 :445.
Sec. 3.6 Streamlines, Path Unes, and Streak Unes 101
Kinematics Chap. 3

~clcmglth is given by Figure 3.6-3 illustrates the streamlines for a two-dimensional flow and also
indicates an arbitrary closed curve along which an integration will be
+ df + dz 2
(3.6-6) performed to define the stream function.
The mass flow rate, per unit length in the z-direction, across the surface
A _. B is given by
(3.6-7)
B {mass flow rate}
more explicitly by noting that the
v is a unit tangent vector to the J
A
pv n ds = across the
surface A -+ B
(3.6-12)

(3.6-8) y

=~ (3.6-9a)
V

=~ (3.6-9b)
V

v.
=- (3.6-9c)
V

dz
(3.6-10)

and we shall find that they are a


equationt presented in Chap. 7.
analysis, streamlines are an aid
ng of fluid motion. To illustrate this
1p and indicate how a qualitative Flg. 3.6-3. Streamlines for a two-dimensional flow.
the flow field.
restrict our discussion to two-dimen-
form where s is the arclength measured counterclockwise along the curve. Similarly
the mass flow rate across the surface B _. A is given by
(3.6-11)

f
A {mass flow rate}
pv 0 ds = across the (3.6-13)
of the equations of motion in the direction surface B -+ A
B
effects are neglected.
to two stream functions. See C. S. Yih,
Since the mass in the region bounded by the curves C1 and C2 is constant if
102 Kinematics Chap. 3 S.C. 3.6 Streamllnes, Path Llnes, and Strea

the flow is incompressible, it follows that 3.6-17, and placement of all the terms

JBV D dslalong C
1
+fA V D dslalong C2 = (3.6-14)
A B
Since this result is true for any arbitr
where the density has been removed, for it must be constant for this statement
arbitrary and the terms in parentheses
to be generally true. Changing the direction of integration along C2 changes
the sign, and Eq. 3.6-14 takes the form otp
B B v.., = oy,
Jvnds = Jvnds (3.6-15)
The derivation presented here proves
A along 0 1 A along 0 2 shows that it must satisfy Eqs. 3.6-22.
For the moment, think of C2 as sorne .fixed curve while C1 may be any While this result is of sorne interes
arbitrary curve. Thus, Eq. 3.6-15 indicates that the integral of v n ds along volumetric flow rate (per unit depth) lj
any arbitrary curve connecting two points, A and B, always has the same
_ volumetric
value-i.e., the integral is independent of the path. Under these circum- 'PB - 'PA - between p<
stances, v n ds is an exact differential and may be represented as
Thus, if A and B are any two points 01
dtp=vnds (3.6-16)
where 1p is independent of path. Expressing tp as a function of x and y, and "PB- ~

expanding the right-hand side of Eq. 3.6-16, we have and we conclude that the value of 1p i.
last result may not be obvious, but it i
(~:) dx + (~:) dy = v.,n., ds + vvnv ds (3.6-17)

The normal n to the curves C1 and C2 must be orthogonal to the tangent VJB- 'PA =f
A
vector given byt
Because the curve connecting A and B.
(3.6-18) line that passes through both points. ,
Equation 3.6-18 requires that V=i
dx dy
DA=n -+nv -ds= 0
"'ds
(3.6-19) by Eq. 3.6-8, and Eq. 3.6-25 becomes
and either
dy dx 'PB- 'PA =
n =- n =-- (3.6-20a)
"' ds ' 11 ds
or Inasmuch as A n is always zero, Eq. :
dy dx In Fig. 3.6-4, we consider two poir
n = - - n11 = - (3.6-20b)
"' ds' ds are parallel. For that case,
depending on which direction we assign to the unit tangent vector, A. If A
is taken in the same direction as s, examination of Fig. 3.6-3 will indica te that or
Eq. 3.6-20a is the desired solution. Substitution of Eq. 3.6-20a into Eq. VJB- 1
(
t Here A is the unit tangent vector to the curves C1 and C1 ll.s
Kinematics Chap. 3
S.C. 3.6 Streamlines, Path Llnes, and Streak Llnes 103
that
3o6-17, and placement of all the terms on the left-hand side, yield
A

J v o n ds/
along C 2
=O (3.6-14) (~: + vv) dx + (~;- vz) dy =O (3.6-21)

it must be constant for this statement Since this result is true for any arbitrary curve, the values of dx and dy are
rection of integration along c2 changes arbitrary and the terms in parentheses must be zero.

dVJ dVJ
V=-
B
ay, V=--
ax (3.6-22)
= Jvonds (3.6-15)
z v

The derivation presented here proves the existence of a stream function and
A along C2
shows that it must satisfy Eqs. 3.6-22.
fixed curve while C1 may be any
While this result is of sorne interest, it is of more interest to note that the
that the integral of v on ds along volumetric ftow rate (per unit depth) between two points A and B is
points, A and B, always has the same
of the path. Under these circum- _ _ {volumetric flow rate per unit depth}
'PB 'PA. - between points A and B (3.6-23)
and may be represented as

(3.6-16) Thus, if A and B are any two points on the same streamline,

tp as a function of x and y, and VJB- VJ.tt = 0 (3.6-24)


3.6-16, we have
and we conclude that the value of VJ is constant along any streamline. This
last result may not be obvious, but it is proved readily.
(3.6-17)
B B
c2 must be orthogonal to the tangent VJB - VJA. = f f
A.
d1p =
A.
V o D ds (3.6-25)

.dy Because the curve connecting A and Bis arbitrary, we can choose the stream-
+ Jds- (3.6-18)
line that passes through both points. Along this particular curve
V=VA (3.6-26)
. dy
+nv -=0
ds (3.6-19) by Eq. 3.6-8, and Eq. 3.6-25 becomes

f
B
dx 'PB - VJA. = VA o n ds (3.6-27)
n=-- (3.6-20a)
v ds
A.

Inasmuch as A on is always zero, Eq. 3.6-24 is proved.


dx
nv=- (3.6-20b) In Fig. 3.6-4, we consider two points, A and B, chosen such that n and .,.
ds are parallel. For that case,
to the unit tangent vector, A. If A VJB- VJ.tt =V l:J..s
of Fig. 3 .63 will indica te that
Substitution of Eq. 3.6-20a into Eq. or (3.6-28)
3-2. If the fluid velocity v is given by

where a has units of sec-1


b has units of n-1
e has units of n-2
u0 has units of ft/sec

obtain an expression for the fluid


Uo.

3-3. If the stream function is given by

Flg. 3.6-4. Qualitative description of a flow field using streamlines.

From this result, we can express the ratio of velocities as


determine the two components of
~ =~s2 (3.6-29) 3-4. (a) Sketch the streamlines in the 1
v2 ~s 1 two-dimensional velocity field.
and conclude that when the streamlines are close together the velocity is
high, and when the streamlines are far apart the velocity is Iow. This idea is V =
"'
especially useful if the streamlines are known; however, if they are sketched
in on the basis of intuition, we must view the results strictly as a qualitative V11 -
description of the ftow field.
(b) Is the continuity equation fo
velocity field?
PROBLEMS
3-5. The Leibnitz rule for differentia1
3-1. If the density of the air in the atmosphere is given by the formula, functions of the independent varia

P = Po[l + a(x2 + y2) + e-l>]


where Po = 0.075 Ibm/ft 3 4>(x) '
a = 1 x I0- 5/ft 2
then
b = 1 X 10--4/ft 1
and the velocity of an airplane flying through the atmosphere is given by,
d4>
dx =
Jb<"'( a) dy
ox + j
w = iw., + jw11 + kw. a(.,)

where w., = 300 mph Derive this result, which is a on


w11 = 20mph theorem. Hint: Start with the defin
w. = -20mph of integration and apply the mear
a(x) and a(x + L\x), and between '
how would the air density measured at the airplane change with time if the
coordinates of the plane are x = y = O, and z = 15,000 ft? 3-6. Express the Reynolds transport
Ans.: 4.9 x lo-~> Ibm/ft3-sec.
notation.
Kinematia Chap. 3
Problems 105

3-2. If the fluid velocity v is given by

v = u0 e-a 1[ibx + jey2]


where a has units of sec-1
b has units of n- 1
e has units of n- 2
u0 has units of ft/sec

obtain an expression for the fluid acceleration Dv/Dt in terms of a, b, e, and


uo.
3-3. If the stream function is given by

of velocities as
determine the two components of the velocity, v., and v11
(3.6-29) 3-4. (a) Sketch the streamlines in the region O ~ y ~ oo and O ~ x ~ L for a
two-dimensional velocity field.
are close together the velocity is
apart the velocity is low. This idea is
known; however, if they are sketched
view the results strictly as a qualitative

(b) Is the continuity equation for incompressible flow satisfied for this
velocity field?
3-5. The Leibnitz rule for differentiating an integral the limits of which are
is given by the formula, functions of the independent variable is given as follows: If
a(x2 + y2) + e-11]
f
b(o:)

rp(x) = f(x, y) dy
a(o:)
then
b(o:)

through the atmosphere is given by, drf>


dx
=J (a) dyox +
db
f[x, b(x)] dx -
da
f[x, a(x)] dx
+ jw11 + kw. a(z)

Derive this result, which is a one-dirnensional analogue of the transport


theorem. Hint: Start with the definition ofthe derivative; rearrange the limits
of integration and apply the mean value theorem for the integrals between
a(x) and a(x + ~x), and between b(x) and b(x + ~x).
at the airplane change with time if the
=O, and z = 15,000 ft? 3-6. Express the Reynolds transport theorem for a vector quantity in index
notation.
106 Klnematla Ct.p. 3

3-7. Greeo's theorem in aplane may be writteo as

f (Rdx + Sdy) = f (::- :;) dxdy


o

where C denotes a closed curve boundiog the region A, oriented such that the
region is on the left as we advance along the curve in the positive direction.
Prove this result using the same approach describe..d In the derivatioo of the Stress in a Flui~
divergence theorem. Note: When R = v., and S = v., this result is known as
Stokes' theorem.
3-8. Express the continuity equation using index notation.
3-9. Considering an arbitrary, fixed volume "f", apply the integral method to Eq.
3.5-13 to derive the continuity equation.
3-10. Using Eq. 3.5-13 and the differ-
ential volume shown in Fig. 3-10,
derive the continuity equation in
cylindrical coordinates.
3-11. If v.,, v., and v. are independent of
time, is the acceleration Dv/ Dt
necessarily zero?
3-12. Using Gibbs notation, prove that
V (Sv) = v VS + SV v 4.1 The Stress Vector
3-13. For a compressible, steady, two-
dimensional flow, show that a Fl. 3-10
In the previous chapter we de
stream function tp* can be defined of a continuum. Our purpose was
such that left-hand side of the linear momen
ay' ay
pV = -

*
"' ay
pV = - -
w ax
.!!._
Dt
J pv dV =
weretp
h -tp* = {mass ftow rate per unit depth} -r.. w
B between points A and B Tlmerate
or change
3-14. Prove that ofmomentum

~Jsdv=Jds
dt dt
dV We must now turn our attention
-;'".(t) -;'".(t)
which act upon the fluid. The b
surface force requires a good dea!
if w is independent of x, y, and z. chapter is to investigate the naturc
We assume that p, v, and t<> f
ordinates and time. In addition, t
of the orientation of the surface e
by the outwardly directed unit nl
order.
Klnematla Chip. 3

=J (as_ aR) dxdy


ax ay
.

the region A, oriented such that the


along the curve in the positive direction.
IPPJ'Oac:h describe_d In the derivation of the
= v., and S = v., this result is known as
Stress in a Fluid 4
1", apply the integral method to Eq.

4.1 The Stress Vector


Fl1. 3-10
In the previous chapter we developed certain ideas about the kinematics
of a continuum. Our purpose was to acquire the tools necessary to treat the
left-hand side of the linear momentum equation.
- a"'
h
pVy =

unit depth}
AandB
:t f pv dV =
;r...w
f pg dV
;r,.(t)
f
+ ~a> dA
~.,.(t)
(4.1-1)

Timerate Body Surface


ofchange force force
ofmomentum

=JdSdt dV We must now turn our attention to the problem of describing the forces
1".(1) which act upon the fluid. The body force presents no difficulty, but the
surface force requires a good deal of study; therefore, our objective in this
chapter is to investigate the nature of the stress vector, t<a>
We assume that p, v, and t<a> are continuous functions of the spatial co-
ordinates and time. In addition, t<a> is assumed to be a continuous function
of the orientation of the surface element under consideration, as designated
by the outwardly directed unit normal n. We shall prove the following in
order.
107
Sec:. 4.1 The Stress Vector
108 Stress in a Fluid Chap. 4

extended to area and volume integrals


l. The stress vectors acting u pon opposite sides of the same surface at a
given point are equal in magnitude and opposite in direction-i.e., (f)A=
t(n) = -~-D)

2. The stress vector may be written in terms of the stress tensor T as


(f)V=
t(n) = n T.

3. The stress tensor is symmetric, Tii = T11 As we mentioned previously, we shall


the mean or average value of sorne fun
In each case, we will be examining a special material volume, and we will to indicate whether an area or a volu
make use of the mean value theorem in applying both the linear and angular obvious.
momentum equations to these volumes. The mean value theorem in one In proving that t(n) = -t<-> we
dimension states that illustrated in Fig. 4.1-2. This volume
b

fj (b - a) =Jf(x) dx, (4.1-2)


"'=s

This result is represented graphically in Fig. 4.1-1, where we see that the area
under the curve of f(x) versus x is given by the integral in Eq. 4.1-2. It is
intuitively obvious that the area can also be represented by the product of
the 1ength b- a and sorne mean value ofthe functionf(x). The idea is easily

f(x)

a
Fl1. 4.1-l. Material volume havi

~------------------------~--------~~~X
x=o x=s x=b t. T. M. Aposta!, Mathematical AnalyJ
lishing Company, Inc., 1957), p. 269.
Fl1. 4.1-1. The mean-value theorem.
Stress In a Fluid Chap. 4 S.C. 4.1 The Stress Vector 109

opposite sides of the same surface at a extended to area and volume integrals,1 which are expressed as
and opposite in direction-i.e.,
(j)A = f fdA (4.1-3a)

in terms of the stress tensor T as


<J)V= jdVf (4.1-3b)
V
As we mentioned previously, we shall use angular brackets (()) to represent
the mean or average value of sorne function. No special notation will be used
special material volume, and we will to indicate whether an area or a volume average is meant, for it is usually
in applying both the linear and angular obvious.
The mean value theorem in one In proving that t<> = -t<-> we will make use of the material volume
illustrated in Fig. 4.1-2. This volume is formed by two parallel surfaces of

(4.1-2)

in Fig. 4.1-1, where we see that the area


by the integral in Eq. 4.1-2. 1t is
also be represented by the product of
ofthe functionf(x). The idea is easily

Fl. 4.1-l. Material volume having the form of an arbitrary slab.


x= & x=b
l. T. M. Aposta), Mathematical Analysis (Reading, Mass.: Addison-Wesley Pub-
1ean-value theorem. lishing Company, Inc., 1957), p. 269.
110 Stress In a Fluid Chap. -4

area A,.(t) connected by a strip perpendicular to the two parallel surfaces of


width L and area A,(t). Application of Eq. 4.1-1 to this volume yields

~~ f pv dV =
-;"',.(t)
f pg dV +
-;"',.(t)
f t 0, 1 dA+
..t,(t)
f t,., dA+
..t.(t)
f t,_., dA
..t.(t)
(4.1-4)
e
where the area integral ofthe stress vector has been split into the three distinct
areas forming the slab. Applying the mean value theorem to the first three
integrals, and putting the last two under the same integral sigo, we get

~~ [(pv)LA,.(t)] = (pg)LA,.(t) + (t0 , 1)A,(t) + f [t,. 1 + t,_.,] dA


..t.(t)
(4.1-5)

Taking the limit L-+ O and noting also that A,(t)-+ O, we obtain

O = lim
L ... O
J ['<-> + tc-a 1dA
1
(4.1-6)
..t.w
Since the limits of the integration are arbitrary, the integrand must be iden- X
tically zero, and the desired result is obtained.

(4.1-7)
Dividing by L 1 and taking the limit L
This result may appear to be intuitively obvious; however, we can easily
become confused when discussing the sigo of a particular stress component,
and it is thus worthwhile to have this result stated formally. In general,
careful use of vector notation helps greatly in eliminating errors in sigo in the
In its present form, Eq. 4.2-3 appears
solution of problems.
when applied to the tetrahedron illu:stral
interesting information. Table 4.2-1
auociated with the four planes of the
4.2 The Stress Tensor

In this section, we wish to demonstrate first that the stress is in equilibrium


Plane Arca
at a point in space, and then use this result to relate the stress vector to the
stress tenso_.!. We assume that any arbitrary material volume may be repre- ABC 4A
sented by a characteristic length L and a shape factor oc(t), thus allowing us BCD Df 4
to express the material volume as ADC aJ 4A
ABD Dk4A
-r"m(t) = oc(t)LS (4.2-1)
In Eq. 4.2-3, we let L1 equal the
Applying the mean value theorem to Eq. 4.1-1 in conjunction with Eq. 4.2-1, theorem to the four plane surfaces
we obtain
!!.. [(pv)oc(t).e] = (pg)oc(t)!J +f t,., dA (4.2-2)
1im
._. {.l.
...,M [dA(t,.,) + n i dA(tc-u) +
Dt
.At,.w
Stress In a Fluid Chap. 4 111
to the two parallel surfaces of
t..... ,,... ,..,... z
Eq. 4.1-1 to this volume yields

dA + f t<al dA + f t<-al dA (4.1-4)


..{ft(l) ..{ft(l)
e
has been split into the three distinct
mean value theorem to the first three
the same integral sign, we get

+ f [t(a) + t(-a)] dA (4.1-5)


A.W

that A,(t) ~O, we obtain

(4.1-6)

X
Fl1. 4.2-1. A material volume in the form of a tetrahedron.
(4.1-7)
Dividing by L 1 and taking the limit L ~ O, we have the desired result.
obvious; however, we can easily
sign of a particular stress component,
result stated formally. In general,
O= 1im ~
L-+o.L;
f
~.w
t(a) dA (4.2-3)

in eliminating errors in sign in the


In its present form, Eq. 4.2-3 appears to be a rather obtuse result; however,
when applied to the tetrahedron illustrated in Fig. 4.2-1, we will obtain sorne
interesting information. Table 4.2-1 indicates the stress vector and areas
usociated with the four planes of the tetrahedron.
Table4.2-1
STIU!SSES AC11No ON A l'lmtAm!DRON
first that the stress is in equilibrium
Plane Arca Normal Stress Vector
result to relate the stress vector to the
material volume may be repre- A.BC D t,.,
a shape factor oc(t), thus allowing us BCD -1 t,_ll
A.DC -J t,_.,
A.BD -k t,_k,
(4.2-1)
In Eq. 4.2-3, we let L1 equal the oblique area dA and apply the mean value
. 4.1-1 in conjunction with Eq. 4.2-1, theorem to the four plane surfaces making up the area integral.

(4.2-2)
~,{ [dA(t<al) + n i dA(t(-11) + n JdA(t<-n) + n k M(t<-kl)]} = O
(4.2-4)
112 Stress in a Fluid Chap . .of Sec. -4.2 The Stress Tensor

Note that taking the limit ~A -4- O is comparable to taking the limit L -4- O, (having magnitude and direction) as
for the characteristic length and the oblique area both go to zero simul- as the stress tensor (being a doubly dir
taneously. Carrying out the division by ~A inciicated in Eq. 4.2-4, we have sor. In this text, the word "tensor" wi
A qualitative description of a tensor is
lim [(t(n) + n i(t(-1) + n j(t<_n) + n k(t(-k)] =O (4.2-5) a vector nicely from the qualitative poin
4A-+O
line segment, which obviously has m
As ~A -4- O, the mean values are all evaluated at a single point, and we write and mathematical definition. Moreo
t(n) = - [(n i)t(-l) + (o j)t(-l) + (o k)t(-k)] (4.2-6) with a piece of chalk and a blackboar
on the blackboard, and its meaning m
Using Eq. 4.1-7 to effect a change ofsigns, we get to the realm of mathematical definiti~
it seems best to view the stress tensor
t(n) = [(n i)t(i) + (n j)t<n + (o k)t(k)] (4.2-7)
with the normal n, yields the stress vect
Equation 4.2-7 indicates that the stress vector acting on a surface having a n. The definition given by Eq. 4.2-9 m
normal n can be expressed in terms of the stress vectors acting on the three can be of any use to us.
coordinate planes. This result is of considerable importance, for now the Since the stress vector acting on a~
dependence of t<n> on n is known, and our analysis of stress is greatly in terms of the stress vectors acting o
simplified. we need only know the scalar compo
We now write Eq. 4.2-7 in the form the surface forces in general. In
the three stress vectors as
(4.2-8)
t< 1> = iT.,., + jT., +k T.,
where the quantity in brackets is defined as the stress tensor T, leading us to
write
t<n> = n T (4.2-9)

The stress tensor is composed of terms of the type it< 1>, which is neither the t<~<> = iT., + jT. 11+ kJ1
scalar (dot) product nor the vector (cross) product of two vectors. It is often
called a dyad 2 and represents a type of multiplication which is generally new The nine scalar components given n
to the student and deserves sorne discussion. In going from Eq. 4.2-7 to the stress tensor T. The stress tensor
Eq. 4.2-8, we have in effect stated that T.,.,
(A B)C = A (BC) (4.2-10) T= T11.,
(
and the question naturally arises, "Why should this be so?" The answer is T.,
that it is so by definition, and the explanation is as follows. The left-hand side The first subscript indicates the pla
of Eq. 4.2-10 consists vf the scalar product A Banda vector C. Multiplica- second subscript indicates the directi
tion of a vector C by a scalar A B is a straightforward operation yielding a the stress acting on the z-surface in the
vector. Obviously, the left-hand side of Eq. 4.2-10 is a well-defined quantity. in Fig. 4.2-2 showing the stresses a
It follows, then, that Eq. 4.2-10 defines the dot product between a vector A accordance with Eq. 4.1-7, the stress
and a tensor (BC)-i.e., it is the operation that yields the vector (A B)C. +i are oppositely directed to the stre
To view the stress vector in its proper perspective, we might note -i. The sign convention adopted her
that a more general description of functions would designate a scalar plane having a positive outwardly d
(a quantity having magnitude only) as a zero order tensor, a vector itself also acts in the positive directi
2. J. W. Gibbs andE. B. Wilson, Vector Analysis (New Haven, Cono.: Y ale University
the stress is also considered positive
Press, 1901). surface having a negative outwardly
Stress in a Fluid Chap. <4 Sec. <4.2 The Stress Tensor 113

is comparable to taking the limit L -+ O, (having magnitude and direction) as a first-order tensor, and a tensor such
oblique area both go to zero simul- as the stress tensor (being a doubly directed quantity) as a second-order ten-
by ~A indicated in Eq. 4.2-4, we have sor. In this text, the word "tensor" will always mean a second-order tensor.
A qualitative description of a tensor is a rather elusive tem. We can describe
(4.2-5) a vector nicely from the qualitative point ofview by referring to itas a directed
line segment, which obviously has meaning consistent with physical rea1ity
evaluated at a single point, and we write and mathematical definition. Moreover, we can construct it very nicely
(n j)tc-JJ + (n k)t(-k)] (4.2-6) with a piece of chalk and a blackboard. However, we cannot draw a tensor
on the blackboard, and its meaning must therefore be more closely confined
signs, we get to the realm of mathematical definition. From the physical point of view,
(4.2-7) it seems best to view the stress tensor as that quantity which, when dotted
with the normal n, yields the stress vector acting on a surface having a normal
vector acting on a surface having a n. The definition given by Eq. 4.2-9 must be expanded considerably before it
of the stress vectors acting on the three can be of any use to us.
considerable importance, for now the Since the stress vector acting on any arbitrary surface may be expressed
and our analysis of stress is greatly in terms of the stress vectors acting on the coordinate planes by Eq. 4.2-8,
we need only know the scalar components of tu, tw, and t(k) to determine
the surface forces in general. In rectangular coordinates, we write
the three stress vectors as
(4.2-8)
Force per unit area
as the stress tensor T, leading us to
tc 1> = iT.,., + jT:tll + kT.,, acting on the x-surface
(4.2-lla)

Force per unit area


acting on the y-surface
(4.2-11b)
(4.2-9)
Force per unit area
of the type itc 1l, which is neither the (4.2-llc)
acting on the z-surface
) product of two vectors. It is often
of multiplication which is generally new The nine scalar components given here represent the nine components of
discussion. In going from Eq. 4.2-7 to the stress tensor T. The stress tensor is often represented in matrix form as
T.,., Tz 11 T.,,)
(4.2-10) T = T!l., T!/11 T!IZ
(
should this be so?" The answer is T., T, 11 T
.Jla11av11 is as follows. The left-hand side The first subscript indicates the plane upon which the stress acts, and the
A Banda vector C. Multiplica- second subscript indicates the direction in which the stress acts. Thus T,z is
a straightforward operation yielding a the stress acting on the z-surface in the x-direction. This notation is illustrated
of Eq. 4.2-10 is a well-defined quantity. in Fig. 4.2-2 showing the stresses acting on the x-surfaces of a cube. In
the dot product between a vector A accordance with Eq. 4.1-7, the stresses acting on the plane having a normal
that yields the vector (A B)C. +i are oppositely directed to the stresses acting on the plane having a normal
proper perspective, we might note -i. The sign convention adopted here is that the stress acting on a coordinate
functions would designate a scalar plane having a positive outwardly directed normal is positive if the stress
as a zero order tensor, a vector itself also acts in the positive direction. A consequence of Eq. 4.1-7 is that
Analysis (New Haven, Conn.: Yale University the stress is also considered positive if it acts in the negative direction on a
surface having a negative outwardly directed normal.
114 Stress in a Fluid Chap. 4 Sec. 4.2 The Stress Tensor

z bracketed term. This correspondence


notation. Writing the three scalar co
allows us to express the stress vector
t(n) = it(n)z
Referring to Eq. 4.2-14, we see that

t<n>., = n.,T.,.,
= n.,T.,
t<n> 11 11

t<nlz = n.,T.,. -

The verbal description of the scalar t


area acting on the n-surface in the x-
interpreted in the same manner as ti
the stress tensor.
Replacement of x, y, and z with
t(n)l = n Tn
Fig. 4.2-2. Stresses on a cube. t<nl2 = n1Tl2

t(n)a = n Ta
The unit normal may be expressed in terms of its scalar components as The repeated ndices are easily expre
n = in., + jn + kn. (4.2-12) tion, and we write
11
l(n)l '
When substituted into Eq. 4.2-7, Eq. 4.2-12 gives
t(n)2 '
(4.2-13)
t(n)a '

lf we now substitute Eqs. 4.2-11 into Eq. 4.2-13, we obtain an expression for Note that any Ietter could be used fo
the stress vector in terms of the three scalar components of the normal and would still have the same meaning. 1
the nine scalar components of the stress tensor. the values 1, 2, and 3, we can repreJ

i[n.,T.,., + n T + n.T.,]
11 11.,
t(n); =

t(nl = +j[n.,T.,11 + n11 T1111 + n.T.11 ] (4.2-14) which is the index notation analogue
veys a great deal of information onc
+k[n.,T.,. + n"T + n.T ]
11 ndices is understood. The only drawt
the scalar components of the stress ve
In arriving at this result, we have made no use of the stress tensor T or the quantity is left to the reader's imagim
operation of dotting the normal with the stress tensor. Since that is a de- a general definition of the dot produc
fined operation, we are simply Iaying the groundwork for a workable defini- vector e is represented as,
tion. Notice that the subscripts in Eq. 4.2-14 occur in a very ordered fashion. e=
The first subscript on the stress component is always the same as the subscript
on the scalar component of the normal, and the second subscript indicates a then the scalar components of e are
direction corresponding to the particular base vector multiplying each C; =
Stress in a Fluid Chap. 4 Sec. 4.2 The Stress Tensor 115

bracketed term. This correspondence willlogically lead us to the use of index


notation. Writing the three scalar components of t(n) as t(n)z t(n)1l and t(n)z
allows us to express the stress vector as

t(n) = it(n)z + jt(n) + kt(n)z


11
(4.2-15)
Referring to Eq. 4.2-14, we see that the. scalar components become

t(n)z = n"'T"'"' + n T 11 11"' + n.T."' (4.2-16a)


t(n)v = n"'T"'" + n T 11 1111 + n.T. 11
(4.2-16b)
t(n)z = n"'T"'. +n T11 11 + n.r (4.2-16c)
The verbal description of the scalar t<n>z might be given as "the force per unit
area acting on the n-surface in the x-direction." Thus, the two subscripts are
interpreted in the same manner as the subscripts on a scalar component of
the stress tensor.
Replacement of x, y, and z with 1, 2, and 3 yields
t(n)l = n1 T11+ n2T21 + n3 T31 (4.2-17a)
t(n)2 = n1Tl2 + n2T22 + naTa2 (4.2-17b)
t(n)a = nTa + n2T2a + naTaa (4.2-17c)
in terms of its scalar components as The repeated ndices are easily expressed in terms of the summation conven-
(4.2-12) tion, and we write
t<nll = n; T; 1 (4.2-18a)
t(n)2 = n; T;2 (4.2-18b)
(4.2-13) t<n>a = n;T;a (4.2-18c)
4.2-13, we obtain an expression for Note that any letter could be used for the repeated ndices and the equations
components of the normal and would still have the same meaning. Remembering that a free index takes on
tensor. the values 1, 2, and 3, we can represen! Eqs. 4.2-18 by the single equation,
(4.2-19)

(4.2-14) which is the index notation analogue of Eq. 4.2-9. The result obviously con-
veys a great deal of information once the convention for repeated and free
ndices is understood. The only drawback to this expression is that it deals with
the scalar components ofthe stress vector, and the directional nature of this
no use of the stress tensor T or the quantity is left to the reader's imagination. Equation 4.2-19 may be taken as
the stress tensor. Since that is a de- a general definition of the dot product between a vector anda tensor. If the
groundwork for a workable defini- vector e is represented as,
14 occur in a very ordered fashion. e=AB (4.2-20)
is always the same as the subscript
, and the second subscript indicates a then the scalar components of e are given by
base vector multiplying each Ci = A;B; (4.2-21)
116 Stras In a fluid Chap. 4 Sec. -4.2 The Stress Tensor

This result could just as well be expressed as Most often we will be concerned wi
(4.2-22) and the normal vector will have o
e~= B~A~
cylindrical surface illustrated in Fig.
for the summation convention is in no way altered by this rearrangement. directed unit normal vector are
Things are not so simple in Gibbs notation, for in general A B and B A
are different, and it will be helpful to examine this difference.
Returning to Eqs. 4.2-8 and 4.2-9, we see that the normal vector n was
used as a prefactor in forming the dot product with the stress tensor. This
means that when the dot product was formed, the scalar multiplication was and Eq. 4.2-19 yields
carried out with the first vector (or the anteceden!) of each dyad. We now t(n)r
examine the process of using nas a postfactor and write an equation analogous
to Eq. 4.2-8. ((D)8

t<> = [it(l) + jt(l) + kt(l<)] n (4.2-23) t<>


Carrying out the scalar multiplication in this case, we get These equations are obviously consj
vector shown in Fig. 4.2-3.
t<> = [i(t<> n) + j(t(l) n) + k(t(l<) n)J (4.2-24)

Substitution of Eqs. 4.2-11 and expansion yield


i[T_n.. + Tzo/111 + T,..n.]
t<> = + j[Tr.JJ,. + T....n, + T11.n.] (4.2-25)
+k[T...n.. + T.,n, + T.n.]
Comparison of this result with Eq. 4.2-14 indicates that the diagonal terms
(n,.T....,, n11 T1111 , n.T.,) are the same; however, the off-diagonal terms have the
subscripts reversed. In index notation, Eq. 4.2-25 becomes

t<>l = Tlinl (4.2-26)


and the Gibbs notation analogue of this result is

t<> = T n (4.2-27)
Fl. 4.1-3. Stress o
In the next section, .we will show that the stress tensor is symmetric, leading
us to the result
(4.2-28) The subscripts for the componet
meaning in cylindrical coordinates a
Although we will not prove it, Eq. 4.2-19 is also valid for curvilinear Thus, Tr 8 represents the stress actit
coordinate systems provided the scalar components of a vector are taken as We should remember that the deri'
the projections of the vector on the unit tangent vectors to the three coordi- stress vector and the stress tensor fo:
nate curves. For a cylindrical system, 1, 2, and 3 represent r, 8 and z; in the realm of understanding of any s
spherical coordinates, these numbers represent r, 8, and cf>. For example, in calculus. However, the results for ot
cylindrical coordinates, Eq. 4.2-17b would be written scope of this text, and Eq. 4.2-19 mt
(4.2-29) linear systems.
Stras In a Fluid Chap. 4 Sec. 4.2 The Stress Tensor 117

Most often we will be concerned with forces acting on coordinate surfaces,


(4.2-22) and the normal vector will have only one nonzero component. For the
cylindrical surface illustrated in Fig. 4.2-3, the components of the outwardly
no way altered by this rearrangement. directed unit normal vector are
for in general A B and B A nr = n1 = 1
examine this difference.
we see that the normal vector n was n2 = n8 =O
product with the stress tensor. This n3 = nz =O
formed, the scalar multiplication was and Eq. 4.2-19 yields
antecedent) of each dyad. We now
t(a)r = Trr
and write an equation analogous
t(a)e = Tre
(4.2-23) t(D)z = Trz
in this case, we get These equations are obviously consistent with the components of the stress
vector shown in Fig. 4.2-3.
n) + k(t<t> n)J (4.2-24)

+ TJI. + T..,,n,]
+ Tr11r + T.,n.] (4.2-25)
+ T,,n. + T,,n,]
4 indicates that the diagonal terms
thc off-diagonal terms have the
On~IIPVPr,
Eq. 4.2-25 becomes
(4.2-26)

(4.2-27)
Flg. 4.2-3. Stress on a cylindrical surface.
the stress tensor is symmetric, Ieading

(4.2-28) The subscripts for the components of the stress tensor have the same
meaning in cylindrical coordinates as they did in rectangular coordinates.
4.2-19 is a1so valid for curvilinear Thus, Tre represents the stress acting on the r-surface in the 0-direction.
components of a vector are taken as We should remember that the derivation of the relationship between the
tangent vectors to the three coordi- stress vector and the stress tensor for rectangular coordinates is well within
1, 2, and 3 represent r, O and z; in the realm of understanding of any student having the usual background in
rcpresent r, O, and cf>. For example, in calculus. However, the results for other coordinate systems are beyond the
be writtcn scope of this text, and Eq. 4.2-19 must be accepted without proof for curvi-
n,T,. + n,T,, (4.2-29) linear systems.
Chap. 4 Sec. 4.3 Symmetry of the Stress Tensor
Stress in a Fluid

certain transformation law. The de


Transformation of tensors group of nine scalar quantities that
Eq. 4.2-38.
In Chap. 1, it was shown that a rotation of coordinates from x, y, z to
x', y', z' (or xi to x;) gave rise to a transformation of the components of a
vector expressed as *4.3 Symmetry of the Stres
(4.2-30)
where 0, 1 was the angle between the X~c-axis and the x;-axis. Thus, 012 is the So far we have gained considera
angle between x 1 and x~. For convenience, we will write Eq. 4.2-30 as continuum by applying the linear m '
volumes. W e can learn still more b
(4.2-31) tion toa third material volume. Fa
the stress tensor, we first prove loe
for a coordinate transformation X~c---+ x~. Similarly, the old components
this result to the differential cube
(unprimed) can be related to the new components (primed) by
(4.2-32) z
Note that the k and j in these two equations could be replaced by any index
since they are repeated or "dummy" indices. In addition, the 1 and i, being
free indices, could be represented by any letter. The only requirement here
is that an index is not repeated three times, for the expression

(4.2-33)
is obviously ambiguous. Expressing the stress vector for a rotation of axes
gives
(4.2-34)
Application of Eq. 4.2-19 yields an expression in terms of the stress tensor
and the normal
(4.2-35)
In liiZht of Eq. 4.2-32, we can express the normal vector n 1 as
(4.2-36)
Substitution of Eq. 4.2-36 into Eq. 4.2-35, and placement of both terms on
Flg. 4.3-1. Stresl
the left-hand side of the equation give us

(4.2-37) material volume, the angular mom


Because n~ is arbitrary and the terms in the bracket are independent of n~,
they must be zero. The transformation law for tensors results. ~t f (r x pv) dV =
.Y m(t)
f(
-y m
(4.2-38) Time rate of T
cbange of angular te
momentum fe
In Chap. 1, we pointed out that the mathematical definition of a vector
required only that it be a group of three scalar quantities which obeyed a provided all torques are the mome
Stress in a Fluid Chap . .of Sec. .of.l Symmetry of the Stress Tensor 119

certain transformation law. The definition of a tensor requires that it be a


group of nine scalar quantities that o bey the law of transformation given by
Eq. 4.2-38.
rotation of coordinates from x, y, z to
transformation of the components of a

*4.3 Symmetry of the Stress Tensor


(4.2-30)
X~;-axis and the x;-axis. Thus, 012 is the So far we have gained considerable knowledge of the nature of stress in a
we will write Eq. 4.2-30 as continuum by applying the linear momentum equation to two special material
(4.2-31) volumes. We can learn still more by applying the angular momentum equa-
tion toa third material volume. Following the procedure used in developing
----. x~. Similarly, the old components the stress tensor, we first prove local equilibrium of torques and then apply
components (primed) by this result to the differential cube shown in Fig. 4.3-1. For an arbitrary

(4.2-32)
z
could be replaced by any index
indices. In addition, the 1 and i, being
any letter. The only requirement here
times, for the expression

(4.2-33)
stress vector for a rotation of axes

(4.2-34)

(n 1T11)rx1; (4.2-35)
the normal vector n 1 as

(4.2-36)
, and placement of both terms on
Fig. 4.3-1. Stresses on a differential cube.

.] =o (4.2-37) material volume, the angular momentum equation is


in the bracket are independent of n~,
law for tensors results. ~t f (r
"Y,.(t)
X pv) dV = f (r
"Y,.(t)
X pg) dV + f (r x t<n>) dA
d ,.(t)
(4.3-1)
(4.2-38) Time rate of Torque owing Torque owing
change of angular to body to suace
mathematical definition of a vector momentum force force
scalar quantities which obeyed a provided all torques are the moments of forces-i.e., there are no local stress
Sec. 4.4 The Stress Equations of Motion
120 Stress in a Fluid Chap. 4

By examining the other two scalar co


couples 3 acting on the surface Sil m(t). Once again we represent the material
volume as T,
"f"'m(t) = 1X(t)L3 (4.3-2)
and apply the mean value theorem to the volume integrals in Eq. 4.3-1.
In index notation, these results are o
.P._ [(r X pV)1X(t)~] = (r X pg)1X(t)~
Dt
+f (r X t(n)) dA
.W.,.(t)
(4.3-3) T;

For the cube illustrated in Fig. 4.3-1, the characteristic length is taken as
L = ~x =~y= ~z. Since the position vector r is a directed line segment *4.4 The Stress Equations o
from the origin to any point on the cube, it tends to zero as L ---+ O. Dividing
Eq. 4.3-3 by L3 and taking the limit L---+ O, we reach the result that torques Our objective here is to derive the
are in local equilibrium. continuum in terms of the compon

O= lim ~
L-+o[;
f (r X t<n>) dA
.W.,.(t)
(4.3-4)
Eq. 4.1-1, and, to keep the mathemat
product of Eq. 4.1-1 with an arbitra
allows the scalar components, b.,, b
As in the case when we proved local equilibrium of stress, the significance they are independent of the spatial o
of this result will not be apparent until we apply it to a special case. lf we can be taken inside integral signs and
form the scalar product of Eq. 4.3-4 with the base vector k, an equation for and Eq. 4.1-1 may be written as
the z-component of the torque will result. Application to the cube shown in
Fig. 4.3-1 gives ~t f
-r.,.(t)
pv b dV = f
-r
Lever Lever Lever
arm Force arm Force arm Force Before continuing with this analys
scalar product with the vector b eli
x-surfaces dealing with the divergence of a ten
flux tensor, pvv. The Iatter will be t ,
defined during the course of this de
y-surfaces (4.3-5) to introduce these new ideas as leis
4.4-1 instead of Eq. 4.1-1 allows us

+ !:[r..,
2
L
2
- ., J
T L
2
- !::[T 1 13- T, 1 13]}
2
11 11 z-surfaces
We now make use of the special
(see Eq. 3.5-12, Sec. 3.5), taking v
i =0
Force
=A
=A
Force
=0.
Eq. 4.4-1 as
Lever
arm
Lever
arm

where the averaging brackets ( ( )) ha ve been dropped to simplify the notation.


f p.P._ (v b) dV =
-r.,.(t)
Dt

The two underlined terms are the only ones which do not cancel as L ---+ O; Expressing the stress vector in term
thus, carrying out the indicated division by L 3 and taking the limit, we get
t(n) b = (n
(4.3-6) yields
or Sea lar

3. D. W. Conditf and J. S. Dahler, "Fluid Mechanical Aspects of Antisyrnrnetrical


(4.3-7a)
J ~t p (v rb) dV = J
Stress," Phys. Fluids, 1964, 7:842:54.
"~'" ..w "~'" ...
Sec. -4.-4 The Stress Equations of Motion 121
Stress in a Fluid Chap. 4

By examining the other two scalar components of Eq. 4.3-4, we can also show
Once again we represent the material
(4.3-7b)
(4.3-2)
(4.3-7c)
volume integrals in Eq. 4.3-l.
In index notation, these results are conveniently expressed as
+ J
d,.(t)
(r X to)) dA (4.3-3) (4.3-8)

the characteristic length is taken as


vector r is a directed line segment *4.4 The Stress Equations of Motion
it tends to zero as L--+ O. Dividing
--+ O, we reach the result that torques Our objective here is to derive the differential equations of motion for an y
continuum in terms of the components of the stress tensor. We start with
Eq. 4.1-1, and, to keep the mathematics ata familiar leve!, we form the scalar
(4.3-4) product of Eq. 4.1-1 with an arbitrary constant vector b. Remember that this
allows the scalar components, b.,, hv, and h. to take on any value provided
equilibrium of stress, the significance they are independent of the spatial coordinates and time. For this reason, b
we apply it to a special case. If we can be taken inside integral signs and differential operators wh~never necessary,
the base vector k, an equation for and Eq. 4.1-1 may be written as

~tf pvbdV= JpgbdV+ Jt<ol bdA


Application to the cube shown in
(4.4-1)
..r,.(t) ..r,.(t) d ,.(t)
Lever
arm Force Before continuing with this analysis, it should be noted that forming the
scalar product with the vector b eliminates, for the present, the necessity of
+L[ Txvi.::.J x-surfaces dealing with the divergence of a tensor, V T, or the convective momentum
flux tensor, pvv. The latter will be treated in Chap. 6, and the former will be
defined during the course of this development; however, it seems reasonable
- L[ rv{:~J y-surfaces ( 4.3-5) to introduce these new ideas as leisurely as possible, and working with Eq.
4.4-1 instead of Eq. 4.1-1 allows us to do this.
- !::[rvl I3 - T.vl 13]} z-surfaces
We now make use of the special form of the Reynolds transport theorem
(see Eq. 3.5-12, Sec. 3.5), taking v b to be the scalar S. We may thus write
2 =A =.
j Force Eq. 4.4-1 as
Lever
arm Jp ~t (v b) dV J = pg b dV + Jt 1ol b dA (4.4-2)
dropped to simplify the notation. "''",.(t) "''",.(t) ,.(t)
ones which do not cancel as L--+ O; Expressing the stress vector in terms of the stress tensor,
by L 3 and taking the limit, we get
(4.4-3)
(4.3-6) yields

(4.3-7a)
(4.4-4)
Mechanical Aspects of Antisymmetrical
122 Stress in a Fluid Chap. 4 Sec. 4.4 The Stress Equations of Motion

The area integral may now be transformed to a volume integral by the diver- Inasmuch as the scalar components ~
gence theorem, and all terms can be put under the same integral sign. must be zero. The scalar components

J{p ~/v
"l'".,.(t)
b)- pg b- V (T b)} dV =O (4.4-5)
Eq. 4.4-10. To convey as much infon
f by the symbol V T and write

Since the limits of integration are arbitrary, the integrand must be identically (aa
1 -
zero and the stress equation of motion for the b-direction is obtained.
D
= pg b +V (T b)
p - (v b) (4.4-6)
Dt
We now wish to eliminate the vector b to obtain the general vector form ofthe
stress equations of motion.
The vector b can be removed easily from the first two terms in Eq. 4.4-6;
however, removal from the last term gives rise to a new mathematical opera-
tion (the divergence of a tensor), and we need to perform this step carefully. We may now rearrange Eq. 4.4-6
Let ils define a vector f by the equation,
Dv
V(Tb)=fb (4.4-7) p - - pg-
[ Dt
Note that the operation T b has been previously defined and yields a vector.
The divergence of a vector is a familiar operation; thus, the left-hand side of Once again, the term in brackets mus
Eq. 4.4-7 is well defined and it follows that the vector fis determined. Follow- stress equations of motion are obtait
ing the result given by Eq. 4.2-25, we express the term T b as
i(T.,.,b., + T"""b11 + T.,.b.) Mass pcr Bod
unit volume unit

+j(TII.,b., + T1111bll + T!l.bZ)


T. b = (4.4-8) ~
+k(T.,b., + T. 11h11 + T..b.) Dv
Taking the divergence of this vector and setting it equal to f b, we get p Dt =
a
- (T.,:r:b., + T., 11 b11 + T.,.b.) i
ax Acoeleration

a
+oy (T,.,b., + T,llbll + T,.b.) = f.,b., + Jllbll + J.b. (4.4-9) Here we see that our result is nothin

+ -oza (T.,b., + T. 11 b11 + T b.)


momentum principie per unit volumc
and the student may well have seeiJ
course on solid mechanics or "strenl
Carrying out the differentiation (remembering that b.,, b11 , and b. are constant) Remember that we started our a
and putting all the terms on the left-hand side, we have formulation of the linear momentum
[( ar.,., oT11., ar.,)
ax + ay + ~ -f., b.,
J was valid for any material volume-
space and time. Use ofthe Reynolds
material derivative inside the inte~
+ [(ar.,11 + ar,ll + ar.!l) - ,]b
11 =o (4.4-to) forces in terms of the stress tensor
ax oy az 11
formed to a volume integral. All t~
+ [ (ar.,.
ax +ay + az
ar11 J
ar ) _ , b

under the same volume integral and
resulted.
Stress in a Fluid Chap. 4 Sec. .of.of The Stress Equations of Motion 123

to a volume integral by the diver- Inasmuch as the scalar components of b are arbitrary, each term in brackets
under the same integral sign. must be zero. The scalar components of the vector f are therefore defined by
Eq. 4.4-10. To convey as much information as possible, we denote the vector
-V (T b)) dV =O (4.4-5)
f by the symbol V T and write

, the integrand must be identically .(o T.,.,- +oT- +oT-.,)


1 11 .,

for the b-direction is obtained. 0X oy oz


b +V (T b) (4.4-6) f -v
- T-
- +.(oT.,
J - +oTy
11ar.ll)
- +oz- 11
(4.4-11)
OX oy
to obtain the general vector form ofthe
ar.,.+oT
+ k (- - +ar )
az-
11
from the first two terms in Eq. 4.4-6; ox oy
rise to a new mathematical opera-
we need to perform this step carefully. We may now rearrange Eq. 4.4-6 into the form

(4.4-7) [p ~; - pg - V T Jb = O (4.4-12)
previously defined and yields a vector.
operation; thus, the left-hand side of Once again, the term in brackets must b~ zero because bis arbitrary, and the
the vector f is determined. F ollow- stress equations of motion are obtained.
express the term T b as
+ T"'~~b11 + T.,.b.) Mass per Body force per
unit volume unit volume
+ Tl/1/bl/ + TI/Zb.) (4.4-8)
~ ~
+ T.yb + T..b.)
11 Dv
p - =pg+VT (4.4-13)
setting it equal to f. b, we get Dt

i
Acceleration
i
Suace force per
unit volume

(4.4-9)
Here we see that our result is nothing more than an expression of the linear
momentum principie per unit volume. The result is valid for any continuum,
and the student may well have seen the right-hand side of Eq. 4.4-13 in a
course on solid mechanics or "strength of materials."
bering that b.,, b11 , and b. are constant)
Remember that we started our analysis with the integral or macroscopic
side, we have
formulation of the linear momentum principie (see Eq. 4.1-1). This equation
ar.,.)
~-J., b.,
J was valid for any material volume-i.e., any body moving and deforming in
space and time. Use ofthe Reynolds transport theorem allowed us to put the
material derivative inside the integral sign, and representing the surface
(4.4-10) forces in terms of the stress tensor permitted the area integral to be trans-
formed to a volume integral. All terms in the equation could then be put
under the same volume integral and the differential or microscopic equation
resulted.
124 Stress in a Fluid Chap. 4 Problems

4-4. If the components of stress T.. , T. 11 ,


PROBLEMS dimensional and is called "plane s
point,
4-1. Starting with Eq. 4.4-l, use index notation and the special form of the Tzz = 1
Reynolds transport theorem given by Eq. 3.5-12 to derive the stress equations TI/V=
of motion. Note: Make use of the result that
Tzv = 1

a aT;i find the plane (designated by n1) u


-;--- (T;;b;)
uX;
=-a-
Xj
b;

when b; is a constant vector. and the plane (designated by n2) U]


is a maximum
4-2. (a) How many distinct components are there in a symmetric tensor? (Tn 8)m
A skew symmetric tensor is one for which
Note that A n2 = O. Hint: Let 1
trigonometric identities
cos 2 O=
(b) How many distinct components are there in a skew symmetric tensor? sin 2 O =
4-3. If the stress distribution in a body is uniform and given by sin 20-
Ans: O = 1651' for maximum non
Tzz = 500 psia Tzv =O O = 61 51' for maximum sh
T1111 = 1000 psia Tzz = 500 psia
r.. = 1000 psia T11 = -500 psia
4-S. Label the stresses shown in Fig. 4
scripts and indicate whether they
sign convention.
what is the normal stress Tnn on the plane ABCD ofthe parallelepiped shown
in Fig. 4-3?
Ans: 561 psia

1
1
)--
,/
, ,/

,l"

X
Fi. 4-l Fi.
Stress in a Fluid Chap. 4 Problems 125

4-4. lf the components of stress T T. y. T.,, are zero, the stress distribution is two-
dimensional and is called "plane stress." Given the following stresses at a
point,
notation and the special form of the T.,., = 100 psia
Eq. 3.5-12 to derive the stress equations T.1111 = - 50 psia
result that T XY = T11., = 50 psia
find the plane (designated by n1) upon which the normal stress is a maximum
(Tnn)max = n1 (n1 T)

and the plane (designated by n2) upon which the tangential (or shear) stress
is a maximum
are there in a symmetric tensor? (Tn,)max = A (n 2 T)
which
Note that A n2 = O. Hint: Let n., = cos O and n11 = sin O, and use the
trigonometric identities
cos 2 O = lO + cos 20)
are there in a skew symmetric tensor? sin 2 O = lO - cos 20)
is uniform and given by sin 20 = 2 sin Ocos O
Ans: O= 16.51' for maximum normal stress
TXI/=0 O = 61 51' for maximum shear stress
T.,. = 500 psia
4-S. Label the stresses shown in Fig. 4-S with alphabetical and numerical sub-
T11 = -500 psia scripts and indicate whether they are positive or negative according to our
sign convention.
plane ABCD ofthe parallelepiped shown

Fil. 4-5
126 Stress in a Fluid Chap. 4 Problems

1
4-6. If the x and y axes are rotated z,z as
counterclockwise 30 about the D
z-axis, as shown in Fig. 4-6, what Dt (pv.,
are the values of T~.,. T~v and T~. because the mass of the cube p A
in the rotated coordinate system y'
for the uniform stress distribution 4-9. An "inviscid" (or nonviscous) flui
described in Prob. 4-3? Use Eq. stress vector is given by
4.2-38 to determine the three stress t(a) = -DP,
components; then use Eq. 4.2-14 Repeat Prob. 4-8 for this case and
to check your value for T~ . obtain
Hint: If i' is the unit vector direct-
ed along the x' -axis, then T.,.,' = P (: +v
X
i' . ten
4-10. If r is the position vector, prove
Flg. 4-6
D
4-7. If D; = -D;;, prove that Dt (r

4-11. lf gis a constant vector, what fo

where J.; is an arbitrary vector. Hint: This may be done either by carrying 4-12. IfF = VS, prove that V xF =

out the double summation and regrouping the terms, or by relabeling the zero.
ndices in an appropriate manner. 4-13. The Kronecker delta d has the p
'
4-8. Apply the linear momentum principie to the differential, material cube shown
'
in Fig. 4-8, to derive the x-direction stress equation of motion. Note that the
Show that
time rate of change ofmomentum ofthecube(D/ Dt)(pv.,AV)may be expressed

z
Hint: Write out the left-hand sid~
4-14. Show that an arbitrary tensor B;,
Differential
material volume symmetric tensor by writing it as
B;1 = !(Bii-
4-15. Given the divergence theorem for

f
-r
VG,

derive the divergence theorem for

f
-r
VA,

~--~--------------------------~Y by letting G = b A. Use this re:


equations of motion without fon
mentum equation with the arbitra
4-16. Rederive Eq. 4.2-14 by first forn
X balances on the tetrahedron shov
Fig. 4-8 to obtain an expression for tea>
Stress in a Fluid Chap. 4 Problems 127
1
z,z as
D Dv~
Dt (pv~ AV)= P AV Dt
because the mass of the cube p A V is a constant.
4-9. An "inviscid" (or nonviscous) fluid is one that exerts no shear stress; thus, the
stress vector is given by
t(a) = -np, for an inviscid fluid

Repeat Prob. 4-8 for tliis case and generalize the result to three dimensions to
obtain

X p ( ~ + v Vv) = - Vp + pg
4-10. lfr is the position vector, prove that
Fig. 4-6
D Dv
-(r x v) = r x-
Dt ' Dt
4-11. lf g is a constant vector, what form must the scalar </> take so that g = -V.,?
This may be done either by carrying 4-12. lf F = VS, prove that V x F = C, i.e., the curl of the gra4ient of a scalar is
IPnmoirU!the terms, or by relabeling the zero.
4-13. The Kronecker delta ~ii has the property that
to the differential, material cube shown d =o, i # j
stress equation of motion. Note that the ~ii = 1, i =j
cube(D/Dt)(pv.,tl V)may beexpressed
Show that
OV Ov
~if-=-
OX OX
Hint: Write out the left-hand side in full.
Differential / 4-14. Show that an arbitrary tensor Bu can be split into a symmetric anda skew
material volume symmetric tensor by writing it as
B = }(Bil + B) + }(B - Bu)
4-15. Given the diyergence theorem for a vector,

f
-r
V G dV =
d
G n dA f
derive the divergence theorem for a tensor,

f V A dV = f A n dA
-r Jl'f

~------------~Y by letting G = b A. Use this result to derive the vector form of the stress
equations of motion without forming the scalar product of the linear mo-
mentum equation with the arbitrary constant vector b.
4-16. Rederive Eq. 4.2-14 by first forming separate x-, y-, and z-direction force
balances on the tetrahedron shown in Fig. 4.2-1 and then adding the results
to obtain an expression for t(a)
Sec. 5.1 The Vlscous Stress Tensor

The statement "stress is a linear functi


tion for solids obeying Hooke's law
surface" is a statement leading to a e
and "stress is a linear function of tl
equation for Newtonian fluids. Consti
The Differential tal postulates discussed in Chap. 1, are

Equations of Motion 5 observation; however, the use of tl


Newton's law of viscosity) to descri
laws of mechanics and the laws of co1
all continua, while constitutive equat
under limited circumstances. In a !
equations as there are specific materi
The first step in formulating a cons
very definite ideas we have regardin
As we indicated in Chap. 2, fluids a
which is independent of orientation.
inasmuch as it acts in a direction o
normal, we write

t<a> = n T
= -np,
We note once again that the minus sign
that t<a> represents the force exerted l
In the previous chapter we derived the three stress equations of motion.
It will be helpful in this developt
Dv parts, the first representing the pres!
p Dt = pg +V . T the stress arising from the deformal
define a symmetric unit tensor, 1, wh
With the continuity equation
A-l=
op + V (pv) = O
ot where A is a vector. Given Eq. 5.1-:
we have four equations which govern fluid motion. However, there are six ponents can be specified on the basi!
unknown components of stress and three unknown components of velocity; the dot product between a vector an'
therefore, we have nine dependent variables and only four equations. We [:ti/ = ~~., = /!/Z =
overcome this excess of unknowns by introducing a constitutive equation
relating the stress to the pressure and the velocity gradients thereby reducing l.,.,=/!/)
the unknowns to four (pressure and the three components of velocity). Thus, the off-diagonal components
ponents are unity. In matrix form, ti

*5.1 The Viscous Stress Ten sor

A constitutive equation is an empirical relationship that attempts to relate


the stress in a continuum to the manner in which the material is deformed.
128
Sec. 5.1 The Vlscous Stress Tensor 129

The statement "stress is a linear function of strain" yields a constitutive equa-


tion for solids obeying Hooke's law. "The stress always acts normal to a
surface" is a statement leading to a constitutive equation for inviscid fluids,
and "stress is a linear function of the rate of strain" yields a constitutive
equation for Newtonian fluids. Constitutive equations, like the four fundamen-
ial tal postulates discussed in Chap. 1, are developed on the basis of experimental

ton 5 observation; however, the use of the word "law" (in Hooke's law and
Newton's law of viscosity) to describe these results is unfortunate. The two
laws of mechanics and the laws of conservation of energy and mass apply to
all continua, while constitutive equations apply only to particular materials
under limited circumstances. In a sense, there are as many constitutive
equations as there are specific materials.
The first step in formulating a constitutive equation is to make use of sorne
very definite ideas we have regarding the state of stress for any fluid at rest.
As we indicated in Chap. 2, fluids at rest experience only a normal stress,
which is independent of orientation. We call this stress the pressure, and
inasmuch as it acts in a direction opposite to the outwardly directed unit
normal, we write
t(D) =o. T
(5.1-1)
= -np, for fluids at rest
We note once again that the minus sign in Eq. 5.1-1 results from our convention
the three stress equations of motion. that t<a> represents the force exerted by the surroundings on the system.
1t will be helpful in this development to split the stress tensor into two
parts, the first representing the pressure stress and the second representing
+VT r the stress arising from the deformation of the fluid. To do this, we must
define a symmetric unit tensor, 1, which has the property

AI=IA=A (5.1-2)
(pv) =O
where A is a vector. Given Eq. 5.1-2 as the definition of 1, the scalar com-
fluid motion. However, there are six ponents can be specified on the basis of the previously derived equation for
unknown components of velocity; the dot product between a vector and tensor. The result is
and only four equations. We
introducing a constitutive equation
/""" = JIIZ = 111. = . 1.11 = I.:r; = I:r;. = o (5.1-3a)

velocity gradients thereby reducing /zz = 11111 = I = 1 (5.1-3b)


three components of velocity).
Thus, the off-diagonal components are zero, and the three diagonal com-
ponents are unity. In matrix form, the unit tensor may be represented as
r
(5.1-4)
relationship that attempts to relate
in which the material is deformed.
Sec. 5.1 The Viscous Stress Tensor
130 The Differential Equations of Motion Chap. S

Using index notation leads us to the conclusion that the scalar components represents the nine scalar equations,
ofthe unit tensor are given by the Kronecker delta c5iJ and Eq. 5.1-2 becomes A.,.,=B.,., A""' =
Aic5H =A; AII.,=Bv., A1111 ;
where A .,= B., A.v
c5i; =o, i -=1= j
(5.1-5) In matrix form, we would represent t
c5ii = 1, i=j
We now define the viscous stress tensor, "' by the equation
T= -pi+"' (5.1-6)
where the condition stated by Eq. 5.1-1 requires that
where the elements ofthe A-matrix ar
"' = O, for fiuids at rest (5.1-7) the B-matrix. Index notation provi<
5.1-11,
Because T and 1 are symmetric tensors, it is necessary that "' be symmetric,
Ai;
and because all the off-diagonal terms of pi are zero, the off-diagonal terms
of T and "' are equal, Returning now to Eq. 5.1-6, we v
T.,II=T""' total stress tensor into the stress equ
TI/Z = TI/Z Dv
p - =pg
T .,= Tzz Dt
or, in index notation, Following Eq. 4.4-11 to obtain the
i -=1= j (5.1-8)
Equation 5.1-6 is the first tensor equation encountered in this text and it i[j__ (pi
OX
deserves further comment. A tensor equation is simply a way to convey a
great deal of information without recourse to an annoying amount of algebra. V (pl) = +i[j__
Consider a scalar equation of the type ox (pi
A=B (5.1-9) +k[j__ (p
ox
What this equation means is obvious: A is equal to B. Consis:ler a vector
equation of the type .____/ Inasmuch as the off-diagonal elem
A= B (5.1-10) diagonal elements are unity, this eq

which means that the sca/ar components of A are equal to the scalar com- v. (pi)= op
ponents of B. For rectangular coordinates we obtain ox
A.,=B., Substitution of Eq. 5.1-16 into Ec
derivative give the viscous stress eq
AII=B11
A.= B.
p(~;+vVv)
Use of vector notation allows us to express these last three equations as a
single equation. Similarly, if A and B are tensors, the tensor equation These equations are listed in Tabl
(5.1-11) cylindrical, and spherical coordina!
Differential Equations of Motion Chap. S Sec. 5.1 The Viscous Stress Tensor 131

conclusion that the scalar components represents the nine scalar equations,
ruu~;~;~ter
delta c5", and Eq. 5.1-2 becomes
A.,.,=B.,., AZ11 = BZ!I A:ez = Bzz
A!IZ = B!IZ A111/=B1111 A!IZ = B!IZ
i i= j A.:e = B.z A.11 = B.11 A = B
(5.1-5) In matrix form, we would represent them as
i=j
, "t', by the equation
(5.1-12)
(5.1-6)
-1 requires that
where the elements ofthe A-matrix are equal to the corresponding elements of
fluids at rest (5.1-7)
the B-matrix. Index notation provides the most informative form of Eq.
it is necessary that "t' be symmetric, 5.1-11,
of pi are zero, the off-diagonal terms (5.1-13)
Returning now to Eq. 5.1-6, we wish to substitute this expression for the
total stress tensor into the stress equations of motion to obtain
Dv
p- = pg- V (pi)+ V "t' (5.1-14)
Dt
Following Eq. 4.4-11 to obtain the divergence of the tensor, pi, we write
i i= j (5.1-8)

{:X (plzz) + ~ (pl11:e) + :z (pi.:e)]


is simply a way to convey a
to an annoying amount of algebra.
V. (pi)= +j[! (piZ!I) +:y (pll/11) +~ (pi.!I)J (5.1-15)

[a a + aza (pi )J
(5.1-9)
~ k ax (pire.)+ ay (pl!IZ)
: A is equal to B. Consider a vector
Inasmuch as the off-diagonal elements of the unit tensor are zero and the
(5.1-10) diagonal elements are unity, this equation becomes
of A are equal to the scalar com- v. (pi) = ap +j ap +k ap = Vp (5.1-16)
we obtain ax ay az
Substitution of Eq. 5.1-16 into Eq. 5.1-14 and expansion of the material
derivative give the viscous stress equations of motion

these last three equations as a


p(~; +vVv) = -Vp+pg+V"t' (5.1-17)

are tensors, the tensor equation These equations are listed in Tables 5.1-1, 5.1-2, and 5.1-3 for rectangular,
{5.1-11) cylindrical, and spherical coordinates.
132 The Differential Equations of Motion Chap. S Sec. 5.1 The Vlscous Stress Tensor

Table 5.1-1 Table


VISCOUS STRESS EQUATIONS OF MOTION IN RECTANGULAR VISCOUS STRESS EQUATIONS
COORDINATES (x, y, z) CooRDINA

x-Direetion r-Direetion

av. av. av. av.) op + pg. + ( -0Tzz + -0Tvz + -OT)


( ar + V"' -ax + Vv-
p -
ay + V az = - -
ax 1 -
ax ay az (a)

y-Direetion

(b)

z-Direction
0-Direction

(e) ov9 v ov9 ~ ov9 --.!:L ~


p ( ot + r or + r ao + r sin o a

Table 5.1-2
1ap [1 a
rOO + pge + -r1 -or (
= - --
VISCOUS STRESS EQUATIONS OF MOTION IN CYLINDRICAL
CooRDINATES (r, 0, z)

r-Direetion
f>-Direction
OVr OVr Ve OVr V~
p -
( ot
+ v,-
or + -
r-ao - r- + v .ovr) op
- = - -
oz or OV. V OV. ~ OV. --.!:L
p( ot + r or + r ao + rsin o
10
+ pgr + [ -- 1 0T9r T99 0Tzr]
r or (rTrr) + -
r ao- - -r + -oz (a)

0-Direction

OVe OVe Ve OVe VrVe OVe~ 1 op


p
( ot + v,-
-
or + -r-ao + - r + v .oz- - - - -
r ao
(b)
In the next section, the viscous stre
z-Direetion
of the rate of strain tensor d for a Ne
by
av. ov, ~ av. av.) _ op 'f = 2,ud + [
p ( ot + v, or + r ao + v. oz - - oz or
(e) 'T = 2,u d + [ (
Differential Equations of Motion Chap. S Sec. 5.1 The Viscous Stress Tensor 133

Table 5.1-3
OF MOTION IN RECTANGULAR VISCOUS STRESS EQUATIONS OF MOTION FOR SPHERICAL
(x, y, z) CooRDINATES (r, IJ, </>)

r-Direetion

op
-- (OTzz OT.. OT)
ox + pg. + -ox + -oy + -oz (a)

1 o 1 o
+ pg, + [ z-;"
r ar
(r1Trr) + - .- "IJ (TIIrsiniJ)
r sm 0 "
(a)

op
-- (OT.. OT.. +-
0Tz)
oy + pg. + -ox +-
oy oz (b)

IJ-Direction
op (-
oT..+oT.. o.,: )
--+pg.+
oz ox oy- +oz- (e)

5.1-2 (b)
OF MOTJON IN CYLINDRICAL
(r, IJ, z)

f>-Direction

oz) =- t
1 o
--(TTrr
) +--
1 OTer
- T-98 +-
OT"] (a)
[ r or r oiJ r oz
(e)

(b)
In the next section, the viscous stress tensor "' will be represented in terms
of the rate of strain tensor d for a Newtonian fluid. The relationship is given
by
op 'f = 2,ud + [(K - i,u)V v]l (5.1-lSa)
- oz
or
(e)
T = 2p d + [ (K-~ ,U) (~~J ~i J (5.1-18b)
Sec. 5.2 Newton's Law of Viscosity
134 The Differential Equations of Motion Chap. S

where we refer to .t and K properly as the shear coefficient of viscosity and the in this section will be difficult to unders
bu/k coefficient of viscosity, respectively. In practice, we refer to .t and K as student. A mathematical description is
the "viscosity" and the "bulk viscosity," respectively. The rate of strain applications and a qualitative description
tensor is given in terms of the velocity gradients ast
The rate of strain
d; = !(Jv, + OV) (5.1-19)
2 OX ox, In order to determine the rate of str
A detailed derivation of this result is presented in Sec. 5.2 followed by a
consider a materialline element of length
qualitative discussion of the rate of strain in Sec. 5.3. The effort required to
trated in Fig. 5.2-1, P and Q are materia
understand the development is large compared to the benefits reaped from
such study; therefore, many students may wish to skip these sections and go z
directly on to Sec. 5.4. They may do this without detriment to the study of
subsequent material; however, any student with a good background in solid
mechanics can cover this material very quickly and should do so.

5.2 Newton's Law of Viscosity


Newton's law of viscosity is based on a linear relationship between the
viscous stress and the rate of strain. In Chap. 1, Newton's law of viscosity
for a simple one-dimensional flow was introduced in the form of a single
r
component of the stress tensor being equal to the coefficient of shear viscosity
times a velocity gradient,
(5.2-1)

Although this expression may appeal to the student's intuition, it is not at all
obvious t~at dv.,fdy is a measure of the rate of strain for a fluid in plane shear
flow, ev_9) though the units (sec-1) are satisfactory. Gaining sorne understand-
ing ofthe rate of strain and its relationship to the viscous stress is undoubtedly
one of the most difficult problems in fluid mechanics; in fact, more than
X
100 years elapsed between Newton's statement in 1686 that,
Flg. 5.2-1. Materi
"The resistance arising from want of lubricity in the parts of a fluid is,
other things being equal, proportional to the velocity with which the
element connecting them. If the fluid
parts of the fluid are separated from one another."
element would be translated and rotated
and the derivation of the Navier-Stokes equations by Navier (1821), Poisson However, if the line element were length
(1831), and Stokes (1845). 1 lt is only natural, then, that the ideas discussed deformed and the rate of change of len
t This relationship should be reminiscent of that between the strain tensor and gradients rate of deformation.
of the displacement vector encountered in solid mechanics. The scalar components of W e note that at any point in space,
11 in rectangular, cylindrical, and spherical coordinates are listed in Table 5.2-1, and the of line elements, each having a differ
relationship between the scalar components of the viscous stress tensor and the velocity tangent vector to the line element
gradients are given in Tables 5.2-2, 5.2-3, and 5.2-4 for Newtonian fiuids.
l. P. F. Nemnyi, "The Main Concepts and Ideas ofFluid Dynamics in Their Historical A.,= li
As-+
Development," Archive for the History of Exact Sciences, 1962, 2:52.
Differential Equations of Motion Chap. S Sec. 5.2 Newton's Law of Viscosity 135

the shear coefficient of viscosity and the in this section will be difficult to understand, requiring extra effort by the
In practice, we refer to p, and K as student. A rnathernatical description is presented first, followed by sorne
," respectively. The rate of strain applications and a qualitative description in Sec. 5.3.
gradients ast

+ ovi) (5.1-19)
The rate of strain
i ox,
is presented in Sec. 5.2 followed by a In order to determine the rate of strain at sorne point in space, let us
in Sec. 5.3. The effort required to consider a rnaterialline element of length !J.s embedded in the fluid. As illus-
compared to the benefits reaped from trated in Fig. 5.2-1, P and Q are material points and !J.s is the rnaterialline
wish to skip these sections and go
z
this without detriment to the study of
with a good background in solid
quickly and should do so.

o
on a linear relationship between the
Chap. l, Newton's law of viscosity Material line
element, ~s
introduced in the form of a single
to the coefficient of shear viscosity

(5.2-1)

the student's intuition, it is not at all


rate of strain for a fluid~
n lane shear
austa1cto:rv. Gaining sorne nderstand-
to the viscous stress ndoubtedly
fluid mechanics; in fact, more than
l"u,;uu;Ju in 1686 that, X

Flg. 5.2-1. Materialline element.

element connecting them. If the fluid were rnoving as a solid body, the line
elernent would be translated and rotated, but the length would never change.
equations by Navier (1821), Poisson However, if the line element were lengthened or shortened, the fluid would be
then, that the ideas discussed deformed and the rate of change of length of !J.s would be a rneasure of the
that between the strain tensor and gradients rate of deformation.
mechanics. The scalar components of We note that at any point in space, we could visualize an infinite number
eoordiDtates are listed in Table 5.2-1, and the
of line elements, each having a different direction A1, where .A, is the unit
the viscous stress tensor and the velocity
5.2-4 for Newtonian fluids. tangent vector to the line elernent
Ideas of Fluid Dynamics in Their Historical
Sciences, 1962, 2:52.
A; = lim
4s-+O
(!J.'
!J.s
) (5.2-2)
The Differential Equations of Motion Chap. S S.C. 5.1 Newton's Law of Viscosity
136

In solid mechanics, the word "strain" denotes a change in length per unit Remembering that the material deriva ti
length. We will follow this nomenclature and call the rate of change in length we obtain
per unit length the "rate of strain." We remember that each line element
passing through a particular point in space may be undergoing a different rate
!(_1 )!!._ (A.f2) .
2 As Dt2
of strain, depending on its direction; therefore, we define the rate of strain
in the following manner: Now, t.v1 represents the change in th
and a Taylor series expansion of v, ab~
the rate of } . {1 D }
strain in the = hm - - (t..s) (5.2-3)
{). -direction
(~a
1 As -o t.s Dt
t.v.=
Following Fig. 5.2-1, we write
(5.2-4) .Here, we have dropped the higher ord
limit t.s-+- O. Noting that !:u1 = t.r1,
where we understand that this equation only holds in limit as t.s-+- O. We form
may rewrite Eq. 5.2-3 in terms of the square of t.s to obtain
l(l)D
- -2 -(t..s)=
2 t..s Dt
2

the ratein of
strain the} = hm - ( - 1 2) -D (t..s 2)}
. { (1) (5.2-5)
{A;-direction 2 t.s Dt
As-o Substitution of this result into Eq. 5.2
In the following steps, it will be understood that we intend to take the Iimit the rate of }
t.s- O. strain in the
{). -direction
=
By Eq. 5.2-4 we obtain 1

Note that our result here is very simih


!(_1
2
)!!.__ (t..s2) = !(_12){!!.__ (t.r . t.r)\
2
t.s Dt t..s Dt
2 ' ' J
(5.2-6) the stress in a continuum. In that cas
by a tensor Tii, and the stress acting
Differentiation of the product yields direction n1 was given by
j 1){!!.__ (t..r;) t.r; + t.r; !!.__ (t..r;)} T.... =
!(_1)!!.__ (t..s2) = !(-
2 t..s 2 Dt 2 t..s 2 Dt Dt
(5.2-7)
In this development, we find that the '
By symmetry, we may write the right-hand side as by the tensor ov,fox1, and the rate of S
jn the direction ).1 is given by Eq. 5.2-1
1(1)D
- - 2 - (t.s 2) = - 2 t.r;-
2 t.s Dt t.s Dt
1{
D (t..r;) } (5.2-8) The tensor ov,fox1 may be rearr
conjugatet ovifox, to obtain
Noting that
(5.2-9) OV = !(OV + ovi)
we write oxi 2 oxi OX;

.P_(t..r;) = Dr;(Q) _ Dr;(P) = t..(Dr;) (5.2-10) t It may be helpful to remember from the
Dt Dt Dt Dt that
tbe atraln In lbe}
{ A,rtirection
and Eq. 5.2-8 takes the form
where u, is the deplacement vector.
(5.2-11) t The transpose of a matrix is obtained b
this operation is applied toa tensor, the resu
Differential Equations of Motion Chap. S S.C. 5.2 Newton'a Law of Vlscoslty 137

" denotes a change in length per unit Remembering that the material deriva ti ve ofthe position vector is the velocity,
and call the rate of change in /ength we obtain
We remember that each line element
may be undergoing a different rate (5.2-12)
therefore, we define the rate of strain
Now, dv, represents the change in the velocity vector between Q and P,
and a Taylor series expansion of v, about the point P gives
lim {_!_ !!_ (~s)} (5.2-3)
&s~o ~s Dt
(av.)
dv, = - dx 1
OX
(5.2-13)

(5.2-4) .Here, we have dropped the higher order terms in anticipation of taking the
only holds in limit as ~s-+- O. We limit ~s-+- O. Noting that dx1 = dr1, we may rearrange Eq. 5.2-12 in the
square of ~s to obtain form
l (-1)!!_ (~s2) = (av,ox )(~r,) (~r
2 ds 2 Dt ~s ~s
1)
(5.2-14)
(5.2-5) 1

Substitution of this result into Eq. 5.2-5 and use of the limit ds-+- O yieldt
that we inte d to take the limit
{!~:a:t: ~~e} (av,
.l.-direction ox
)A.,A.
=
i
(5.2-15)

Note that our result here is very similar to that obtained when we analyzed
(5.2-6) the stress in a continuum. In that case, the "state of stress" was determined
by a tensor T,1, and the stress acting on a plane having a normal n, in the
direction n1 was given by
(5.2-16)
(5.2-7)
In this development, we find that the "state of rate of str.ain" is determined
side as by the tensor ov,jox;. and the rate of strain on a surface having a normal A.,
jn the direction A.1 is given by Eq. 5.2-15.
(5.2-8) The tensor ov,Jox1 may be rearranged by adding and subtracting its
conjugatet ov1Jox, to obtain

- r;(P) (5.2-9)
(5.2-17)

(5.2-10) t lt may be helpful to remember from the small deformation theory of solid mechanics
that
= ( iJu,)
tbe s.train in tbe}
{..1.-du:ecuon iJx ).).
1
1

where u, is the deplacement vector.


_!_{~' ~(D'')}
~s 2
Dt
'
(5.2-11) t The transpose of a matrix is obtained by interchanging the rows and columns. When
this operation is applied to a tensor, the result is called the conjugate tensor.
138 The Differential Equations of Motion Chap. S Sec. 5.2 Newton's Law of Viscosity

The first term on the right-hand side is called the rate of strain tensor, d;, Table 5.
and is symmetric. SCALAR CoMPONENTS OF THE

d; = !(OV + OV;) (5.2-18) Rectangular Coordinates (x, y, z)


2 OX; OX .
(d;; = d;;)
The components of d;; for rectangular, cylindrical, and spherical coordinates
are listed in Table 5.2-1. The second term in Eq. 5.2-17 is called the vorticity
tensor, 01, and is skew symmetric d., =d.

o .. = !(ov _ ov;) (5.2-19) av.


" 2 OX ; OX d.. = oz
(O;;= -0;;)
Cylindrical Coordinates (r, O, z)
We now wish to show that the vorticity tensor O;; makes no contribution to
the rate of strain. Later we shall show that the components of this tensor
represent the rotational motion of the fluid.
Returning to Eq. 5.2-15 and making use of the definitions given by Eqs.
5.2-18 and 5.2-19, we obtain 1 ove v,
des = -r -ao + -r ds. =
the rate of }
{
strain in the = d;;A;A; + O;;A;A; (5.2-20) av.
1.;-direction d.. = oz'
where the unit tangent vector A; is completely arbitrary. We can show the last
term on the right-hand si de of Eq. 5.2-20 to be identically zero by carrying out Spherical Coordinates (r, O, r/>)
the double summation and suitably regrouping the terms; however, the proof
is somewhat easier if we use the technique of relabeling the repeated or dummy
ndices. We first split the term into two parts,
dr 8 = d8 r = ~2 [r !_ (~) -
or r

(5.2-21) 1 ov8 v,

and in the second term on the right-hand si de we relabel i as j and j as i to


d88=--
r ao +-.
r

obtain
= (-
1 OV. V,
+-r + -r- '
V9 COt 0)
O;;A;A; = !O;; A; A; + !O;;A;A (5.2-22)
d
r sin-O--
a,,
Changing the order of multiplication by A; and A; allows us to express this
resultas
(5.2-23) Newton's law of vlscosity
Because O;; is skew symmetric, we know that (O;; + O;;) is identically zero
and the proof is obtained. We begin our task of formulatin
O;;A;A; = O (5.2-24) by assuming that the viscous stress te
tensor. We then restrict ourselves to
Our expression for the rate of strain is now given entirely in terms of the rate rate of strain to obtain Newton's la
of strain tensor similar procedure in treating solids-i.
the rate of } of the strain tensor, and then restrict
strain in the
{1.;-direction
= d;;A;A; (5.2-25)
pendence on the strain to obtain Hoo
Differential Equations of Motion Chap. 5 Sec. 5.2 Newton's Law of Viscosity 139

is called the rate of strain tensor, d; 1, Table 5.2-1


SCALAR CoMPONENTS OF THE RATE OF STRAIN TENSOR

i
+ av.)
'
ax;
(5.2-18) Rectangular Coordinates (x, y, z)

, cylindrical, and spherical coordinates


term in Eq. 5.2-17 is called the vorticity

(5.2-19) ov.
d.. = oz '
Cylindrical Coordinates (r, 8, z)
tensor n;; makes no contribution to
that the components of this tensor
fluid.
use of the definitions given by Eqs.
/\ 1 OVe v,
dee=--
r o8
+-,
r
(5.2-20) ov.
d.. = oz' d., = d.. = 21 (ov. ov,)
a,: + oz
,..,.,,.,.,.,.,," arbitrary. We can show the last
to be identically zero by carrying out Spherical Coordinates (r, 8, </>)
the terms; however, the proof
of relabeling the repeated or dummy dre = der = !2 [' .!_ (~) + !r ov,J
or r o8
parts,
(5.2-21) 1 OVe Vr
d88=--
r o8
+-.
r
si de we relabel i as j and j as i to

(5.2-22)
_(
d- 1
ov.p v, ve cot
----+-+--
,sin 8 o,, r r
8) '

(5.2-23) Newton's law of viscosity


that (.Q;; + il;;) is identically zero
We begin our task of formulating a constitutive equation for fluids
(5.2-24) by assuming that the viscous stress tensor is a function of the rate of strain
tensor. We then restrict ourselves to a linear, isotropic dependence on the
now given entirely in terms of the rate rate of strain to obtain Newton's law of viscosity. We generally follow a
similar procedure in treating solids-i.e., assume the stress tensor is a function
of the strain tensor, and then restrict the analysis to a linear, isotropic de-
(5.2-25)
pendence on the strain to obtain Hooke's law.
140 The Differential Equatlons of Motion Chap. 5 S.C. 5.2 Newton'1 Law of VIICOIIty

Our starting point is, then, of Eq. 5.2-6 subject to the restriction th
(5.2-26) .,.,1 =O when
Thus, we have a tensor function of a tensor as our first step in formulating a is given by the following six equations:
constitutive equation. This idea is undoubtedly new to the student, and it is
best to back up a bit to begin at the beginning. We shall consider first a
.,.,. = Cud,. + C1rtiw + C1ad +
sca/ar function of a scalar, .,." = C11d,. + c.adw + c..c~ +
p = f(T) (5.2-27) .,. = C11d,. + C1rtiw + Cud +
We have indicated here that the density is a function of the temperature. .,... = C,1d,. + C,adw + C,ad +
This function is easily plotted as p versus T; derivatives are readily identified
as the slope of this curve and integrals as the area under this curve. .,. = Cud,. + Caadn + C,ad., +
Next we shall consider a scalar function of a vector, .,... = C11d,. + C1 rtiw + C,ad +
(5.2-28) We see that the general linear relationshi
Buried in this array is the special form of
We have indicated here that the density is a function of the spatial coordinates,
in Sec. 1.5 for a simple shearing flow-i
X, x 1 , x 8 (or x, y, z). If p were independent of one of the coordinates, we
could plot it as a function of x, to obtain a surface in 3-space. Derivatives
are understood as the slopes of tangents to this surface and integrals as vol- "=p
umes under the surface. We might ask, "What of a vector function of a which is simply a special form of Eq. S.
vector?" described in Sec. 1.5 we can deduce that
(5.2-29)
c.. =
Certainly this idea is readily acceptable and simply states that each scalar We can now show that by restrictin
component of the velocity depends on the spatial coordinates. tropic, only 2 of the 36 coefficients in
v1 = / 1(X, x 1, x 8) or v. = v.(x, y, z) (5.2-30a) of isotropy for Eq. 5.2-26 is expressed

v1 = h(x, x 2, x 8) or v. = v.(x, y, z) (5.2-30b) .,.;1 =/,


where .,.;1 and d1 are the viscous stress
V8 = / 3(X, x1, x8) or v. = v.(x, y, z) (5.2-30c) rotated coordinate system x 1 - x~. Eq
Although we may feel quite comfortable with such function, we cannot draw relationship between .,.,1 and d11 is in
a "picture" of it easily, so we must rely on mathematical formalism. idea of isotropy is sometimes difficult
Now let us return to our tensor function of a tensor and state more further before we apply this restriction t
explicitly what this terminology means. In words, we could say that each Let us consider two observers locatec!
component of .,.,1 is a function of all the components of d11 Symbolically, respectively. Each observer is going ~
we may write cxperiment on the same body of fluid. ol
1tress tensor and the rate of strain tens~
Tu = fn (dw du, daa, du, d2B du) (5.2-3la)
will reduce to
"u= / 11(d11 , d22 , d 83, d12 , d18, d 81) etc. (5.2-Jlb)
and
Owing to the symmetry of both "H and dii, we have only six equations repre-
senting the depende.ce of "'~ on the six distinct values of d11 .,.:,. = e~ d:W, in the
Having established what we mean by a tensor function of a tensor, we Now, if the fluid is isotropic we ex
need only write down the linear form of this relationship and require it to be ayatem to obtain the same dependence
lsotropic to obtain Newton's Iaw of viscosity. The most general linear form t A ICalar zero in a tensor equation shoul
Differential Equations of Motion Chap. 5 S.C. 5.2 Newton'a Law of Vlscoslty 141

of Eq. 5.2-6 subject to the restriction thatt


(5.2-26) .,.,1 =O when d11 =O (5.2-32)
tensor as our first step in formulating a is given by the foUowing six equations:
new to the student, and it is
beginning. We shall consider first a
.,..,. = Cud.,. + C1rd., + C1ad + Cudn + C1r,d,. + C1,d.f/e (5.2-33a)
'~"n = Clld + C.rd., + C,d + CtAdn + C dv + C,.d"11 (5.2-33b)
'~" = C11d.,. + C1rd., + C,d + C"d" + C d,. + C,.d.
(5.2-27) (5.2-33c)
81
is a function of the temperature. .,... = C,ld.,. + C,sdr, + C.,d + C"d" + C.,d,. + C.,d. (5.2-33d)
T; derivatives are readily identifi.ed
as the area under this curve. .,.,. = C61d.. + C6adn + C6,d + Cudn + C d,. + C.,d.66 (5.2-33e)
.,... = C, 1d.,. + c,r14, + c,,d + c"d" + C86dv + c ..d.f/e (5.2-33f)
(5.2-28) We see that the general linear relationship gives rise to 36 unknown constants.
Buried in this array is the special form of Newton's law of viscosity discussed
is a function of the spatial coordinates,
in Sec. 1.5 for a simple shearing flow-i.e.,
eoc:nd,ent of one of the coordinates, we
a surface in 3-space. Derivatives dvf/e
.,..f/e=p dy (5.2-34)
to this surface and integrals as vol-
ask, "What of a vector function of a which is simply a special form of Eq. 5.2-33d. From experiments of the type

dxs) (5.2-29)
described in Sec. 1.5 we can deduce that
c.. = 2p (5.2-35)
and simply states that each scalar We can now show that by restricting this linear relationship to be iso-
the spatial coordinates. tropic, only 2 of the 36 coefficients in Eqs. 5.2-33 are distinct. The condition
or vf/e = vix, y, z) (5.2-30a) of isotropy for Eq. 5.2-26 is expressed as

or v11 = v11(x, y, z) (5.2-30b) '1": = f,(d~,) (5.2-36)


where .,.;1 and d;1 are the viscous stress and rate of strain tensors in sorne
or v. = v.(x, y, z) (5.2-30c) rotated coordinate system 1 -x x;. Equation 5.2-36 states that the functional
with such function, we cannot draw relationship between .,.,1 and d11 is independent of the coordinate system. The
on mathematical formalism. idea of isotropy is somctimes difficult to grasp, and we should explore it
function of a tensor and state mqre further before we apply this restriction to simplifying Eqs. 5.2-33.
In words, we could say that each x x;
Let us consider two observers located in the 1 and coordinate systems,
the components of d11 Symbolically, respectively. Each observer is going to perform a simple shearing flow
experiment on the samc body of fluid. Only a single component of thc viscoua
(5.2-31a) stress tensor and the rate of strain tensor will be nonzero; thus, Eqs. 5.2-33
wiU reduce to
etc. (5.2-31b) in thc x 1-coordinate system (5.2-37)
and
d11 , we have only six equations reprc-
distinct values of d11 .,.:,_ = e~ d:,_, in thc x;-coordinate system (5.2-38)
by a tensor function of a tensor, wc Now, if thc fluid is isotropic we expect thc observcr in the x;-coordinate
of this relationship and require it to be l)'ltem to obtain thc same dcpendencc of stress upon rate of strain as tbe
,Qi'~!liltv, Thc most general linear form t A ecaJar zero in a tensor equation ahould be read u the null temor.
142 The Differential Equations of Motion Chap. S Sec. 5.2 Newton's Law of Viscosity

observer in the x,-coordinate system. Thus, C~, = C", and Eq. 5.2-38 We now return to examine the first of Eq
beco mes that the C terms are invariant; thus,
(5.2-39) 5.2-33a becomes
which is justa special case of Eq. 5.2-36.
We are now in a position to examine Eqs. 5.2-33 for a series of coordinate
transformations, for which we will require that the Cii remain constant. The
a.
transformed values of -r,1 and 1 will be related to the original values by Substitution of the results from Eqs. 5.2
Eqs. 5.2-40
-r; = CX~r:CX'TIJ: (5.2-40a)
Z 1Z
1
a;i = Ol.i/r.Cltjlakl (5.2-40b)
1 1
1 /)( First, let us examine a coordinate Comparing this result with Eq. 5.2-33a,
1 /
1 / transformation consisting of a 180
1 /
1 // rotation about the z-axis as shown
1 // in Fig. 5.2-2. The direction cosines
1 1 ,"' cxm,. for this transformation are given
By examining the remaining Eqs. 5.2-33,
Y----------~----------Y also be zero, and the array of coefficien
byt
cx11 = -1 CX12 = 0 cx13 =O C11 C12 G
C1a
cx21 =O CX22 = -1 cx 28 =O
X c21 C22 c2a G
cxal =O cxa2 =O CXaa = 1
Fl. 5.1-l. Coordinate transformation. (5.2-41) Cal Ca2 Caa
c41 c,2 C,a (j
We may now apply these results in forming the double sum indicated in
Eqs. 5.2-40. For -r~ 1 , we write
o o o o
T~l = cxucxu-r~~:
o o o o
Noting that only cxw cx22, and cx38 are nonzero, we quickly find that Finding the other zeros in this array
T~1 = ex u CXuTn = -( -1 ) 2-r11 = T 11
is a straightforward but rather tedious
task, and we will only outline briefiy
Tbe remaining terms for -r;1 and a;1 are cals_ulated readily, and we find the remaining steps. For the rotaqon
1 shown in Fig. 5.2-3, the array is
-r.,., = -r.,., '
-r.,=-r.,ll
reduced further to
1
'T;II = 'TIIII 'TI/1 = -TIIS (5.2-42)
1
'T = -r -r' ., = --r.:x: Cu c12 c13 o o o
and c21 C22 C2a o o o
a~z = azz a~=a., Cal Ca2 Caa o o o
a;"= a"" a;.= -a". (5.2-43) o o o e" o o
a~.= a a~z = -a.z o o o o e55 o
t Remember that according to Eq. 5.2-40, the first index on Ot refers to the x/-coordinate o o o o o Css
system, and the second index refers to the x,-coordinate system. Thus, Otu is the cosine of
the angle between the x'-axis and the y-axis. (5.2-48)
Differential Equations of Motlon Chap. S Sec. 5.2 Newton's Law of Viscosity 143

C~, = Cu, and Eq. 5.2-38 We now return to examine the first of Eqs. 5.2-33. We have already decided
that the C,1 terms are invariant; thus, in the x;-coordinate system, Eq.
(5.2-39) 5.2-33a becomes

Eqs. 5.2-33 for a series of coordinate


that the C11 remain constant. The (5.2-44)
be related to the original values by Substitution of the results from Eqs. 5.2-42 and 5.2-43 gives
Eqs. 5.2-40
T = tXj~~:ClTic (5.2-40a)
d - Otj/t.Otjldkl (5.2-40b) (5.2-45)
First, let us examine a coordinate Comparing this result with Eq. 5.2-33a, we find that
transformation consisting of a 180
rotation about the z-axis as shown (5.2-46)
in Fig. 5.2-2. The direction cosines
By examining the remaining Eqs. 5.2-33, we find that other coefficients must
cx:mn for this transformation are given
also be zero, and the array of coefficients takes the forro
byt
cx:11 = -1 Cl12 =o Cl13 =o Cu cl2 cl3 Cu o o
<X21 =o Cl22 = -1 Cl23 =o
C21 C22 C2s c2, o o
<Xsl =o CX:s2 =O <Xss = 1
(5.2-41) c81 Cs2 Css Cu o o
(5.2-47)
forming the double sum indicated in
en e,2 e,s e" o o
o o o o e66 e56
o o o o e65 e66
nonzero, we quickly find that Finding the other zeros in this array
=(-1) T11 =Tu 2 is a straightforward but rather tedious
task, and we will only outline briefly z
cals.ulated readily, and we find the remaining steps. For the rota,qon
shown in Fig. 5.2-3, the array is
'
T:~: 11 =T:~: 11
reduced further to
T;. = -T!Iz (5.2-42)
T~:t: = -Tz:~:
en e12 el3 o o o
-7-~------y
C21 e22 e2s o o o 1
d:W = d:l:ll c31 es2 esa o o o 1
1
1
d;. =-di/. (5.2-43) o o o e,, o o 1
1
d;:~: = -d.z o o o o e55 o 1
1
1 ,
the first index on at refers to the x/-coordinate o o o o o e66 IZ

rx,...coordiilate system. Thus, at 11 is the cosine of


(5.2-48) Flg. 5.1-3. Coordinate transformation.
1+4 The Differential Equations of Motion Chap. 5 Sec. 5.2 Newton's Law of Viscosity

y'Jz z'1z As we noted previously, in Eq. 5.2-3,


1 the coefficient of shear viscosity, anc
1 1
1 1
1 1
expressed as
1 1
1 1 cl2 = (K - t/l) (5.2-53)
1 1
1
1 where K is caBed the coefficient of
1
1
1
1
.r bulk viscosity. Expressing Eqs. 5.2-52
. in index notation and incorporating
the coefficients of viscosity, we get
'Tii = 2.tdi; + ((K - f.t))dkki;
(5.2-54)
Fig. 5.2-4. Coordinate transformation. Fig. 5.2-5. Coordinate transformation. In terms of the velocity gradients,
this equation reduces to
Continuing this process, we utilize the rotation shown in Fig. 5.2-4 to obtain
Cn C12 C12 o o o 'T. = !l(avi + ov;)
" OX; OX;
C21
c21
C22
C2a
c2a
C22
o
o
o
o
o
o (5.2-49)
+ ((K - i.t)] (!::) ;;

o o o c44 o o (5.2-55)
o o o o Css o This result is listed in Tables 5.2-2, 3,
o o o o o c44 spherical coordinates.
Here, we have reduced the array to seven distinct coefficients. For the co- Table
ordinate transformation shown in Fig. 5.2-5, we obtain ScALAR CoMPONENTS
TENSOR FOR NE
Cn C12 C12 o o o
cl2 Cn C12 o o o Rectangular Coordinates (x, y, z)
o o o
(av.)
cl2 cl2 Cn (5.2-50)
o o o c44 o o .,. = 2!-l ox
o o o o eH o av.)
o o o o o c44 .,. = 2p (ay

We obtain our final simplification of the arra y by using the coordina te trans- av,)
formation shown in Fig. 5.2-6, which a11ows us to prove that .,.., = 21-l
(oz
Cn = C44 + C12 (5.2-51)
Our final form of Eqs. 5.2-33 is
.,. = .,. = .t(,
'Txx = C44dxx + C12(d.,., + dVV +d..) (5.2-52a)
7 vv = c44dvv + cl2(d.,., + dvv + d,,) (5.2-52b) .,.., = .,.,. = 1'(-
'Tzz = c44d + cl2(d.,., + dvv + d ) (5.2-52c)
'T XV = C44dXV (5.2-52d)
7 11z = Cudvz (5.2-52e)
a In thesc cquations,
'Tzx = c44d., (5.2-52f)
The Differential Equations of Motion Chap. S Sec. 5.2 Newton's Law of Viscosity 145

z'l z As we noted previously, in Eq. 5.2-35 the coefficient C44 is identified with
1
1 the coefficient of shear viscosity, and the coefficient C12 is traditionally
1
1 expressed as
1
1
1 C12 = (K - J.t)
(5.2-53) 1
1 ZtZ

y
1
1
.
i
where K is called the coefficient of
bulk viscosity. ExpressingEqs. 5.2-52
1
1
1
.. in index notation and incorporating 1
1
1
the coefficients of viscosity, we get 1
Ti; = 2.tdii + [(K - J.t)]dkkbii
1
1
1
(5.2-54)
Fig. 5.2-5. Coordinate transformation.
In terms of the velocity gradients,
this equation reduces to \
the rotation shown in Fig. 5.2-4 to obtain \

o o o 'T;; = .t (OV; + OV;) \


\
X \
. OX; OX; \
o o o \ 1

o o o + [(K _ i!J)] (ovk) <5;; \X

22
(5.2-49) oxk Fig. 5.2-6. Coordinate transformation.
c44 o o (5.2-55)
o Css o
This result is listed in Tables 5.2-2, 3, and 4 for rectangular, cylindrical, and
o o c44 spherical coordinates.
seven distinct coefficients. For the co-
Table 5.2-2
5.2-5, we obtain SCALAR COMPONENTS OF THE VISCOUS STRESS
o o o TENSOR FOR NEWTONIAN FLUIDS

o o Rectangular Coordinates (x, y, z)


o o o (5.2-50)
c44 o o Tzz =
(
OVz)
2, OX + ((K - fp,)V V] (a)
o Cu o
o o c44 -r_. =
( Jv) + [(K -
2, oy f,t)V v] (b)
the array by using the coordinate trans-
allows us to prove that Jv,)
2, oz + [(K - f.u)V v]
.,... =
( (e)
(5.2-51)
(d)

Cl2(dzz + dvv + d) (5.2-52a)


Cl2(dzz + dvv + d ) (5.2-52b) (e)

C12(dzz + dv + d )
11
(5.2-52c)
(5.2-52d) (f)

(5.2-52e)
a In these equations.
(5.2-52f)
146 The Differential Equations of Motlon Chap. S Sec. 5.3 Qualitative Description of the

Table 5.2-3
ScALAR CoMPONENTS OF THE VISCOUS STRESS 5.3 Qual itative Descri ption
TENSOR FOR NEWTONIAN FLUIDS
Strain
Cylindrical Coordinates4 (r, O, z)
While our treatment of Newton's
(a) the analysis presented in the previous
will be worthwhile to discuss the su
1 1ve
T 88 = 2p, - -
( r 10
+ v,)
- + [(K -
r
fp,)V V] (b)

Tu = 2p, (av.) + [(K -


)z fp,)V v] (e)

1 (ve)
Tre =Te = p, r- - 1 1v,J
+-- (d)
r [ 1r r r 10

1ve 1 1v,)
Te = Tze = 11- -
( 1z +--
r 10
(e)

(f)
Fig. 5.3-1. Rates of defon

In thesc cquatlons.

rate of strain for the three line elem(


components of the velocity vector fo
Table 5.2-4
SCALAR CoMPONENTS OF THE VISCOUS STRESS
TENSOR FOR NEWTONIAN FLUIDS v.,
Spherieal Coordinates (r, O,</>)
4

V11 =
(av.) + ((K -
Trr = 2p, a; fp,)V V] (a) For each of the three cases, we wish
equation,
1 1ve
= 2p, ( - - + -v,) + [(K - fp,)V V] (b)
T 88
r 10 r rate of strain} l
{
inthe = --
1v. A-direction ds(>..l
T = 2,t ( -r sm
1
. - --
O a,,
+-rVr + Ve-eot
r-
0) + [(K - fp,)V V] (e)
where the unit tangent vector A is gi
1 (ve) 1 1v,J
+ --
re = Te r = p, [ r-
(d)
T
1r -r r 10

(e)
CASE 1 : The unit tangent vector

(f)

1 a 1 a . 1 ao.,
In thcsc cquattons, V ' = ;1 a,. (r1r> + r sin 8 ati <e sm B) + r sin 8 "i-
Differential Equations of Motion Chap. S Sec. 5.3 Qualitative Description of the Rate of Strain 147

5.3 Qualitative Description of the Rate of


Strain

While our treatment of Newton's law of viscosity is essentially complete,


- i,u)V v] (a)
the analysis presented in the previous section has been rather abstract, and it
will be worthwhile to discuss the subject further. Let us now examine the
(b)

- i,u)V v] (e)

~) + ~ ov,] (d)
r r o8

~ ov,) (e)
r o8

ov,) (f)
oz
Fig. 5.3-1. Rates of deformation for plane Couette fiow.
1 ave a
+--+-
r ao az
rate of strain for the three line elements illustrated in Fig. 5.3-1. The scalar
components of the velocity vector for this flow are

v"' =u o h (K) (5.3-la)

V11 = Vz = 0 (5.3-1b)
(a)
For each of the three cases, we wish to compute the rate of strain using the
equation,
- i.u)V v] (b)
rate of strain} 1 D
{inthe =--(ds<>->)=(dA.)A (5.3-2)
ve eot 8)
- r - + [(K - i,u)V v] (e)
A.-direction ds<>-> Dt
where the unit tangent vector A is given by
1 ov,] (d)
r o8 A..= dr;
(5.3-3)
ds<>->
1 ove]
8 +---
) (e)
r sin 8 otf>
CASE 1 : The unit tangent vector has only one nonzero component

r;,(1)] (f)
dr.
1, i = 1
a . 1 a.~ A.; = -- = O, i = 2 (5.3-4)
T8 <o sm O) + r sin o "i- ds(l) {
o, i = 3
1-48 The Differential Equations of Motion Chap. 5 Sec. 5.3 Qualitative Description of the

and Eq. 5.3-2 reduces to Letting (} = O, TT/2 gives the rate of


tively. The maximum rate of strain
(5.3-5)
angle of 45; thus,

From Table 5.2-1 we may obtain an expression for d11 (listed as d.), and the .1 D (d S(a
--
rate of strain becomes ds 181 Dt
-
1 D
-(ds1u) =_:e= O
ov (5.3-6) If we compare this result with Eq.
ds(l) Dt ox acting on the y-surface in the x-direct
This result should be intuitively obvious. strain of a line element inclined at
CASE 2: The unit tangent vector again has a single nonzero component This situation is rather curious,
following discussion is rather quali
0, i= 1 that basis.
1
A1 = dr = 1, i= 2 (5.3-7)
dS(t.)
{0 . -3
' J-
and we obtain
1 D
- - (ds 121) = duAA 1 = du (5.3-8)
ds 111 Dt
From the results in Table 5.2-1, Eq. 5.3-8 becomes
1 D
- - (ds 121) =-=O
ov. (5.3-9)
ds121 Dt oy Tren!
Once again, the result should be intuitively obvious, for although the line
clement ds121 is rotating, it is certainly not deforming.
CASE 3: The components of the unit tangent vector for this case are

cos 6, i= 1
dr1 . . (5.3-10)
A(=;--=
1

{
8
Sin' J= 2
s<al O, i= 3
and the rate of strain is given by

1 D du cos1 8 d 11 cos 8 sin 8 + (5.3-11)


- - ( d s111) -=
ds11 1 Dt +du sin 8 cos 8
1
dn sin 8 + Rot
Flc. 5.3-2. S
Noting that d11 is symmetric and using the results of Table 5.2-1, we gct

(ov.,) (ov., ov) O . O + (ov) . Let us consider, now, the two


ox cos + -oy + -ox cos sm
1
.- 1 -D (d s111) = 1 (}
- O
ds 111 Dt
-
oy sm 5.3-2, of a small fluid element whi
motions consist of pure translation
(5.3-12)
which immcdiately reduces to solid body motion. The element ma
shown in Fig. 5.3-3, or it may be s
- 1- !!._ (ds111) = (dvz) cos 8sin 8 (5.3-13) illustrated in Fig. 5.3-4. Note that
for an incompressiblc flow by requ
ds 111 Dt dy
Oifferential Equatlons of Motlon Chap. 5
Sec. 5.3 Qualitative Oescription of the Rate of Strain 149

Letting O = O, TT/2 gives the rate of strain (i.e., zero) for cases 1 and 2, respec-
(5.3-5) tive!y. The maximum rate of strain occurs for a Iine element situated atan
angle of 45; thus,

,1 D (d )j 1(dv.,) (5.3-14)
ds(s) Dt S(s) B=rr/4 = l dy
(5.3-6) If we compare this result with Eq. 5.2-1 or Eq. 1.5-7, we find that the stress
acting on the y-surface in the x-direction is directly proportional to the rate of
strain of a line element inclined at a 45 angle between the x- and y-axes.
has a single nonzero component This situation is rather curious, and requires further explanation. The
following discussion is rather qualitative in nature and must be accepted on
0, i = 1 that basis.
1, i = 2 (5.3-7)
{
o, i = 3
(5.3-8)

At

(5.3-9)
Tronslotion
obvious, for although the line
not deforming.
tangent vector for this case are

i =1
i=2 (5.3-10)
i=3

(5.3-11)
OcosO + d11 sin1 8 Rototion
Flg. 5.3-l. Solid-body motion.
the results of Table 5.2-1, we get

+ av") cos osin o+ (ov") sin 1


8 Let us consider, now, the two-dimensional motion, illustrated in Fig.
ax oy 5.3-2, of a small fluid element which does not undergo deformation. These
(5.3-12) motions consist of pure translation and rotation, and may be referred to as
salid body motion. The element may undergo extension (or compression), as
shown in Fig. 5.3-3, or it may be subjected to what is known as pure shear,
(5.3-13) illustrated in Fig. 5.3-4. Note that these illustrations indicate deformation
for an incompressible flow by requiring the volume of the fluid element to
Sec. 5.3 Qualltative Descriptlon of the
150 The Differential Equations of Motion Chap. S

We cannot visualize the rate of str


we are therefore forced either to e
coordinate transformation (rotatio
-------, Eq. 5.3-2. Plane shear is illustrate~
not analytically) that it is actually
----Txx
y y

'Tyy

t
Flg. 5.3

rotation. If we remember that vise


tion of the fluid and not from any s
Fig. 5.3-3. Extension. able that the shear stress T 11., for p
proportional to the rate of strain
remain constant. We may easily visualize that the pure shear shown in Fig.
5.3-4 could be obtained by superposition of the extensions shown in Fig.
5.3-3. For extension in the x-direction, the rate of strain depends on the Rotation
difference between the velocity at x = O and at x = !1x. Thus,
str~in . ) = r (/1v.,) (av.,) We can leam something about 1
{m~atetheofX-drrectlon Jm
l!.z-o uX
A
=
O X
(5.3-15) ing the time rate of change of the
define an average rotation in the e<
where !1v., represents the incremental increase in length per unit time. Simi-
larly, for the y-direction, we obtain M=

(~ate of str~in . ) =
m the y-d1rect10n
1' (!1v11 )
Im uy
A
l!.11-o
= (OV
Oy
11 )
(5.3-16)

Tyx

r Fig. 5.3-4. Pure shear.


tlf

Fig. 5.3-6. An
Sec. 5.3 Qualltative Description of the Rate of Strain 151
Differential Equations of Motion Chap. S

We cannot visualize the rate of strain in pure shear in this simple manner;
we are therefore forced either to examine Eq. 5.3-15 or Eq. 5.3-16 under a
coordinate transformation (rotation through 45), or return t the use of
Eq. 5.3-2. Plane shear is illustrated in Fig. 5.3-5, which shows (graphically,
not analytically) that it is actually composed of pure shear plus solid body

y y

X X

Flg. 5.3-5. Plane shear.

rotation. lf we remember that viscous stresses result only from\ the deforma-
tion of the fluid and not from any solid body motion, it should appear reason-
able that the shear stress -r11., for plane Couette flow (plane shear) is directly
that the pure shear shown in Fig. proportional to the rate of strain along the 45 line (pure shear).
of the extensions shown in Fig.
the rate of strain depends on the Rotatlon
and at x = !l.x. Thus,
We can learn something about the rotational motion offluids by examin-
(5.3-15)
ing the time rate of change of the angles ex and {3 shown in Fig. 5.3-6. We
define an average rotation in the counterclockwise direction by
in length per unit time. Simi-
f:l.() = !({3 - ex) (5.3-17)

(5.3-16)

At

shear. Fig. 5.3-6. An arbitrary deformation.


The Differential Equations of Motion Chap. 5 S.C. 5.4 The Equatlons of Motion
152

and the average rate of rotation around the z-axis as components are the same as those
construct a vector from the coJmo,one
n. = lim
&e-+o
(A.())
A.t
the "axial vector" of that tensor. In
vorticity tensor yields the vorticity ve
For small angles (i.e., At- 0), (1.. and Pare given by of the tensor av;/ax
1 is a measure of

(5.3-18a) 5.4 The Equations of Motio

Our objective in this section is t<


(5.3-18b) Newtonian fluid to the viscous stres!
will be valid only for incompressi
Starting with Eq. 5.1-17,
and the average rate of rotation becomes

O

= !_(av
2 ax
11 _ av.,)
ay
(5.3-19a) p(:; +Y VY) =

In a similar manner, we can show that we make use of the constitutive eqll

(5.3-19b)
p(~ +Y Vv) = -Vp +A
and
n., = ~ (av. _ av 11 ) (5.3-19c)
2 oy az At this point, the analysis is restrict
notation,
These threeterms may be recognized as one-half the three scalar components
of the vorticity vector w, which is the curl of the velocity vector. av; avi) ap + pg
(ot + Vax
p - ;- = - -
axi
VXY=I(av. av'll) (av., av.) k(av1/ av.,) 1
- - - +J - - - + -
W=
ay az az ax ax - -
ay Considering only the divergence of
(5.3-20) Eq. 5.2-18 for d,1 and write
Examining the scalar components of the vorticity tensor,
ad; a [1(av a
2.t.-= 2.t.---+
o _ ~ (av 11 _ av.,) ~ev.,- av.) ox ax ax a
1 2 1
2 ax ay 2 oz ox 1

Here, we restrict the analysis to inc


~ev'll- ov.,) o _ ~ev _ av'll)
Sl=
2 axay 2 ay az a2
_v; _ =

- ~ev.,- av.)
2 az ax
~ev
2 ay
_avu)az o
ax; ax;
We obtain this result by changing t
(5.3-21) the continuity equation for incomp

we note that the diagonal elements are zero and only three of the six off-
Vv
diagonal elements are distinct. Except for the factor t. these three distinct
S.C. 5.4 The Equatlons of Motlon 153

components are the same as those making up the vorticity vector. We can
construct a vector from the components of a skew-symmetric tensor, called
the "axial vector" of that tensor. In this case, twice the axial vector of the
vorticity tensor yields the vorticity vector w; thus, the skew-symmetric part
of the tensor ov,j ox
1 is a measure of the rotational motion of the fluid.

(5.3-lSa) 5.4 The Equations of Motion

Our objective in this section is to apply the constitutive equation for a


(5.3-lSb) Newtonian fluid to the viscous stress equations of motion. Our final result
will be valid only for incompressible flows and constant viscosity fluids.
Starting with Eq. 5.1-17,

(5.3-19a)
p(~; + v Vv) = -Vp + pg +V~ (5.4-1)

we make use of the constitutive equation for a Newtonian fluid to obtain


_ av.) (5.3-19b)
ox p(~ +V. vv) = -Vp + pg +V. [2,ud +(K- i,u)(V. v) 1]
(5.4-2)
_ ovv) (5.3-19c)
az At this point, the analysis is restricted to constant viscosity, yielding, in index
one-half the three scalar components notation,
of the velocity vector.
ovi ovi) - -op + od ( o (ovk)
(ot + OX =
p - V- pg + 2.t- + K -
2-,U) -- (5.4-3)
ov., _ ov,) + k(ovv _ ov.,) OX OX 3 OX oxk
(az ax ax oy Considering only the divergence of the rate of strain tensor, we substitute
(5.3-20) Eq. 5.2-18 for di;
and write
vorticity tensor,

l(ov., _ ov.) od
ox
[l(OV
2,u-= 2,u---+-
OX 2 OX
o1) ] ov [ v,
=,u--+-- (o
ov1 ) ]
2 2
(5.4-4)
2 oz ox OX OX; OX; OX OX

Here, we restrict the analysis to incompressible flows and write


_ ! (av. _ ovv)
2 oy oz ~ _ i_(ov 1) _O (5.4-5)
OX; OX OX OX;
o
We obtain this result by changing the order of differentiation and noting that
(5.3-21) the continuity equation for incompressible flows reduces to
zero and only three of the six off-
for the factor t. these three distinct (5.4-6)
Sec. S..of The Equations of Motlon
154 The Differential Equations of Motion Chap. S

Table
Applying Eq. 5.4-5 and substituting Eq. 5.4-4 into Eq. 5.4-3, we have the
differential equations of motion for the incompressible ftow of a constant
viscosity, Newtonian fluid
x-Direction
avi avi) ap
P(-a + v; a- = --a + pg +p.-a
a2v a (5.4-7)
t X; X X; X;

If we carry out the double Sl,lmmation for the term a2 ax1ax1,we obtain
y-Direction
a2 a2 a2 a2
--=-+-+- (5.4-8)
ax;ax; ax~ ax~ ax~
This scalar operator is known as the Laplacian and designated in Gibbs
z-Direction
notation by V2 where
a2 a2 a2
V2 =VV=-+-+-
ax2 al az 2 (5.4-9)

In Gibbs notation, Eq. 5.4-7 becomes


Local Pressure force Viscous force Tabl
acccleration

per u1t volume per rnit volume THE EQuATIONS oF Mono


CYUNDRICAL C

p(~: +V. vv) = -Vp + pg +,u Vv 2


(5.4-10) r-Direction

t
Convective
acceleration
i
Body force
per unit vlume

We should note that this result is nothing more than a differential form
of the linear momentum principle-i.e., the time rate of change of momentum
(represented in terms of local and convective acceleration) is equal to the 6-Direction
body force and the surface force (represented in terms of both the pressure
and viscous forces). Equation 5.4-10 is often known as the Navier-Stokes
equation in honor of the two scientists who first developed this final form. It
will serve as a starting point for many of the subsequent developments and
applications. These equations are tabulated in Tables 5.4-1, 5.4-2, and 5.4-3
for rectangular cylindrical and spherical coordinates.
z-Direction

Moving reference frames

Inherent in the development of the equations of continuity and motion


was the idea of a fixed reference frame (or coordinate system) with respect op
=--+
oz
to which distances and velocities were measured. In the analysis of sorne
prob1ems, it is convenient to view the system as if one were an observer
Sec. 5.4 The Equations of Motion 155
Differential Equations of Motion Chap. S

Table 5.4-1
. 5.4-4 into Eq. 5.4-3, we have the THE EQUATIONS OF MOTION FOR CONSTANT p. AND p IN
incompressible flow of a constant RECTANGULAR CooRDINATES (X, y, z)

x-Direction
02V
+pgi+!-l~ (5.4-7)
UX; UX;

for the term o2fox; ox1, we obtain


y-Direction

(5.4-8)
(b)

Laplacian and designated in Gibbs


z-Direction

(5.4-9) (e)

force Viscous force Table 5.4-2


unit volume per rnit
volume
THE EQUATIONS OF MOTION FOR CONSTANT p. AND p IN
t CYLINDRICAL CooRDINATES (r, IJ, z) .

(5.4-10) r-Direction

i
Body force
per unit vlume
P(
iJv, iJv, ve iJv, v~
at + v, Tr + 7 iJIJ - 7 + v. iJz
iJv,)

ap [ a( 1 a ) 1 av, 2 ove av,J


more than a differential form = - or + pg, + p. a; -a; (rv,) + ;: ao - ;: ao + oz (a)
the time rate of change of momentum
acceleration) is equal to the IJ-Direction
in terms of both the pressure
often known as the Navier-Stokes
first developed this final form. It
of the subsequent developments and
in Tables 5.4-1, 5.4-2, and 5.4-3 (b)

z-Direction

equations of continuity and motion


(or coordinate system) with respect
= -
op
iJz
[1 a( av.) 1
av. av.]
+ pg. +P. ; a,. 'Tr" + ;:t 001 + oz (e)
measured. In the analysis of sorne
system as if one were an observer
156 The Dlfferentlal Equatlons of Motlon Chap. 5
Sec. 5.4 The Equations of Motion

Table 5.4-3
THI EQUATIONS OF MonoN CoNSTANT s AND p IN Inasmuch as w is a constant vector,
SPHERICAL CooRDINATES (r, O, ~)t
op
-+wVp
r-Direction at
Referring to Eq. 3.2-20, we note that
total derivati ve; thus,
dp +V
(a) dt
and we see that the continuity equaf
0-Direction
same form as that for a fixed referen
We We Ve We
P( -+v-+--+---+-
V. iJvl V,.Ve V. COt
- -r-
0) we require only that the fluid vel
iJt r iJr r iJO ,. sin a~ r o Velocity V r and the time derivative
ojot be replaced with the time deriv
(b) ing frame constant djdt. For an obs
time rate of change of the density
however, for an observer moving at
~Direction
density at a fixed point is given by
to describe the continuity equation
that Eq. 5.4-16
op +~
(e) at
applies to both fixed reference frame
t For apberlcal coordbWea tbe Laplaclan la,
constant velocity. For moving fra
v 1 a (" a) 1 a1 in a_\ 1 (a' )
evaluated in terms of an observer m
- ; a, a, + 'In 8 j6\1 8 afJ} + ,. ain1 8 a~
If we replace v by vr + w in the

moving with a constant velocity w rather than fixed in space. If an observer


is moving with a velocity w, then the relative fluid velocity " that he measures
is
'fr = ' f - W (5.4-11) Since w is constant, this equation re

As an example, consider an observer situated on a ship moving through a


quiescent ocean. The fluid velocity v is zero and the relative velocity is just
equal and opposite to the velocity of the observer.
Once again the first two terms on ti
"r = -w (5.4-12) the total derivative, and we write
If we substitute Tr + w for the fluid velocity in the continuity equation, we
obtain p ( -dv. + 'fr VV
dt'
op op
-ot + V (pv) = -ot + V (pvr ) + V (pw) = O (5.4-13) This result indicates that the equati
observer moving at a constant vel
Oifferentlal Equatlons of Motlon Chap. 5
Sec. 5.4 The Equations of Motion 157

Inasmuch as w is a constant vector, we may write this equation as

op + w V p +V (pv) =O (5.4-14)
ot r

Referring to Eq. 3.2-20, we note that the first two terms ofEq. 5.4-14 are the
total derivative; thus,
2 av,
; a -
2 v,r'cot 6- 2
~
iJv)
a- (a)
dp
dt
+ V (pv) =
r
O (5.4-15)

and we see that the continuity equation for a moving reference frame has the
same form as that for a fixed reference frame. In changing reference frames,
_v.cot6) we require only that the fluid velocity v be replaced by the relative fluid
velocity vr and the time derivative holding the spatial coordina tes constant
r r
ofot be replaced with the time derivative holding the coordinates in the mov-
~ av, - _v,_ - 2 cos 6 iJv) ing frame constant dfdt. For an observer fixed in space, dpfdt represents the
+ r1 iJ6 r1 sin' 6 r 1 sin' 6iJ~ (b)
time rate of change of the density at a point moving with a velocity w;
however, for an observer moving ata velocity w the time rate of change ofthe
density ata fixed point is given by dpfdt. In general, Eq. 5:4-15 is not used
to describe the continuity equation for a moving observer. We simply state
that Eq. 5.4-16
op
-+V (pv) =O (5.4-16)
(e) at
applies to both fixed reference frames, and reference frames which move at a
constant velocity. For moving frames, velocities and time derivatives are
evaluated in terms of an observer moving with the frame.
If we replace V by Vr +"in the Navier-Stokes equations, we obtain

than fixed in space. If an observer


fluid velocity Tr that he measures
p[~t (vr + 2
w) + (vr + w) V(vr + w)J = -Vp + pg + p V (vr + w)
(5.4-17)
(5.4-11) Since w is constant, this equation reduces to

~ltllatc:<l
on a ship moving through a
zero and the relative velocity is just p(~r + W Vvr + Vr Vvr) = -Vp + pg + p V2vr (5.4-18)
observer.
Once again the first two terms on the left-hand side of Eq. 5.4-18 represent
-w (5.4-12) the total derivative, and we write
~locity in the continuity equation, we
(5.4-19)

' (pvr) + V (pw) = O (5.4-13) This result indicates that the equations of motion take the same form for an
observer moving at a constant velocity as they do for an observer fixed in
S.C. 5.5 Dimensional Analysls
158 The Differential Equations of Motion Chap. S

The first two terms on the right-hand


space. The situation clearly becomes more complex for accelerating reference as
frames, and that case is left for more advanced study. -Vp+pg=

*5.5 Dimen~ional Analysis


where p 0 is a constant generally b
Before going on to the solution of the equations of motion for uniformly rearrangement is possible only if the
accelerated flows and one-dimensionallaminar flows, we shall investigate the sible flows. Substitution of Eq. 5.5-1
dimensionless form of these equations. Each term in Eq. 5.4-10 has units of
either force per unit volume or time rate of change of momentum per unit vol-
ume, both being equivalent by definition. By dividing each term in Eq. 5.4-1 O
p(!: + vVv) =-V
by a constant having these same dimensions, we will obtain the dimensionless For any given flow there may be seve
equations that will give rise to two important dimensionless numbers-the length and velocity. In closed-con
Reynolds number, NRe and the Froude number, NFr In forming the dimen- traditionally chosen to be four time
sionless equations we shall make use of sorne characteristic length, L, and acteristic velocity is taken to be t
sorne characteristic time, Lfu0 , where U0 is the characteristic velocity. radius is defined as the cross-section
. In a sense, we can look upon this process as one of choosing a special set of
fundamental units (L instead of 1 cm and Lfu0 instead of 1 sec) for each ero!
R~~.=
particular problem. Such a choice leads to the scale factors, NRe and NFr wc
The main purpose of dimensional analysis is to aid us in interpreting For a pipe, the hydraulic radius is e
the experiments often performed to gain information about flows not sus-
ceptible to analysis. For incompressible flows, the vetocity and pressure are
determined by:

l. the equations of motion;


2. the continuity equation;
3. the boundary conditions.

From these requirements follows the guiding principie behind dimensional For the present we will simpl~
analysis, which states that the dependent variables (v and p) are functions of: velocity by L and u0, and multiply

l. the independent variables (x, y, z, and t);


2. the parameters that occur in the differential equations (p, fl, and g);
3. any parameters that occur in the boundary conditions.
a(;J + (.!.) .(LV)(.!.) = -e
Oe~O) Uo Uo
By putting the equations in dimensionless form, we can reduce the number
of parameters in the differential equations from three to one, and show that
the Froude number is important only for two-phase flows or free-surface Each term in parentheses in Eq. 5
flows. cumbersome equation, we need t
In analyzing Eq. 5.4-10, we first represent the gravity vector as the defining these variables, it is helpful
gradient of a scalar, cp the quantity in question. Insofar l
g =-Ve/> (5.5-1) using the capital letter or script le
sional quantity. Often, such a cho
where cp = -(g.,x + g y + g.z)
11
Sec:. 5.5 Dimensional Analysls 159
Differential Equations of Motion Chap. S

The first two terms on the right-hand side ofEq. 5.4-10 may now be expressed
complex for accelerating reference as
LUYQU\'-'U study.
-Vp+pg=-Vp-pV4>
= -V(p + P4>) (5.5-2)
= -V[(p- Po)+ P4>l

where p 0 is a constant generally taken to be atmospheric pressure. This


equations of motion for uniformly rearrangement is possible only if the density is constant-i.e., for incompres-
taLwu'" flows, we shall investigate the sible flows. Substitution of Eq. 5.5-2 into Eq. 5.4-10 gives
Each term in Eq. 5.4-10 has units of
of change of momentum per unit vol- p(~: + v Vv) = -V[(p- Po)+ p4>] +,u V2v (5.5-3)
By dividing each term in Eq. 5.4-10
we will obtain the dimensionless For any given flow there may be several possible choices for the characteristic
1nnrt"'" dimensionless numbers-the length and velocity. In closed-conduit flow, the characteristic length is
number, N Fr. In forming the dimen- traditionally chosen to be four times the hydraulic radius, R 11 , and the char-
some characteristic length, L, and acteristic velocity is taken to be the average velocity, (vs> The hydraulic
u0 is the characteristic velocity. radius is defined as the cross-sectional area divided by the wetted perimeter.
as one of choosing a special set of
and Lfu0 instead of 1 sec) for each R = cross-sectional area
11 (5.5-4)
wetted perimeter
to the scale factors, NRe and NFr
analysis is to aid us in interpreting For a pipe, the hydraulic radius is equal to one-fourth the diameter,
information about flows not sus-
flows, the vdocity and pressure are (1T~2)
R,.=--- (5.5-5)
1TD
D
4

~~'>"''~..E>
principie behind dimensional For the present we will simply designate the cbaracteristic length and
variables (v and p) are functions of: velocity by L and u0 , and multiply every term in Eq. 5.5-3 by L/ pu~ to obtain

and t);
differential equations (p, ,u, and g);
boundary conditions.
o(;J + (..!.) (LV)(.!.)
)r~o) Uo Uo
= -(LV)(p- {
PUo
0
+ 1> + (_l!:_)(r,2V )(..!.)
2
Uo
)
puoL
2

u0
form, we can reduce the number
(5.5-6)
from three to one, and show that
for two-phase flows or free-surface Each term in parentheses in Eq. 5.5-6 is dimensionless; to work with a less
cumbersome equation, we need to define sorne dimensionless variables. In
represent the gravity vector as the defining these variables, it is helpful to choose symbols which clearly represent
the quantity in question. Insofar as it is possible, we shall try to do so by
(5.5-1) using the capital letter or script letter associated with the particular dimen-
sional quantity. Often, such a choice would lead to confusion, as it would if
160 The Differential Equations of Motion Chap. 5 Sec. 5.5 Dimensional Analysis

V were used to represent the dimensionless velocity, or if T were used to


represent the dimensionless time. Under these circumstances, sorne other
variable is chosen to represent the dimensionless variable. The dimensionless
variables in Eq. 5.5-6 are defined as

U = (:J {dlmenslonleas veloclty}

0 = c~o) {dimensionleastime}

V = LV {dlmensionleas vec:tor operator}t

p = (p - Po)
pu~
{dlmensionleas preasure}

N _ PUoL {dlmensionleas parameter called the}


Re - -- Reynolds number
p.
[JJ =p + 1!._2 {dimensionleas
the body
preasure whlch lncludes}
force term
Uo
Use of these dimensionless variables allows us to write Eq. 5.5-6 as
Flg. 5.5-1. Sudden
1
oU +U. VU =-V&'+ - - V2U (5.5-7)
o0 NRe
sudden contraction. For practica]
For incompressible flows, the dimensionless continuity equation takes the the mathematical point of view w
form parabolic at z = oo.
(5.5-8) While it is possible to solve tn
So far .we have reduced the number of parameters in the differential thus determine the pressure variati
equations from three (p., p, g) to one (NRe); however, both U and &' will sive to investigate this effect experiJJ
depend on the parameters that appear in the boundary conditions, and we pressure variation. To obtain the n
need to examine two specific cases to indicate how the boundary conditions mnimum amount of experimenta
are treated in dimensional analysis. From these two examples, we shall see dimensionless form. The differe
that the Froude number enters the solution as a parameter in the normal stress dimensionless form and we must J
condition at a fluid-fluid interface. For confined flows, the dimensionless form. The boundary conditions a
velocity and pressure, U and &', are independent of the Froude number.
B.C.l: v, = 2(v,){l- 4
Sudden contraction in a pipeline B.C.2: v. =o, r = T,
As an example of flow in a closed conduit, we consider the sudden B.C. 3: v. =O,
contraction in a pipeline such as that illustrated in Fig. 5.5-l. We assume
(~~
2

that the Reynolds number is less than 2100; the flow is therefore laminar, B.C. 4: v, = 2(v,)I
and the velocity profile is essentially parabolic at sorne distance from the
t Numerical methods have been use
t Lack of a suitable symbol for the dimensionless "del" operator compels us to use found in J. E. Fromm, "A Method for
thc samc symbol for both the dimcnsionless and thc dimensional form. Flows" (Los AJamos, Calif.: Los Alam
Differential Equations of Motion Chap. S Sec. 5.5 Dimensional Analysis 161

velocity, or if T were used to


these circumstances, sorne other
~~~vUJll;;~~ variable. The dimensionless

~+oo

Grovity

Flg. 5.5-1. Sudden contraction in a pipeline.


(5.5-7)
sudden contraction. For practica! purposes this distance is finite, but from
continuity equation takes the the mathematical point of view we simply state that the velocity profile is
parabolic at z = oo.
(5.5-8) While it is possible to solve the equations of motion for this flowt and
of parameters in the differential thus determine the pressure variation along the z-axis, it may be less expen-
(NRe); however, both U and 9 will sive to investigate this effect experimentally and develop a correlation for this
in the boundary conditions, and we pressure variation. To obtain the maximum amount ofinformation from the
how the boundary conditions mnimum amount of experimental effort, we must examine the problem in
these two examples, we shall see dimensionless form. The differential equations are already available in
as a parameter in the normal stress dimensionless form and we must now put the boundary conditions in this
confined flows, the dimensionless form. The boundary conditions are
independent of the Froude number.
B.C.1: v, = 2(v.>{ 1 - 4(~J} z= -00 (5.5-9a)

B.C. 2: v. =o, r =r, O>z:::::-oo (5.5-9b)


B.C. 3: v, =O, r = r2 , +oo:::::z;;:::O (5.5-9c)
conduit, we consider the sudden

(~~T1- 4(~JJ
illustrated in Fig. 5.5-1. We assume
2100; the fiow is therefore laminar, B.C.4: v, = 2(v.) z = +oo (5.5-9d)
uaauJll'-' at sorne distance from the
t Numerical methods have been used with success on such problems. Details may be
nsi<onless "del" operator compels us to use found in J. E. Fromm, "A Method for Computing Nonsteady, Incompressible, Viscous
and the dimensional form. Flows" (Los Alamos, Calif.: Los Alamos Scientific Laboratory, 1963), LA-2910.
162 The Differential Equations of Motion Chap. S
Sec. 5.5 Dimensional Analysis

where we ha ve made use of conservation of mass (or vo1ume in this case of


incompres~ib1e fl.ow) For the configuration shown in Fig. ~

1r D~ (v.)1 = D~ (v. ) 2
1r (5.5-10) <4>) = -
4 4 and the pressure becomes
in formu1ating boundary condition 4. Choosing (v, )1 as the characteristic
ve1ocity :md D 1 as the characteristic 1ength 1eads to the following dimension-
less boundary conditions, Here we see that gravity simply gives
B.C. 1': U,= 2(1- 4R2 ), Z = -oo (5.5-11a) term, pgz cos Ot, and does not infl.uen
B.C. 2': U, = O, R = t, O> Z - oo (5.5-11b) wou1d seek on1y to determine the fu

B.C. 3': u. = o, R = t&lD 00 z o (5.5-11c)


Racing sloop hull design
B.C. 4': Z= +oo (5.5-lld)
where To illustrate how the Froude nu1
1et us assume that we are assigned t
R = ..!._ hull shape for an America's Cup defe
DI
s1oop is a subt1e and comp1ex task gen
Z=.!.... sailor-designers; however, sorne use
DI obtained by towing models through <
Fig. 5.5-2. The model is a scaled-do
lD= D2t
DI dimension-such as the 1ength, beam.
On the basis ofEqs. 5.5-7, 5.5-8, and 5.5-11, we may express the functiona1
dependence of U and & as
U= U(R, (), Z, 0, NRe lD) (5.5-12)
f!P = f!P(R, (), Z, 0, NRe lD) (5.5-13)
Confining our attention to the pressure term, we express the time average
and area average ast
(5.5-14)
where the dependence on R, (), and 0 is removed by the averaging process.
This resu1t indicates that the dimension1ess pressure depends on Z and the
two parameters, NRe and lD. Note that one ofthese parameters carne from
the differentia1 equation, whi1e the second entered the solution via a boundary
condition. Position of the free
Actual experiments will be concerned with the dimensional pressure p, =
surface is z, z, (x, y)
which may be expressed as
(5.5-15)
t A script 9t with an appropriate subscript will be used throughout the text to denote
a dimensionless ratio of two like quantities.
t Time averages will be discussed in Chap. 6 and will always be denoted with the
overbar ().
Fig. 5.5-2. M
Chap. S
Sec. 5.5 Dimensional Analysis 163
of mass (or volume in this case of
For the configuration shown in Fig. 5.5-1 , the scala r (c/J ) is given by

(5.5-10) (cjJ ) = -gz sinn: (5 .5-16)


and the pressure becomes
. Choosing (v.) 1 as the characteristic
leads to the following dimension- (p) = p(z)U(Z, NRe !J?n) + pgz sinn: + Po (5.5-17)

Here we see that gravity simply gives rise to an additive hydrostatic pressure
(5.5-11 a) term, pgz cos ex, and does not influence the velocity U. An experimental study
(5.5-11b) would seek only to determine the function f(Z, N Re !Ji n).
(5.5-11c)
Racing sloop hull design
Z = + oo (5.5-lld)
To illustrate how the Froude number enters into free surface problems,
Jet us assume that we are assigned the job of determining the best possible
hull shape for an America's Cup defender. The design of a hull for a racing
sloop is a subtle and complex task generally undertaken by wise and venerable
sailor-designers; however, sorne useful information about hull design can be
obtained by towing models through a testing tank, such as that illustrated in
Fig. 5.5-2. The model is a scaled-down version of a proposed hull, and every
dimension- such as the Jength, beam, etc.-is related to the dimensions ofthe
5.5-11, we may express the functional
Towing mechonism

(5.5-12)
(5.5-13)
term, we express the time average

(5.5-14)
is removed by the averaging process.
ess pressure depends on Z and the
one of these parameters carne from
entered the solution via a boundary Model hull

Position of the free


with the dimensional pressure p, =
surfoce is Z5 z5 (x, y)

z
(5.5-15)
will be used throughout the text to denote

6 and will always be denoted with the


L.
Fig. S.S-2. Model towing tank.
Sec. 5.5 Dimensional Analysis
164 The Differential Equations of Motion Chap. S

The Froude number is defined ast


actual hull by
(5.5-18)
where ~ L < l. Thus, the model is geometrically similar to the actual hull.
Our experiment rnight consist of towing severa! models, all having the thus yielding the final form of boundar
same scale factor ~ L> through the tank to determine which produces the z
least drag ata given speed. Presumably we know approximately at what speed B.C. 3": f!JJ = - ,
NFr
the actual hull will be moving, and we know the physical properties of the
fluid (salt water) through which it will move. The question to be answered is: On the basis of Eqs. 5.5-7 and 5.5-8 a
What should be the conditions in the tank so that the tests are dynamically Eqs. 5.5-20a and b and Eq. 5.5-24, the
similar to the actual motion? The answer is that the Reynolds number and is represented by
Froude number must be the same; our objec.tive is to prove it. U= U[X, Y,Z, 0, A1
lt is easiest to view this problem as if the hull were fixed in space and the
fluid were flowing past it with a uniform velocity. Under these conditions f!JJ = f!JJ[X, Y, Z, 0, A
the boundary conditions are Since the models are geometrically si
B.C. 1: x,y, z-.. oo (5.5-19a) function A( X, Y, Z) is identical for bot
flow to be dynarnically similar, we req
(i.e., the flow is undisturbed far from the hull, the origin of the coordinate Froude number be the same; thus,
system being on the hull), and
B.C. 2 : V= 0, a(x,y, z) =O (5.5-19b) _
N Re- (u L)
0
11 actUI
Here, a(x, y, z) is a function which describes the surface of the hull; thus, and
boundary condition 2 indicates that the tluid velocity is zero on the surface 2

of the hull. NF - u )
--.
r - ( gL actu

B.C. 3: p =po, z = z,(x,y) (5.5-19c)


When these conditions are subjected t
This last boundary condition indicates that the fluid pressure is equal to the we find
atmospheric pressure at the free surface, designated by z,(x, y). This boundary
condition neglects viscous and surface tension effects, but this is a reasonable and
assumption. Rewriting the boundary conditions in dimensionless form, we get 11model = 9i
B.C. 1': U = -i, X , Y, Z--.. oo (5.5-20a) The condition on the velocity is easil
kinematic viscosity is difficult if not iJ
B.C. 2': U = O, A( X, Y, Z) = O (5.5-20b)
cases. Since ~ L is necessarily much sm
B. C. 3': f!JJ = 4> 1 , Z = Z,(X, Y) (5.5-20c) of the fluid in the towing tank must b
u 02 z= z, (o:,!ll than the kinematic viscosity of water.
The function 4> is given by and in practice we can maintain eithe '
4> = gz (5.5-21) number constant but not both. Under
intuition are required to determine the
and to incorporate the dimensionless variable Z, we multiply and divide 4> ance of a hull.
by the characteristic length L , and rewrite boundary condition 3' as
t Most often the Froude number is defiii
Eq. 5.5-23 arises naturally from the dimensio
B.C. 3": f!JJ = (~~)z. Z = Z,(X, Y) (5.5-22) be used throughout this text.
Sec. 5.5 Dimensional Analysis 165
Differential Equations of Motion Chap. S

The Froude number is defined ast


u2
(5.5-18) NFr = _Q, (5.5-23)
gL
similar to the actual hull.
severa! models, al! having the thus yielding the final forro of boundary condition 3,
to determine which produces the
we know approximately at what speed B.C. 3" : Z = Z,(X, Y) (5.5-24)
know the physical properties of the
The question to be answered is: On the basis of Eqs. 5.5-7 and 5.5-8 and the boundary conditions given by
so that the tests are dynamically Eqs. 5.5-20a and b and Eq. 5.5-24, the functional dependence of U and &
is that the Reynolds number and is represented by
objec.tive is to prove it. U = U[X, Y, Z, e, A(X, Y, Z), NRe NFr] (5.5-25)
if the hull were fixed in space and the
velocity. Under these conditions & = &[X, Y, Z, e, A(X, Y, Z), NR e NFrl (5.5-26)
Since the models are geometrically similar to the proposed actual hull, the
(5.5-19a) function A(X, Y, Z) is identical for both. For the actual flow and the model
flow to be dynamically similar, we require that the Reynolds number and the
the hull, the origin of the coordinate Froude number be the same; thus,

(5.5-19b) N _
Re -
(u L)0
v actual -
_(u L\0
v }model
(5.5-27)
the surface of the hull; thus, and
uid velocity is zero on the surface
NF- -u~)
r - (- (u~)
-
gL actual-- gL model
(5.5-28)

(5.5-19c)
When these conditions are subjected to the restriction given by Eq. 5.5-18,
that the fluid pressure is equal to the we find
designated by z.(x, y). This boundary (5.5-29)
effects, but this is a reasonable and
3/2
vuuuoJu~ in dimensionless form, we get Vmodel = :71-L Vactual
/M)
(5.5-30)
(5.5-20a) The condition on the velocity is easily met; however, the condition on the
Y,Z) =O (5.5-20b) kinematic viscosity is difficult if not impossible to meet for most practica!
cases. Since ~Lis necessarily much smaller than one, the kinematic viscosity
Z = z.(X, Y) (5.5-20c) of the fluid in the towing tank must be one to two orders of magnitude Iess
than the kinematic viscosity of water. Such a fluid is rather difficult to find,
and in practice we can maintain either the Reynolds number or the Froude
gz (5.5-21) number constant but not both. U nder such conditions considerable skill and
intuition are required to determine the effect of a free surface on the perform-
Z, we multiply and divide tfo ance of a hull.
boundary condition 3' as
t Most often the Froude number is defined as u0 /VgL; however, the form given by
Eq. S.S-23 arises naturally from the dimensionless form of the equations of motion and will
Z = Z,(X, Y) (5.5-22) be used throughout this text.
166 The Differential Equations of Motion Chap. S
Se c. 5.6 Appl ications of the Differential E

The treatment of dimensional analysis given here is brief in comparison


For uniform acceleration,
to the usefulness of the method. A more thorough discussion is available
elsewhere. 2 - 3 Dv
-=a
Dt '
and
5.6 Applications of the Differential
Equations of Motion In light of Eqs. 5.6-4 and 5.6-5, Eq. 5

There are only four classes of problems for which we may easily obtain 0= -Vp
analytic solutions of the equations of motion: fluids at rest or moving at a for uniform acceleration. Equation
constant velocity; uniformly accelerated flows; one-dimensionallaminar flow; except that the gravity vector has bee
and irrotational flow. With the aid of advanced analytical or numerical three scalar forms of Eq. 5.6-6 by tald
methods, practically all laminar flows are susceptible to analysis; however, respectively,
these four types of flow represent the limits of the undergraduate at this time. op .
Although irrotational flow theory is a key tool in the analysis of flow - = pl
around immersed bodies, and wave propagation in open channel flow, it will
ox
not be treated in this text. A readable account of this particular area of fluid op .
-= PJ
mechanics is given by Temple, 4 although the classic in the field is still the oy
detailed treatment of Milne-Thomson. 6
lf the fluid velocity is constant, op
-=p
v=b (5 .6-1) oz
Noting that the total differential of th
where the constant vector bis zero if the fluid is at rest. Substitution of Eq.
5.6-1 into Eq. 5.4-10 yields the previously derived result,
dp = op dx +
O= -Vp + pg (5.6-2) ox
The solution of this equation was treated in detail in Chap. 2; we shall not we may use Eqs. 5.6-7 to obtain
discuss it further here. dp = p[i (g - a) dx +j
Integration of Eq. 5.6-9 between x 0 , Yo
Uniformly accelerated flow
the general expression for the pressur
Uniform acceleration means that each element of the fluid experiences p = Po + p [ A.,(x - x0)
the same acceleration; thus, v is a function of time but not position. Making
where p = p(x, y, z)
use of the material derivative, we may write Eq. 5.4-10 as
Po = p(xo, Yo Zo)
Dv ~-
p-=-Vp+pg+.tV-v (5.6-3)
Dt A.,= i (g- a)
2. G. Birkhoff, Hydrodynamics, A Study in Logic, Fact and Similitude (Princeton, N.J.: A 11 = j (g- a)
Princeton University Press, 1960).
3. P. W. Bridgman, Dimensional Analysis, (New Haven, Conn.: Yale University Press, Az =k (g- a)
1931).
4. G. Temple, An Introduction to Fluid Dynamics (New York : Oxford Press, 1958). If the pressure p 0 is known or specifie(
5. L. M. Milne-Thomson, Theoretica/ Hydrodynamics, 4th ed. (New York : The where in the system. Ifthe system con1
Macmillan Company, 1960). is constant, Eq. 5.6-10 may be used t(
Chap. S
Sec. 5.6 Applications of the Differential Equations of Motion 167
given here is brief in comparison
thorough discussion is available For uniform acceleration,
Dv
-=a a constant vector (5.6-4)
Dt '
and
V 2 v =O (5.6-5)
In Iight of Eqs. 5.6-4 and 5.6-5, Eq. 5.6-3 reduces to
for which we may easily obtain O= -Vp + p(g- a) (5.6-6)
motion: fluids at rest or moving at a
for uniform acceleration. Equation 5.6-6 has the same form as Eq. 5.6-2,
flows; one-dimensionallaminar flow; except that the gravity vector has been replaced by (g - a). We obtain the
of advanced analytical or numerical
three scalar forms of Eq. 5.6-6 by taking the scalar product with i, j, and k,
are susceptible to analysis; however, respective! y,
of the undergraduate at this time.
is a key too! in the analysis of flow op= p i. <g- a )
- (5.6-7a)
vv"l'."uvu in open channel flow, it will
ox
account of this particular area of fluid
-op = . <g- a)
PJ. (5.6-7b)
the classic in the field is still the ay
op = pk. (g - a) (5.6-7c)
(5.6-1) oz
the fluid is at rest. Substitution of Eq. Noting that the total differential of the pressure is
derived result,
dp = op dx + op dy + op dz (5.6-8)
(5.6-2) ax oy az
in detail in Chap. 2; we shall not we may use Eqs. 5.6-7 to obtain
dp = p[i (g- a) dx + j (g- a) dy +k (g- a) dz] (5.6-9)
Integration of Eq. 5.6-9 between x 0 , y 0 , z 0 and any arbitrary point x, y, z gives
the general expression for the pressure in a uniformly accelerated fluid.
each element of the fluid experiences
of time but not position. Making
p =Po+ p [ A.,(x- x 0 ) + Av(y- y + A.(z- z
0) 0) J (5.6-10)
write Eq. 5.4-10 as
where p = p(x, y, z)
Po = p(xo, Yo Zo)
+ pg + ,uV2v (5.6-3)
A.,= i (g- a)
in Logic, Fact and Similitude (Princeton, N.J.: A 11 = j (g- a)
(New Haven, Conn.: Yale University Press, A.= k (g- a)
Dynamics (New York: Oxford Press, 1958). If the pressure p 0 is known or specified, Eq. 5.6-10 defines the pressure every-
Hydrodynamics, 4th ed. (New York: The where in the system. If the system contains a free surface at which the pressure
is constant, Eq. 5.6-10 may be used to determine the position ofthe surface.
Sec. 5.6 Appllcatlons of the Oifferentlal Equ
168 The Differential Equations of Motion Chap. S

remains constant. Since the flow is incon


For example, let us specify that x 0 , y 0 , z0 is a point on the free surface where L L
the pressure is p 0 Now let z = z,(x, y) represent the position of the free
surface. Then h0L = f
o
z, dx = f[
o
Zo

p[x, y, z.(x, y)] =Po (5.6-11)


= z0L - ~(~
2

and Eq. 5.6-10 reduces to -


A.,(x - x 0 ) + A (y - y 0) + A.(z, - z0) =O (5.6-12) Solving for z0 gives
11
z0 = h 0 + ~g
Consider the problem of a rectangular tank, illustrated in Fig. 5.6-1;
which is accelerated in the x-direction. Initially the fluid in the tank would not and substitution of Eq. 5.6-17 into Eq ..
a.
Position of the free z, = ho+-
surfoce when the g
tonk is acceleroted
z One-dimensional laminar flow

L. As the first example of one-dimensiot


detail laminar flow in a pipe. We have
Sec. 2.8, where we applied the linear m

1
Grovity ..,__ _ _...,Accelerotion
volume to yield the equation of motion a
investigation. That method of analysis h!
~~~~~~~----~~--~~ a=l~ tion with a mnimum of effort, tending t<J
~---------- L -----~ stress distribution and the shear stress
Fig. 5.6-1. An accelerated rectangular tank. advantage is its tendency to gloss ove
must be made.
In this example, we shall start wit~
be uniformly accelerated, for it would slosh back and forth and the surface Stokes equations in cylindrical form and
would be wavy. However, after a period oftime this secondary motion would wc may simplify them to obtain the pre
die out and movement would be uniform. For this system, In starting with the complete form of
clearly the assumptions that go into the s
A.,= i ( -kg - ia.,) = -a., (5.6-13a)
of this method of analysis. We wish to ,
A11 = j (-kg- ia.,) =O (5.6-13b) illustrated in Fig. 5.6-2.
l. r-direction:
A, = k ( -kg - ia.,) = -g (5.6-13c)

and Eq. 5.6-12 becomes p


OVr OVr
-+v-+---- V9 OVr v: + v-OV,
( ot r or r ao r oz
-a.,(x - x 0)
Solving Eq. 5.6-14 for z, gives
- g(z, - z 0) = O (5.6-14)
+ pgr + f' [:re :r (rvr))
2. 0-direction:
z, = z0 - a., (x - x 0 ) (5.6-15)
g p ( -OVe + V rOV9
-
+ V9-OV9
-
+ VrVII
-
+ V, .
ot or r ao r
We see that the liquid depth is a linear function of x but the exact position of
the interface is not yet determined. To locate the interface, we must make use + pg, + f' [~(! ~ (rv11))
or r or
?f the principie of conservation of mass-i.e., the mass of the fluid in the tank
Sec. 5.6 Appllcations of the Differential Equatlons of Motion 169
Differential Equations of Motion Chap. S

remains constant. Since the fiow is incompressible, we require that


z0 is a point on the free surface where
L L
y) represent the position of the free
h0 L = Jz,dx = J[zo- '<x - x)J dx
o o
0

(5.6-11) (5.6-16)
L- ~(~ - x L)
2

= z0 0

Solving for z0 gives


(5.6-12)
tank, illustrated in Fig. 5.6-1,
z=
0 h0 + 7(i- Xo) (5.6-17)

Initially the fluid in the tank would not and substitution of Eq. 5.6-17 into Eq. 5.6-15 yields the final result

z, = h0 + ~ (~ - x) (5.6-18)

z One-dimenslonal laminar flow

L, As the first example of one-dimensionallaminar flow, we shall examine in


detail laminar fiow in a pipe. We have treated this problem previously in
Sec. 2.8, where we applied the linear momentum principie to a differential
volume to yield the equation of motion applicable to the particular flow \mder
1------+- Accelerotion investigation. That method of analysis has the advantage of providing a solu-
.....-.~ o =iax tion with a mnimum of effort, tending to focus student attention on the shear
stress distribution and the shear stress-rate of strain relationship. lts dis-
advantage is its tendency to gloss over sorne important assumptions that
must be made.
In this example, we shall start with the complete forro of the Navier-
slosh back and forth and the surface Stokes equations in cylindrical forro and demonstrate under what conditions
of time this secondary motion would we may simplify them to obtain the previously derived Hagen-Poiseuille law.
F or this system, In starting with the complete forro of the equations, we are forced to state
- ia.,) =-a., (5.6-13a) clearly the assumptions that go into the solution. This, then, is the advantage
of this method of analysis. We wish to apply the listed equations to the flow
- ia.,) =O (5.6-13b) illustrated in Fig. 5.6-2.
- ia.,) = -g (5.6-13c) l. r-direction:

p (OVr +V OVr + ~ OVr- ~+V OVr) = - op


ot r or r ao r oz or
g(z,- z0) =O (5.6-14) 2 2
a (1 o ) 1 0 Vr 2 OVe 0 Vr] (5.6-19)
+ PKr + p. [ or ~ or (rvr) + r2 (J(J2 - r2 o() + oz2
2. 8-direction:
a"'(x- Xo) (5.6-15)
g OVe+ vOVe
p (- - +-
- +Ve-OVe VrVe+ vOVe) - = - 1-op -
ot r or r o() r oz r o()
function of x but the exact position of
locate the interface, we must make use a o ) (1 2
0 Ve 1
Ovr
+ pg, + p. [or ~ or (rve) + ;:; o82 + ;:; o()
2 2
0 Ve]
+ OZ2 (5.6-20)
.,.,--,.,,.. the mass of the fluid in the tank
170 The Differential Equations of Motion Chap. 5 Sec. 5.6 Appiica~ions of the Differential Equa

2. The flow is steady. This would


conditions were changing with ti
3. The flow is laminar. For this assu

1
Grovity
Reynolds number be less than 210
flow becomes turbulent. (The tn
flow is discussed in Chap. 6.)
4. The velocity components, v9 and v
be true at the en trance of the pipe
zero at the entrance, they will
precise treatment of the flow in the
of this text; nevertheless, we may
experimental studies which indio
given by
L.= O
Entronce of pipe Fully developed

The entrance length is generally ~


centerline velocity is within 1 per

If we restrict our analysis to values o


valid. Imposing assumptions 3 and 4 o

l. r-direction:

0=--
op
or
2. 0-direction:
1 of
Fl. 5.6-l. Laminar ftow through a pipe. 0= - - -
r o6
3. z-direction:
3. z-direction:

p
av.)
(v.-
oz = - - +
oz
op pg.
[1 or
+ fL -o-
r

4. continuity equation:
4. continuity equation: ov.
-=
o oz
1
- - (rvr) + -1 -ov8 + -ov, = O (5.6-22)
ror roe oz Equation 5.6-27 indicates that v. is inde
In attacking this problem, we make four assumptions. the flow is axisymmetric-i.e., v. is inde
the problem is reduced to solving Eqs. ~
l. The density and viscosity are constant. This assumption is, of course,
reflected in the forro of the equations of motion and the continuity
op = P-[! ~
equation. oz ro.
Chap. 5
Sec. 5.6 Applications of the Differential Equations of Motion 171

2. The flow is steady. This would not be true if the entrance or exit
conditions were changing with time.
3. The flow is laminar. For this assumption to hold, we require that the
Reynolds number be less than 2100; at higher Reynolds numbers, the
flow becomes turbulent. (The transition from laminar to turbulent
flow is discussed in Chap. 6.)
4. The velocity components, v8 and vr, are zero. This condition need not
p= be true at the en trance of the pipe; however, even if v8 and vr are non-
zero at the entrance, they will tend toward zero as z increases. A
precise treatment of the flow in the en trance region is beyond the scope
of this text; nevertheless, we may draw upon other mathematical and
experimental studies which indicate that the entrance length L. is
given by
Le = 0.058 DNRe (5.6-23)

The entrance length is generally defined as the distance at which the


centerline velocity is within 1 per cent of its final value.

If we restrict our analysis to values of z greater than L., assumption 4 is


valid. Imposing assumptions 3 and 4 on Eqs. 5.6-19 through 5.6-22 yields

l. r-direction:

(5.6-24)

2. 0-direction:

(5.6-25)

3. z-direction:

4. continuity equation:

av. =o (5.6-27)
oz
(5.6-22)
Equation 5.6-27 indicates that v. is independent of z. If we also assume that
four assumptions. the flow is axisymmetric-i.e., v. is independent of 0-and note that g. = O,
the problem is reduced to solving Eqs. 5.6-24, 5.6-25, and 5.6-28.
om;tan1t. This assumption is, of course,
uations of motion and the continuity
op =P.[!~('
oz ror av.)J
or (5.6-28)
Sec. 5.6 Applications of the Differential Eq
172 The Differential Equatlons of Motion Chap. 5

and the dimensionless pressure gradie


Equations 5.6-24 and 5.6-25 may be integrated to give

f
p = pr g8 dO + C (r, z)1
(5.6-29)
f = tJ.p
lp(v,

p = f
p gr dr + C (0, z)
2 (5.6-30)
we obtain
u.= 2(1
and
where C1 and C2 are the constants of integration and may be functions of r
and z, and () and z, respectively. Differentiating Eqs. 5.6-29 and 5.6-30 with
respect to z and equating the results, we have
oC1(r, z) oC 2(0, z) where R = r/D. In this particular case
(5.6-31) putting the results in dimensionless f:
oz oz complex problem of turbulent flow in
Inasmuch as the right-hand side of Eq. 5.6-31 is nota function of r, the left- 5.6-38 for reference. The term f is kno
hand side cannot be a function of r. Similar reasoning with regard to the () consider its dependence on the Rey
dependence leads us to the conclusion that the derivatives of C1 and C2 may Chap. 8.
only be functions of z. Thus,
op oC (z)
- = -1 - = -2 -
oC (z) (5.6-32)
of
Flow two 1m miscible flulds betwe
oz oz oz pi ates
Referring now to Eq. 5.6-28, we conclude that the left-hand side may only
be a function of z, while the right-hand side may only be a function of r. Analyzing the flow of two immi
There is only one possible solution to this situation-i.e., both sides are equal. questions the student might have rega ,
to a constant-and we write of continuity of velocity and stress in fi
- tJ.p = ll! ~(r
L r dr
dv,)
dr
(5.6-33)
ftow illustrated in Fig. 5.6-3 consists of.
a pressure gradient in a channel of hei
analysis, we will assume that thc cha
where the partial derivati ves have been replaced with total deriva ti ves, because the pressure and velocity can be taken
v, is only a function of r. In writing Eq. 5.6-33, we have assumed that dpjdz At this point we need to say sometl
is a constant (equal to -tJ.pjL) over the entire region O ~ z ~ L; actually, it ticc of reducing a threc-dimensional p1
is not. If the entrance length L. is much smaller than the total length L, the
error in Eq. 5.6-33 is negligible. We solved this equation in Chap. 2 t yield

v. = (tJ.:)~~[l _ (~n (5.6-34)


and
Q = 1TT~
8/l
(tJ.p)
L
Very often it is beneficia} to put derived results in dimensionless form to
ndicate clearly what dimensionless parameters must be considered. If we
define the dimensionless velocity u. as
u=...!:!.. (5.6-35) FIJ. s.w. FJow or
(v,)
Differential Equatlons of Motlon Chap. 5
Sec. 5.6 Appllcatlons of the Differential Equations of Motion 173

and the dimensionless pressure gradient fas


integrated to give

(5.6-29) !-~(!!.)
- lp(v.)1 L
(5.6-36)

we obtain
(5.6-30) U1 = 2(1 - 4R1) (5.6-37)
and
integration and may be functions of r
ciu.lauJ11~ Eqs. 5.6-29 and 5.6-30 with
64
J=- (5.6-38)
NRe

where R = rfD. In this particular case, there is not much to be gained by


(5.6-31)
oz putting the results in dimensionless form; however, in treating the more
complex problem of turbulent flow in a pipe, it will be helpful to have Eq .
. 5.6-31 is nota function of r, the Ieft- 5.6-38 for reference. The termfis known as the friction factor, and we will
Similar reasoning with regard to the O consider its dependence on the Reynolds number for turbulent flow in
that the derivatives of el and c2 may Chap. 8.

=--
oCb) (5.6-32) Flow of two 1m miscible flulds between parallel
oz pi ates
that the left-hand side may only
side may only be a function of r. Analyzing the flow of two immiscible fluids will help to answer any
situation-i.e., both sides are equal. questions the student might have regarding the application of the principies
of continuity of velocity and stress in formulating boundary conditions. The
flow illustrated in Fig. 5.6-3 consists of two fluids flowing under the action of
(5.6-33) a pressure gradient in a channel of height h1 +hu. For the purpose of this
analysis, we will assume that the channel is infinite in the z-direction; thus,
replaced with total derivatives, beca use the pressure and velocity can be taken as independent of z.
5.6-33, we have assumed that dpfdz At this point we need to say something of the common engineering prac-
entire region O ~ z ~ L; actually, it tice of reducing a three-dimensional problem to a two-dimensional problem
smaller than the total length L, the
this equation in Chap. 2 t yield

(5.6-34)

must be considered. If we

(5.6-35) Fil. S.W. Flow of two immiscible ftuids.


(v.)
174 The Differential Equations of Motion Chap. S
Sec. 5.6 Applications of the Differential Equ

by assurning that the system is infinite in extent in one direction. Since no and Eq. 5.6-44 may be written as
real system extends infinitely in any direction, we would like to be able .to
I II
answer the question, "When will this technique provide a reasonable descnp- (D) = t(D)

tion of the real system ?" There is, in fact, no set answer to this question; An implied assumption in this treatn
each case must be considered individually. For the problem under considera- mathematical surface separating the two
tion, reasonable results are obtained with a two-dimensional analysis, pro- itself may be treated as a two-dimensio
vided the width of the channel is at Ieast an order of magnitude (factor of ten) for the interface can be incorporated into
Iarger than the channel depth, h1 +hu. We might press the issue an~ ask, tension, for example, may be treated ra
"Why is this so?" The answer les in examining the complete three-dtmen- such a treatment is beyond the scope of
sional solution and comparing it with the two-dimensional result. Because the to interfaces acting only as mathematica
mathematical techniques required to solve the three-dimensional case are Returning now to the solution of
beyond the scope of this text, we must limit o~rselves to .the simpler case. integrate Eqs. 5.6-40 and 5.6-42 to obtai
This restriction is satisfactory, for our purpose JS to examme the boundary
conditions for two-phase flow rather than to obtain an exact solution. Pr = -pgy
Assuming steady, one-dimensional flow of incompressible, constant- Prr = -prrgy
viscosity fluids allows us to simplify the equations of motion in rectangular
Equation 5.6-46 may be expressed as
coordinates to
B.C.l: r
t{l) = tn
O= -(opr)
ax
+ .ur(a2v~)
oy 2 (5.6-39) or, in component form,
(l)

) Fluid 1 TII
opr)
o=- (oy - Prg
(5.6-40)
B.C.la: I
T 1/Z=
I
1/Z

TII
B.C. lb: T 1111= 1111

B.C.lc:
0= - (~~II) + ,Un(~~~I) ) (5.6-41) Referring to Eq. 5.1-6 we see that the ce
Fluid 11 given by
O= - (oyopn) - Prrg .
(5.6-42) Tn =-pi
Making use of Eq. 5.1-3 and the resul
Before solving these equations, we must re-examine the behavior of the Eq. 5.6-51 as
stress vector at a phase interface. In Sec. 4.1, it was shown that

(5.6-43)
We may now write boundary condition
This result was based on the assumption that both the velocity vector and the
stress vector were continuous functions. Applying this result to the interface -Pr
av;) =
+ 2,ur (oy -Prr
between the two immiscible fluids, we write
Noting that
y=O (5.6-44)
we may substitute Eqs. 5.6-47 and 5.6-4
where the unit nqrmal, n, has been arbitrarily taken as the outwardly directed
unit normal for :fluid l. At any point in fluid II, CI(x) =
6. R. Aris, Vectors, Tensors, and the Basic
(5.6-45) Cliffs, N.J.: Prentice-Hall, Inc., 1962), Chap. 1
Differential Equations of Motion Chap. 5
Sec. 5.6 Applications of the Differential Equations of Motion 175
in extent in one direction. Since no
direction, we would like to be able to and Eq. 5.6-44 may be written as
provide a reasonable descrip- t~n) = tf;) Y= O (5.6-46)
fact, no set answer to this question;
An implied assumption in this treatment is that the interface is simply a
. For the problem under considera-
mathematical surface separating the two fluids. In actual fact, the interface
with a two-dimensional analysis, pro-
itself may be treated as a two-dimensional fluid and the equation of motion
an order of magnitude (factor of ten)
for the interface can be incorporated into the analysis. 6 The effect of surface
. We might press the issue and ask,
tension, for example, may be treated rationally by this method. However,
examining the complete three-dimen-
such a treatment is beyond the seo pe of this text, and we will limit ourselves
two-dimensional result. Because the
to interfaces acting only as mathematical surfaces separating two fluids.
solve the three-dimensional case are
Returning now to the solution of Eqs. 5.6-39 through 5.6-42, we first
limit ourselves to the simpler case.
integrate Eqs. 5.6-40 and 5.6-42 to obtain
purpose is to examine the boundary
than to obtain an exact solution. p 1 = - p1gy + C1(x) (5.6-47)
flow of incompressible, constant- Pn = - Pugy + Cu(x) (5.6-48)
equations of motion in rectangular
Equation 5.6-46 may be expressed as
B.C.l: 1 tll
tm = <ll y=O (5.6-49)
(5.6-39) or, in component form,

(5.6-40) B.C. la: T~., = r;! ) (5.6-50a)


n.c. lb: T! 11 = r;; y= o (5.6-50b)
B.C. le: T!. = T!! (5.6-50c)
(5.6-41) Referring to Eq. 5.1-6 we see that the component of the stress tensor, T1111 , is
given by
(5.6-42) (5.6-51)
Making use of Eq. 5.1-3 and the results given in Table 5.2-2, we express
must re-examine the behavior of the Eq. 5.6-51 as
4.1, it was shown that
Tv!l = -p + 2p(~;ll) (5.6-52)
(5.6-43)
We may now write boundary condition lb as
that both the velocity vector and the
Applying this result to the interface -p + 2pl (av
ay!/
1
) = -pu + 2pn (av a:- '
11
) y= o (5.6-53)
Noting that
y=O (5.6-44) (5.6-54)
taken as the outwardly directed we may substitute Eqs. 5.6-47 and 5.6-48 into Eq. 5.6-53 to obtain
fluid 11, C1(x) = C11 (x) (5.6-55)

(5.6-45) 6. R. Aris, Vectors, Tensors, and the Basic Equations of Fluid Mechanics (Englewood
Qiffs, N.J.: Prentice-Hall, lnc., 1962), Chap. 10.
S.C. 5.6 Applications of the Dilferential
176 The Dlfferential Equatlons of Motion Chap. S

where
Equ~tions 5.6-47 and 5.6-48 indicate that the pressure gradients, opfox
and op11fox, are not functions of y. Because v~ and v1! are not functions of x,
Eqs. 5.6-39 and 5.6-41 indicate that the pressure gradients must be constant,
and the solutions of these two equations are
ll.p) y2 If the pressure drop ll.p, length L,
L 2p.1 (
V~=-- -+Ay+BI (5.6-56)
h1 and hu are given, Eqs. 5.6-63 and 5.6
ity profiles and the flow rates. In pra
(5.6-57) with the volumetric flow rates for fluid
where pressure drop for a given system. This
solution where we assume values of h1 a1
(::1) = (::11) = - (/j.f) calculated by both Eq. 5.6-65 and 5.6

The principie of continuity of velocity yields the following boundary condi-


q =
f
\
dp) h~I [ ( /-'I) 1(h
L 2p.1 1-'u - 3 hu
1

tions:
B.C.2: V~= 0, y= -h (5.6-58a) qii = (ll.p) h~1 (2h !(hu)
11) _

VI= L 2p.u 3 2 h1
B.C. 3: ., VII., ' y=O (5.6-58b)
B.C.4: 1 =O,
v! y= hu (5.6-58c) 11
Here, q1 imd q are the volumetric flo
In general, the pressure drops will not
For a Newtonian fluid, boundary condition la becomes h11 must be assumed until the comput

ov! ov!) (av~


1
ov! 1)
B.C.la: 1-'1 (;-- + ;-- = 1-'n ;--- + ;- ' y= o (5.6-58d) Flow In a Couette vlscometer
uy ux uy ux
The four boundary conditions expressed by Eqs. 5.6-58 are used to determine In this example, we wish to det
the constants of integration as follows: annular region and the torque actin
shown in Fig. 5.6-4. We start with thq
B.C.2: - ep)M
- - -
L 2p.1
A1h1 + B1 = O (5.6-59) coordinates. If the flow is steady, and
infinite in the z-direction, the velocity
components vr and v, are zero and the
B.C. 3: B1 =Bu (5.6-60) Navier-Stokes equations reduce to

- ( P) h~1
11 1. r-direction:
B.C.4: + A11 h11 + B11 =O (5.6-61)
L 2p.n - P~ = -
r
(ap)
or
(5.6-67)
B.C. la: p.1A1 = p. A11 11 (5.6-62)
2. 0-direction:
Use ofthese four equations to evaluate the four constants ofintegration gives
the velocity profiles for the two phases: o= - !(op) or ror (rv,))J
r ao +p.[~(!~
v~ = (');;:[ {::) _(~JJ + n[ (:)- G: ~~:) J (5.6-63) 1
) 3. z-direction:
(5.6-68)

o = - (:~) -
v!l () 2~:Jl- (~JJ- (:I1I)(;:Jn[t- (~)]
pg (5.6-69)
(5.6-64)
S.C. 5.6 Appllcatlons of the Dlfferentlal Equatlons of Motlon 177
The Olfferentlal Equatlons ol Motlon Chap. S

where
that the pressure gradients, op1fox
l:ip) ( 1 ) (h~ h~) h
Because v~ and v1! are not functions of x,
the pressure gradients must be constant,
n= (1: ~ ~-~ 1

{~1 + hu)
\pi Pu
(5.6-56) If the pressure drop l:ip,
length L,
viscosities p 1 and p 11, and fluid depths,
h1 and h11 are given, Eqs. 5.6-63 and 5.6-64 may be used to calculate the veloc-
ity profiles and the ftow rates. In practice, we are more likely to be supplied
(5.6-57)
with the volumetric ftow rates for ftuids I and 11 and be asked to compute the
pressure drop for a given system. This problem would require a trial and error
solution where we assume values of h1 and hn, allowing the pressure drop to be
calculated by both Eq. 5.6-65 and 5.6-66.

= 1 !:ip)h~I [(PI)
2
yields the following boundary condi- Q _ !(!2.) ]h1 + nh 1[! + hn(PI )] (5.6-65)
\ L 2p Pn 3 hu 2 h1 Pn
y= -h (5.6-58a)
qu = (l:ip) h;I (2hn) _!(hu) (!!L.) !lhu (5.6-66)
y=O (5.6-58b) L 2p11 3 2 h1 Pn
(5.6-58c) Here, q1 and q11 are the volumetric ftow rates per unit width of the channel.
In general, the pressure drops will not be the same and new values of h1 and
h11 must be assumed until the computed pressure drops are equal.

y =o (5.6-58d) Flow In a Couette vlscometer

by Eqs. 5.6-58 are used to detennine In this example, we wish to detennine the velocity distribution in the
annular region and the torque acting on the inner cylinder for the system
shown in Fig. 5.6-4. We start with the Navier-Stokes equations in cylindrical
(5.6-59) coordinates. lf the ftow is steady, and the annular region is considered to be
infinite in the z-direction, the velocity
components vr and v, are zero and the
(5.6-60)
Navier-Stokes equations reduce to
1. r-direction:
A11 hu + Bn = O (5.6-61)
v
2
-p ..l.=- - (op) (5.6-67) Cup (rototesl
= ,UnAn (5.6-62) r or
2. 0-direction: Fluid to be
the four constants of integration gives tested in gap

o= - !(op)
r ao
+ p[!.(! ~ (rv~~))J
or r or
+ n[ (~) - {:: ) ~:) J
1 (5.6-68)
(5.6-63) 3. z-direction:

o = - (:~) -
-
1
{:: ) (;:Jn[1- (~J J (5.6-64)
pg (5.6-69) Flt. 5.6-4. Couette viscometer.
178 The Differential Equations of Motion Chap. S
Problems

Inasmuch as Remembering now that the express


9+2r

f
9
(~~)dO= O (5.6-70)
coordinates, provided we interpret
projections on the tangent vectors to
surface force per unit }
the pressure gradient in the 8-direction must be zero,t and Eq. 5.6'-68 may be flacting
area in the 0-direction
on the fluid
= th
written as
(1
-d -d- (rv 9 ) )
dr r dr
=O (5.6-71)
The components of the unit outwardl
inner cylinder are given by
Integration gives 11

ld and we obtain
- - (rv 9 ) = C1 (5.6-72)
r dr surface force per unit area )
exerted by the inner cylinder
Multiplying by r dr and integrating again, we get {on the fluid in the 0-direction

v9 = Cr + C2 The torque that the inner cylinder e


(5.6-73)
2 r
which is subject to the boundary conditions
B.C. 1: v9 = O, r = r1 (5.6-74) Force/unlt area

(5.6-75) From Eq. (d) in Table 5.2-3, we ob


B.C.2:
Application of these boundary conditions yields the velocity distribution,
Thus,

(5.6-76) .,.
r9 1r=r 1 -
- rd
--(V
P. [ dr l

Although we have assumed that the annular region extends to infinity and the torque is
in the z-direction, it is, of course, not the case for any real Couette viscometer. !Y".=-
In general, the viscometer is made such that r 2 - r 1 is much smaller than the
length, L. Under these conditions, the end effects are negligible and Eq.
5.6-76 is valid. If we let n represent the outwardly directed unit normal for
In this example, we treat the torque :
the fluid, the surface force acting on the fluid is given by formulate Eq. 5.6-82 as,
jsurface force per unit } = t
larea acting on the fluid (DI
(5.6-77)
!Y"= f
and in index notation we would write .A
but inasmuch as it requires taking
jsutface force per unit } _ t
larea acting on the fluid - (Dli cylindrical coordinates, it was not d<
= n1T;i (5.6-78)
= nTli + n2T2i + naTat PRO
t .The only other alternative is that opfoO takes on both positivc and negativc values 5-l. (a) Show that {ji = 3.
in thc rcgion O ~ O ~ 21r. (b) Show that {jiiAI = At.
Differential Equations of Motion Chap. S
Problems 179

Remembering now that the expression t(ali = n1T;1 applies to curvilinear


coordinates, provided we interpret the scalar components of vectors as
(5.6-70) projections on the tangent vectors to the coordinate curves, we may write
surface force per unit }
must be zero,t and Eq. 5.6'-68 may be farca in the 0-direction = t(alB = nrTre + n8 T 88 + n.Tze (5.6-79)
lacting on the fluid
The components of the unit outwardly directed normal n at the surface of the
(5.6-71) inner cylinder are given by
n 8 =O, n. =O (5.6-80)
and we obtain
(5.6-72)
surface force per unit arca }
{
exerted by the inner cylinder = - Tr 8 = -T rB r = r1 (5.6-81)
on the fluid in the 0-direction
The torque that the inner cylinder exerts on the fluid is
(5.6-73)
Area

T (5.6-82)
(5.6-74) Force/unlt area Lever arm

(5.6-75) From Eq. (d) in Table 5.2-3, we obtain

yields the velocity distribution, (5.6-83)


Thus,
2p.w0
(5.6-76) (5.6-84)

and the torque is


[t- (~r J
annular region extends to infinity
case for any real Couette viscometer.
(5.6-85)
that r2 - r1 is much smaller than the
end effects are negligible and Eq.
outwardly directed unit normal for
fluid is given by In this example, we treat the torque as a scalar. A rigorous treatment would
formulate Eq. 5.6-82 as,
(5.6-77)
9" = fr X t(a) dA (5.6-86)
A
but inasmuch as it requires taking the cross product of two vectors in
cylindrical coordinates, it was not done.
n;T;i (5.6-78)
n1 T + n2T21 + naTa, PROBLEMS
takes on both positive and negative values
5-1. (a) Show that 6 = 3.
(b) Show that f1 11A 1 =A,.
Problems
180 The Differential Equations of Motion Chap. 5

5-2. lf B11 = B22 = B33 =t. B12 = B21 = !, B31 = B13 = 1, and B32 = B23 =O, tested. (d) The height of the ba
write out the nine components of the tensor A;1, given that is varied.
5-9. Determine the Reynolds number
A;; = B; 1 + O;;
film shown in Fig. 5-9; show tha
5-3. Prove, using Gibbs notation, that

(n T) b = (b T) n
provided T is symmetric. Expand both sides and show that they are equivalent.
5-4. Prove
V (pi)= Vp
using index notation.
5-5. A metal rod, shown in Fig. 5-5, is
being stretched and compressed
sinusoidally. The displacement ~x
of the end of the rod is given by
~x = -l ~L (cos wt)
where

----
1
' Grovity
~L ~ L
x=O x =L
lf the displacement is a linear
function of x, what is the rate of Fig. 5-S
strain as a function of x and t?
5-6. Is the tensor Vv symmetric?
5-7. If A is a tangent vector to a free surface, what is the value of A Vp at the free
F
surface?
5-8. A stirred tank such as that shown Obtain the functional relationshi
in Fig. 5-8 is a common device Stirred tonk
is laminar and one-dimensional,
for blending or mixing two fluid
streams. The rate of mixing de- Ans: NFr =N Re sin 0/3
pends on the propeller geometry 5-10. When two immiscible fluids, su
and the size and number of baftles. simultaneously pumped throug
In designing a mixer, one must occur depending on the relative
weigh the cost of the original two fluids. Indicate what para
equipment versus the operating dynamic similarity between a m
cost (i.e., power input). To do - d - tension effects are to be neglecte
this, experiments must be per effect for sorne systems.
formed to determine the power O
input as a function of propeller
1 5-11. A problem of sorne practica! im
geometry and the baftling arrange- Flg. S-8 another in a pipe. Such a situati
ment. What dimensionless groups replacement taking place from t
will the power input depend upon? Assume: (a) The propellers to be used domnate (i.e., the left-hand sid
in the experiments are all geometrically similar. (b} The number of baflles what parameters must be held co
is fixed, but the size may vary. (e) A number of tank diameters are to be tension.
Differential Equations of Motion Chap. 5 Problems 181

= J, B31 = B13 = 1, and B32 = B23 = O, tested. (d) The height of the baffies remains constant; only the width b
tensor A 0 , given that is varied.
5-9. Determine the Reynolds number and Froude number for the falling Iiquid
film shown in Fig. 5-9; show that they cannot be specified independently.

sides and show that they are equivalen t.

(pi) = Vp

-
1
......--
x=O x =L
Fig. 5-5

what is the value ofA Vp at the free


Fig. 5-9

Stirred tonk Obtain the functional relationship between NRe and NFr Assume the flow
is laminar and one-dimensional, and that q is given.
Ans: NFr = NRe sin 0/3
5-10. When two immiscible fluids, such as oil and water or water and air, are
simultaneously pumped through a pipe, a variety of flow patterns may
occur depending on the relative flow rates and the physical properties of the
two fluids. Indicate what parameters you would fix in order to maintain

.._,.- - 0
-d- dynamic similarity between a model system and an actual system. Surface
tension effects are to be neglected, although this may be the most important
effect for sorne systems .
5-11. A problem of sorne practica) importance is the replacement of one fluid by
Flg. 5-8 another in a pipe. Such a situation is illustrated in Fig. 5-11, which shows the
replacement taking place from the action of gravity. If viscous forces pre-
? Assume: (a) The propellers to be used dominate (i.e., the left-hand side of the equations of motion is negligible),
similar. (b) The number of baffies what parameters must be held constant to model this system? Neglect surface
A number of tank diameters are to be tension.
182 The Dlfferential Equations of Motion Chap. 5 Problems

!
Grovity
f>O

5-14. Determine the velocity profile an


the plane Couette flow illustrate
parallel plates, the lower fixed
applied pressure gradient 3pf3x
Volve closed ot Volve open ot required to reduce the volumetri
downstreom end downstreom end profile for this condition.
5-15. A film of water is flowing dow
Specify how large the air gap hn
5-12. A rectangular tank containing a fluid is sliding down the inclined plane shown water velocity profile can be negl
in Fig. 5-12. The mass of the fluid plus the tank is M, and the coefficient of
friction between the tank and the plane is ~. Derive an expression for the
angle p, obtained during uniform acceleration. Water
Ans: {3 = tan-1 ~

Fig. 5-12 Fig, 5-13

5-13. If a cylindrical tank of liquid is rotated at a constant angular velocity w 0 ,


the free surface will assume sorne shape, such as that illustrated in Fig. 5-13.
Use the equations of motion in cylindrical coordinates to determine the
F
position of the surface. The fluid depth at rest is h0
Ditrerential Equations of Motion Chap. 5 Problems 183

~~~~~~~~~~~~~~~~~~
-uo
p,

t>O

Fig. 5-14

5-14. Determine the velocity profile and the volumetric flow rate per unit width for
the plane Couette flow illustrated in Fig. 5-14. The system consists of two
parallel plates, the lower fixed and the upper moving at a velocity u 0 The
applied pressure gradient apax is a constan t. Determine the value of apax
Volve open ot required to reduce the volumetric flow rate to zero, and sketch the velocity
downstreom end
profile for this condition.
5-15. A film of water is flowing down a vertical wall as illustrated in Fig. 5-15.
Specify how large the air gap hu must be so that the effect of the air on the
is sliding down the inclined plane shown water velocity profile can be neglected.
plus the tan k is M, and the coefficient of
is ~- Derive an expression for the
Water Air ot otmospheric pressure

Fig. 5-13

at a constant angular velocity w 0 ,


such as that illustrated in Fig. 5-13.
""'n"'".'~ coordinates to determine the Air ot otmospheric pressure
Fig. 5-15
Problems
la.. The Dlfferentlal Equatlons of Motlon Chap. 5
S-18. A wire is pulled through a circular
S-16. Determine the magnitude and direction of the maximum normal viscous stress as shown in Fig. 5-18. The two
in Prob. S-14-i.e., maximize Tflfl for the case of zero ftow rate. pressure, p0 Derive an expression
chambers, and an expression for
the cylinder.
S-19. If inertial effects can be neglected
the stream function 'P defined in
J41p
ax' + 2 a.r
5-20. Given a tensor

write out in matrix form the comp(!


tensors into which Aii can be dec
5-21. A long circular cylinder of radius
infinite fluid. Determine the vel

Fl1. 5-17

S-17. A falling liquid film is formed on the cylindrical rod illustrated in Fig. S-17.
Determine the velocity profile in the region of one-dirnensional flow.
2
Ans: v. = -PK (r: - rl) + -pgr2 Jn (-' )
4J.' 2p r1

Fil

5-22. If the velocity of the cylindrical 1


determine the force which must b
Assume that viscous effects predc
may be approximated by steady,
force, remember to include both v
that you cannot obtain an exact s
Fil. 5-18
Problems 185
Dlfrerentlal Equatlons of Motlon Chap. 5
S-18. A wire is pulled through a circular cylinder which connects two large charnbers
ofthe maximurn normal viscous stress as shown in Fig. 5-18. The two chambers are maintained at a constant
the case of zero ftow rate. pressure, p0 Derive an expression for the volumetric ftow rate between the two
charnbers, and an expression for the force required to pull the wire through
the cylinder.
5-19. Ifinertial effects can be neglected and the ftow is two-dimensional, show that
the stream function 'P defined in Sec. 3.6 must satisfy the biharmonic equation,
~'P ~'P ~'P
ax' + 2 a.x~ ay2 + ay4 = 0
5-20. Given a tensor

write out in matrix form the components of the symmetric and skew symmetric
tensors into which Aii can be decomposed.
5-21. A long circular cylinder of radius r0 rotates with an angular velocity w0 in an
infinite fluid. Determine the velocity profile.

cylindrical rod illustrated in Fig. 5-17.


region of one-dimensional ftow. uo,---+l

o -o,
- 2- < < 1
o,
Flg. s-n
5-22. If the velocity of the cylindrical ram shown in Fig. 5-22 is a constant u0 ,
determine the force which must be applied to the ram as a function of time.
Assume that viscous effects predominate and the ftow in the annular region
may be approximated by steady, one-dimensional ftow. In computing the
force, remember to include both viscous and pressure forces, keeping in mind
that you cannot obtain an exact solution.
11
Sec. 6.1 Time Averages

*6.1 Time Averages

We define the time average of s~

Turbulent Flow 6 S=-


2

The time interval !:::..t is arbitrary, bu


be made large enough so that S is 1
on At is illustrated in Fig. 6.1-1, -v

In the preceding chapter the equations of motion were solved for one-
dimensional laminar flows and uniforrnly accelerated flows. The differential
equations proved useful in analyzing these problems only because the equa-
tions could be reduced to comparatively simple forms. When the flow is
turbulent, the fluid motion is still govemed by the differential equations;
however, all three components ofthe velocity vector will be nonzero functions
of the spatial coordinates and time. Under these conditions, we are not able
to obtain exact solutions to the equations of motion.
Since so many practica} engineering problems deal with turbulent flow,
we must develop sorne method of analyzing such flows, the key tools of which
will be: the time-averaged equations of motion and continuity; the macro-
scopic mass, momentum, and mechanical energy balances; and experimental
or empirical friction factors and drag coefficients. We will first deal with the Fl1. 6.1-1. Effect of At
time-averaged equations of motion and then present a qualitative description
ofturbulent flow. Chapter 7 is devoted to the formulation ofthe macroscopic
balances and their application to problems where viscous effects may be function S. Not all quantities ha
neglected. Chapter 8 deals with the definition of friction factors and drag independent of At; however, the ,
coefficients and the application of the macroscopic balances to problems possess meaningful averages for tur
where viscous effects are important. immersed bodies. For other cases,
not hold true, and the analysis pre
186
Sec. 6.1 Time Averages 187

*6.1 Time Averages

We define the time average of sorne function S ata time t as

6 - 1 ft+4t
S = - Sdt
2dt
t-4t
(6.1-1)

The time interval dt is arbitrary, but in general we would hope that dt could
be made large enough so that S is independentof dt. The dependence of S
on dt is illustrated in Fig. 6.1-1, where M* depends on the nature of the

of motion were solved for one-


accelerated flows. The differential
problems only because the equa-
simple forms. When the flow is
by the differential equations;
vector will be nonzero functions
these conditions, we are not able
ofmotion.
problems deal with turbulent flow,
such flows, the key tools of which
motion and continuity; the macro-
energy balances; and experimental dt---
toelficients. We will first deal with the Fl. 6.1-1. Effect of tl.t on the time-averaged value of S.
then present a qualitative description
the formulation ofthe macroscopic
where viscous effects may be function S. Not all quantities have meaningful time averages in that S is
p1u1uuuu of friction factors and drag independent of dt; however, the components of velocity and pressure do
macroscopic balances to problems possess meaningful averages for turbulent flows in closed conduits and around
immersed bodies. For other cases, such as atmospheric turbulence, it may
not hold true, and the analysis presented in this section would not be valid.
Turbulent Flow Chap. 6 Sec. 6.2 Time-Averaged Equations of Co1
188

lf we measured one component of the velocity in a slowly changing tur- In the time interval 2 D..t, the time-av
bulent stream, the result might look like the curve in Fig. 6.1-2, where the stant, and Eq. 6.1-3 reduces to
average velocity v. changes only slowly with time compared to the ftuctuating
component v;.To separa te the velocity into an average value anda fluctuating v. = v. +l
value, we require that D..t be Iarge enough for us to average the ftuctuations
We thus see that the time average o

v'
However, the average of the square
quantity

is used as a convenient measure of the


It is known as the "intensity of turbt
0.01 to 0.10 for most turbulent ftows.

percentage turbul

and is therefore in the neighborhood

6.2 Time-Averaged Equatio


Continuity ~nd Motion
Time
Flg. 6.1-l. The turbulent velocity.
Contlnulty equatlon

effectively, yet small compared to the time during which significant variations
in v. occur. Note that if v. is nota function of time D..t is unbounded and this Restricting ourselves to incompt
condition obviously can be satisfied. From experimental studies we know equation as
that the turbulent fluctuations are very rapid compared to variations in the
v.
mean ftow, and we will assume that sorne time interval can always be chosen Writing the velocity vector in te
so that the fluctuations can be averaged without obscuring the time variations components,
of the mean flow. V=
Expressing the velocity in terms of the time average plus a fluctuating allows us to write Eq. 6.2-1 as
component,
V:= V:+ V~ (6.1-2) Vv +
and substituting into Eq. 6.1-1, we have Taking the time average yields

-Jv
. l+AI
t+At t+At
1
v

= -2 1D..t
- Jv ii-_ -2 D..t J(v + v') dt
1
- (6.1-3) -
2 D..t
1-At
vdt +
t-Al t-Al
Turbulent Flow Chap. 6 Sec. 6.2 Time-Averaged Equations of Continuity and Motion 189

the velocity in a slowly changing tur- In the time interval 2 11t, the time-averaged velocity may be considered con-
the curve in Fig. 6.1-2, where the stant, and Eq. 6.1-3 reduces to
with time compared to the fluctuating 1+41
into an average value anda fluctuating
for us to average the fluctuations
v. = v. +
1
-2 11t
- Jv;
1-M
dt (6.1-4)

We thus see that the time average of the fiuctuation is zero.

v; =O (6.1-5)
However, the average of the square of the fiuctuation is not zero, and the
quantity

v.
is used as a convenient measure ofthe magnitude ofthe turbulent fluctuations.
It is known as the "intensity of turbulence" and ranges from approximately
0.01 to 0.10 for most turbulent fiows. The percentage turbulence is defined by

percentage turbulence = 100 x -_--


J(:)i
(6.1-6)
v.
and is therefore in the neighborhood of 1 to 10 per cent.

6.2 Time-Averaged Equations of


Contlnulty ;ind Motion
velocity.
Contlnulty equatlon
during which significant variations
of time 11t is unbounded and this Restricting ourselves to incompressible flows, we write the continuity
From experimental studies we know _ equation as
rapid compared to variations in the (6.2-1)
time interval can always be chosen
Writing the velocity vector in terms of the time average and fluctuating
without obscuring the time variations components,

the time average plus a fluctuating


v = f v' + (6.2-2)
allows us to write Eq. 6.2-1 as

(6.1-2) V f +V v' =O (6.2-3)


Taking the time average yields

f f
- 1+41 1+41
1 1
(6.1-3) - - V vdt + - - V v' dt = O (6.2-4)
2 11t 2 11t
1-41 1-41
190 Turbulent Flow Chap. 6 Sec. 6.2 Tlme-Averaged Equatlons of

We may change the order of differentiation and integration to obtain Similarly, the pressure term reduce

V[
2
~t fvt+At ]

t-At
dt +V
[

2
~t ft+At ]

t-At
v' dt = O (6.2-5) 1
2M
t-

Since the second integral is zero by definition and the first integral yields the In treating the nonlinear, convecti
time-averaged velocity vector, the time-averaged continuity equation is velocity in terms of vand v'
Vv =O (6.2-6) t+At
and from Eq. 6.2-3 we note that 1
- -
2dt
Jv Vv dt = -2
t-At
Vv' =0 (6.2-7)
and expand the result to obtain
Following the discussion in Sec. 3.6, we may use the time-averaged velocity
vector to define a time-averaged streamline given by t+At

dx
-=-=-
dy dz (6.2-8)
1
- -
2dt
Jv Vv dt =
t-At

Equations of motlon

Starting with the Navier-Stokes equations

(6.2-9)

we form the time average to obtain

p(- 1 Jt+(av) dt
2 dt ot
t-At
+-1
2 dt
';;'.
t-At
vv) dt)
Starting with term (a) on the righ
comment that time-average terms :
f f
t+At t+At
1 1 thus,
= - - -
2 dt
Vp dt + pg + -2dt
- p, V2v dt (6.2-10)
t-At t-At

In analyzing Eq. 6.2-10, we shall start with the term farthest to the right and
work steadily to the left, thereby dealing with the easiest terms first. Starting
with the viscous term, we change the order of differentiation and integration Term (b) can be handled by rem
to obtain obtain
t+At
1
- -
2 dt
Jt+At
"'v2v dt
t-Al
= "'V 2
[
-
2dt
1
- Jv
t-At
t+At ]
dt = "'V v 2
(6.2-11) 1
- -
2dt
Jv' Vv dt =
t-At
Turbulent Flow Chap. 6
Sec. 6.2 Tlme-Averaged Equations of Continuity and Motion 191

l
and integration to obtain
Similarly, the pressure term reduces to

~t Jv' dt
1+41 1+41
[ = O (6.2-5) 1
2 - -JVpdt = Vp (6.2-12)
1-41 2~t
1-41
and the first integral yields the
continuity equation is In treating the nonlinear, convective acceleration term, we must express the
velocity in terms of vand v'
=0 (6.2-6)
1+41 1+41 '
1 1
- - J v Vv dt = - - J (v
2~t 2~t
+ v') V(v + v') dt (6.2-13a)
=0 (6.2-7) 1-41 1-41

may use the time-averaged velocity and expand the result to obtain
given by
1+41 1+41

dz
(6.2-8)
1
- -
2~t
J 1
v Vv dt = - - J v Vv dt
2~t
(a)
v. 1-41 1-41

1+41

+ 2~t- 1- J
1-41
v' Vv dt (b)

(6.2-13b)
1+41

(6.2-9) + 2~t
1
J
- - v Vv' dt (e)
1-41

t+4t

+ - 1- J v' Vv' dt (d)


) 2~t
t-41

Starting with term (a) on the right-hand side of Eq. 6.2-13b, we need only
comment that time-average terms are essentially constant in the period 2 ~~;
f
1+41
1
dt + pg + -2~t
- p. V2v dt (6.2-10) thus,
1-41 1+41

the term farthest to the right and


-
2~t
1
J-V v-a - v-V
V t=V (6.2-14a)
t-4t
with the easiest terms first. Starting
of differentiation and integration

l
Term (b) can be handled by removing the tensor Vv from the integral to
obtain
1
- -
2at
J dt 1+41
v = p. V 2v (6.2-11) 1
- - Jv' Vv dt =
t+4t [
1
-2~t
- Jv' dt
1+41 ]
Vv =o (6.2-14b)
1-41 2~t
1-41 t-41
192 Turbulent Flow Chap. 6 Sec. 6.2 Time-Averaged Equations of Co

since the time average of the fluctuating component is zero. Term (e) is We must keep in mind that we ares
treated in a similar manner by removing the scalar operator v V from the functionf(t) as
integral. a= lim [f(t -ll
at
1
- -
2 ~~
Jv t+4t
Vv' dt = v V
[
1
- -
2 ~~
J t+4t ]
v' dt = O (6.2-14c)
dt-o

The limits of integration in Eq. 6.2-l


t-4t t-4t to yield
~=t
Term (d) represents the turbulent convective transport, and for the present
we simply express this term as
av . ~=H4
J vd
- = 1Im
1 Jt+4t - at dt-o
- v' Vv' dt = v' Vv' (6.2-14d)
2~t
t-4t Applying the mean value theorem to
In treating the remaining term in Eq. 6.2-10, we must proceed carefully, av = lim {(2~t)vlt+
paying proper attention to the use of dummy variables of integration. What at dt ... o
we wish to prove here is that the average of the derivative is equal to the
derivative ofthe average. This statement has a great deal of.intuitive appeal, where r:x and Plie between l. Ca
and, beca use the proof is a bit laborious, the student may simply wish to sean limit, we get
the development quickly and move directly to Eq. 6.2-23. Expressing the av 1
integral properly in terms of a dummy variable of integration, we find at = 2~/

_1_
2 ~~
J t+4t
av dt -
at - 2 ~~
_1_ f ~=t+4t

av d
a'T} '1J
(6.2-15)
Comparison of Eqs. 6.2-16 and 6.2-2
a ti ve is equal to the deriva ti ve of th
t-4t lf=t-4t t+

Carrying out the integration yields an expression for the average of the
deriva ti ve,
2~t J t-4t

The average Collection of all the terms containe


ofthe (6.2-16) the time-averaged equations of moti
derivative

We now turn our attention to the derivative of the average.


P (av
at +-V. v-) V =- p v-
~; ~ ~~ ~~ 1-::,:
We see that this result has precise
The derivative
ofthe average (6.2-17) equation with the exception of the
(2
that this term can be represented as

In this part of the analysis, we must remember that ~~ is a parameter and not pv' Vv'
a variable. Expressing the derivative in terms of its definition yields
We can prove Eq. 6.2-25 by expand'
lf=H4Hdt lf=H4t-dt equivalent; however, the proof can

av D
J V dr - J V dr Todo so, we first express Eq. 6.2-7

- = l i1
. q=t-4t+dt q=t-4t-dt (6.2-18) a
(2~t) (2 ~t)
at dt-o
a
Turbulent Flow Chap. 6 Sec. 6.2 Time-Averaged Equations of Continuity and Motion 193

component is zero. Term (e) is We must keep in mind that we are simply expressing the derivative of sorne
the scalar operator v V from the function J(t) as
' of = lim [f(t + M- f(t - MJ (6.2-19)

2 ~1 f
t+b.t ] ot dt-oo 2bt
v V [ v' dt =O (6.2-14c) The limits of integration in Eq. 6.2-18 may be rearranged by a series of steps
t-dt to yield
~=t+4t+dt ~=t-4t+dt

I V dr - V dr I
-ov = l. ~=t+dl-61
tm
~=t-dt-dt (6.2-20)
ot dt--o (2M (2 ~~)
(6.2-14d)
Applying the mean value theorem to these two integrals gives
Eq. 6.2-10, we must proceed carefully, ov = lim {(2bt)vjt+4t+adt- (2bt)vlt-b.t+/Mt}
dummy variables of integration. What (6.2-21)
ot dt--o (2M (2 ~t)
1n"''rmo of the derivative is equa/ to the

has a great deal of intuitive appeal, where oc and ~ lie between l. Cancelling the factor of 2bt and taking the
the student may simply wish to sean limit, we get
directly to Eq. 6.2-23. Expressing the The derivative ( . _ )
6 2 22
variable of integration, we find of the average

1
2 ~~
f ~=t+dt

ov d
or r
(6.2-15)
Comparison of Eqs. 6.2-16 and 6.2-22 indica tes that the average of the deriv-
ative is equal to the deriva ti ve of the average, and we write
~=t-dt

an expression for the average of the _1_


2 ~~
It+4t

av dt = ~
ot ot (6.2-23)
t-dt

The average Collection of all the terms contained in Eqs. 6.2-11 through Eq. 6.2-23 yields
of the (6.2-16) the time-averaged equations of motion.
derivative

of the average.
p av + _
( Ot V
v-) = - v _+
V p pg
+ fl t'72-
v V -
~v
pV
V
, (6.2-24)

We see that this result has precisely the same form as the Navier-Stokes
The derivative
of the average (6.2-17) equation with the exception of the term pv' Vv'. Our next task is to show
that this term can be represented as the divergence of a tensor, i.e.,

~member that ~~ is a parameter and not pv' Vv' = V (pv' v') (6.2-25)
terms of its definition yields
We can prove Eq. 6.2-25 by expanding both sides and showing that they are
equivalent; however, the proof can be obtained easily using index notation.
To do so, we first express Eq. 6.2-7 in index notation
(6.2-18)
ov~ =o (6.2-26)
OX
194 Sec. 6.3 A Qualitative Description of Turb1
Turbulent Flow Chap. 6

and then an~lyze the right-hand side of Eq. 6.2-25 using index notation and This result indicates that we may ti
disregarding the time-averaging. the same way that we treat laminar flo
are replaced by the time-averaged qua
aa (
X;
pV;V;
, ') = pV;-
, av; + V, ;a- (pV;')
axi ax;
(6.2-27) with the total time-averaged stress tens
gradients of vin a manner analogous t1
Here, we ha ve simply used the rule for differentia:ting a product of two func- calculate turbulent velocity profiles i1
tions. Since we are only 'reating incompressible fl.ows, the second term on laminar flow. It would require a cons
the right-hand side of Eq. 6.2-27 is zero by Eq. 6.2-26 however, we now have at our disposa
which must be used in conjunction
, a (pV;') = pV;,(-av;)
V;- = (6.2-28) velocity profiles. Even though we do n
axi axi it is extremely important to know that!
and our proof is complete. form indicated by Eqs. 6.2-6 and 6.2-3
a ( , ') '(av;) mind when we derive and apply the m
ax; pV;V; = pV; -axi
lndex notation -

(6.2-29)
Gibbs notation V (pv' v') = pv' Vv' *6.3 A Qualitative Descriptio
We now define the turbulent stress tensor Turbulent Flow

i 10 = -pv' v' (6.2-30) As we illustrated in Chap. l with a


and write the time-averaged equations of motion as classified as either laminar or turbulent.
tion from laminar to turbulent flow, is
av
P ( at+vVv ) =-VfJ+pg+t-tV2v+Vi't) (6.2-31) rather difficult to discuss in a quantit~
related to the Reynolds number, which
The terms represented by pv' v' are often called the Reynolds stresses, 1 forces, pu~, to viscous forces, .tu0/L.
for it was Reynolds who first applied the Navier-Stok~~in deriving the general ~ pu 0 L _ pug _ inert
equations of motion for turbulent fl.ow. N Re- - -.
ft .tU o V!SCC
Returning to Chap. 5, we note from Eqs. 5.4-1 and 5.4-10 that the viscous
L
term may be written
(6.2-32) At Reynolds numbers greater than abe
duits become unstable, which means t
and the time-averaged form is
will tend to become larger and the la
(6.2-33) the growth of turbulent eddies. When
Substitution of Eq. 6.2-33 into Eq. 6.2-31 yields the time-averaged stress force that the fluid exerts on solid su
equations of motion pressure drop is required to maintain a

p(:: + v Vv) = -V[J + pg +V i!Tl (6.2-34)


is not the sole criterion for predicting
fl.ows (i.e., fl.ows in a converging secti
decelerating flows. Indeed, a transitior
where the total time-averaged stress tensor i<Tl is simply the sum of the time- diverging section of conduit at Reyn1
averaged molecular stress tensor and the turbulent stress tensor, 2 x 103 This fact should not be part
(6.2-35) that the velocity is a functioJ;l of the i
the parameters appearing in the equat1
1. O. Reynolds, "On the Dynamical Theory of Incompressible Viscous Fluids and
the Determination of the Criterion," Phi/. Trans. Roy. Soc. (London) Ser. Al, 1895,
eters appearing in the boundary con'
186:123. conduit, a parameter indicating the rat'
Turbulent Flow Chap. 6 Sec. 6.3 A Qualitative Description of Turbulent Flow 195

Eq. 6.2-25 using index n,otation and This result indicates that we may treat time-averaged turbulent ftows in
the same way that we treat laminar flows, provided the pressure and velocity
are replaced by the time-averaged quantities, and the viscous stress tensor
-ov~ a (pv ')
+ v, -1 (6.2-27) with the total time-averaged stress tensor. If we could specify -r<t) in terms of
OX OX
gradients of v in a manner analogous to Newton's law of viscosity, we could
differentiating a product of two func- calculate turbulent velocity profiles in just the same manner as we did
"ble ftows, the second term on laminar flow. It would require a constitutive equation for turbulent ftow;
by Eq. 6.2-26 however, we now have at our disposal only empirical expressions for .:r<t),
which must be used in conjunction with experimental data to determine
,(av;) = 0
pv 1. - (6.2-28) velocity profiles. Even though we do not know how to evaluate .:r<t) a priori,
OX
it is extremely important to know that the time-averaged equations take the
form indicated by Eqs. 6.2-6 and 6.2-34. These will be key points to keep in
pv;-,
(av~) mind when we derive and apply the macroscopic balances.
OX
(6.2-29)
*6.3 A Qualitative Description of
Turbulent Flow
(6.2-30) As we illustrated in Chap. 1 with a burning cigarette, most ftows can be
of motion as classified as either laminar or turbulent. The onset of turbulence, or the transi-
tion from laminar to turbulent flow, is an active area of current research and
(6.2-31) rather difficult to discuss in a quantitative manner. The transition is closely
related to the Reynolds number, which may be considered the ratio of inertial
often called the Reynolds stresses, 1 forces, pu~, to viscous forces, ,tu0 / L.
Navier-Stoke_~_in deriving the general
N Re ~ pu 0 L = pu~ = inertial force per unit area ( . _
630
,t ,tu 0 viscous force per unit area
Eqs. 5.4-1 and 5.4-1 Othat the viscous
L
(6.2-32) At Reynolds numbers greater than about 2 x 103 , most flows in closed con-
duits become unstable, which means that any small disturbance in the flow
will tend to become larger and the laminar velocity profile is destroyed by
(6.2-33) the growth of turbulent eddies. When the flow becomes turbulent, the drag
1 yields the time-averaged stress force that the fluid exerts on solid surfaces increases sharply, and a larger
pressure drop is required to maintain a given flow rate. The Reynolds number
is not the sole criterion for predicting the onset of turbulence. Accelerating
(6.2-34) ftows (i.e., flows in a converging section of conduit) are more stable than
decelerating flows. Indeed, a transition to turbulent ftow can take place in a
r .:r<T) is simply the sum of the time- diverging section of conduit at Reynolds numbers far below the value of
turbulent stress tensor, 2 x 103 . This fact should not be particularly surprising, because we know
(6.2-35) that the velocity is a function of the independent variables (x, y, z, and t),
of Incompressible Viscous Fluids and
the parameters appearing in the equations of motion (NRe), and the param-
. Roy. Soc. (London) Ser. AI, 1895, eters appearing in the boundary conditions. In a diverging or converging
conduit, a parameter indicating the rate of change of cross-sectional area will
196 Turbulent Flow Chap. 6 Sec. 6.3 A Qualitative Description of Turbu

appear in the boundary conditions; thus, we should expect the transition to


turbulence to be affected by this parameter.

Turbulent flow in a tu be

When the flow is laminar and the fluid is Newtonian, the velocity profile
in a tube is
(6.3-2) 1
1
1
Although we cannot compute the velocity profile directly for turbulent flow, 1
1
we can obtain experimental measurements of the time-averaged velocity and 1
1
the fluctuating components. A representative curve for the velocity is shown 1
1
in Fig. 6.3-1 based on the approximate expression 1
1
..--Lom
~r
7

v. = v.,max ( 1 - (6.3-3) 1
1

y= dist
Equation 6.3-3 could not be used to predict the shear stress at the tube wall,
because the derivative of v. with respect to r tends to infinity as r-+ r 0 Fig. 6.3-2. Laminar sublay
a tube.
Tube center--J
1
Nevertheless, it is a reasonably accur
profile.
In the central region of the tube, t
the turbulent stress far exceeds the vis1

1
Since the velocity fluctuations are zero
k---+----Turbulent core ---~Ir------<""\
-pv' v' = ~(tl =
= 0.6
o;;," e. there must be a transition region wh1
1;;;." comparable. The region outside the c1
regions, as shown in Fig. 6.3-2.

Stresses in turbulent tube flow

If we examine the time-averaged st


tion for cylindrical coordinates, we ot

+ v av. + ~ av. + v. av.)


1.0 p (av.
ot ror r i)() oz
o
Fig. 6.3-1. Laminar and turbulent velocity profiles.
= -(~~) + pg. + [;
Turbulent Flow Chap. 6 Sec. 6.3 A Qualitative Description of Turbulent Flow 197

we should expect the transition to

;.---Turbulent core

is Newtonian, the velocity profile

(6.3-2)
1
1
1
profile directly for turbulent flow, 1
of the time-averaged velocity and 1
1
ve curve for the velocity is shown 1
1
1
1
- .!:..)1/7 (6.3-3)
1
1,. Laminar sublayer
1
ro 1

the shear stress at the tube wall, y= distance from the wall
to r tends to infinity as r-+ r 0
Fig. 6.3-2. Laminar sublayer and transition region in
a tube.

Nevertheless, it is a reasonably accurate expression of the turbulent velocity


profile.
In the central region of the tube, the velocity ftuctuations are large and
the turbulent stress far exceeds the viscous stress-i.e.,

i< 1> ~ i In the core (6.3-4)


Since the velocity ftuctuations are zero at the wall,

- pv' v' = i<t> = O At the wall (6.3-5)


there must be a transition region where turbulent and viscous stresses are
comparable. The region outside the core can be arbitrarily divided into two
regions, as shown in Fig. 6.3-2.

Stresses in turbulent tube flow

If we examine the time-averaged stress equation of motion in the z-direc-


tion for cylindrical coordinates, we obtain, from Table 5.1-2,

ov. + _ov. + v ov. + _ov.)


6
p-(ot vor- -r o()- voz-
r
L'o ---
= _ (ofi) + pg. + [l ~ (rf~;>) + l o-r~;> + o-r~;>] (6.3-6)
velocity profiles.
oz r or r o() oz
Sec. 6.3 A Qualitative Description of Turb~
198 Turbulent Flow Chap. 6

Following the analysis for laminar flow, we make four assumptions:


l. vr = v8 = O, the time-averaged flow is one-dimensional;
2. av.foz = o, following from assumption 1 and the time-averaged con-
tinuity equation;
3. f~';>, f~';>, and -r~;> are independent of () and z;
4. g. =o.
Equation 6.3-5 immediately reduces to

(aozft) = 1 ~ (r-r~;'>)
r or
(6.3-7)

Once again we argue that the right-hand si de of Eq. 6.3-7 is independent of z


while the left-hand side is independent of r. We conclude that both sides are
equal to a constant and write

- t.p = 1 !!._ (r-r<T>) (6.3-8)


L r dr rz

1.0

Fig. 6.3-4. Variation of intensity o


_(f) -,-,
Trr = - p v, Vr

where .p is the time-average pressure


tube. We have again neglected entran
result. Integration and application of t

B.C. 1 : f~;'> is fini


yields an equation for the total time-av
11>
11>

~
11>

.,....
o
.e
(/) Thus, the total shear stress distribution
pressure drop.
Experimental studies of the turbulei

- (t)
Trz =-
0 .6 0.4 0.2 o have been made by a number of inve
-----rtr0 shown in Fig. 6.3-3. The variation of ii
Fl. 6.3-3. Total shear stress and turbulent shear stress for ftow Fig. 6.3-4. We find from these two fi
in a tube.
Turbulent Flow Chap. 6 Sec. 6.3 A Qualitative Description of Turbulent Flow 199

, we make four assumptions:


flow is one-dimensional;
uu'f'"'vu1 and the time-averaged con-

of Oand z;

(6.3-7)

side of Eq. 6.3-7 is independent of z


of r. We conclude that both sides are

(6.3-8)

1.0

- - - - - r;r0
Fig. 6.3-4. Variation of intensity of turbulence with radial position.

where !J.p is the time-average pressure difference and L is the length of the
tube. We have again neglected entrance and end effects in obtaining this
result. Integration and application of the boundary condition

B. C. 1: -r;;> is finite for O $;; r $;; r0 (6.3-9)


yields an equation for the total time-average shear stress

fCTJ
rz
= f
rz
+ fCrztl = - (tlp)
L 2
!_ (6.3-10)

Thus, the total shear stress distribution is determined by measurement of the


pressure drop.
Experimental studies of the turbulent shear stress,

(6.3-11)
have been made by a number of investigators, and illustrative results are
shown in Fig. 6.3-3. The variation of intensity of turbulence is illustrated in
turbulent shear stress for flow
Fig. 6.3-4. We find from these two figures that the turbulent shear stress
200 Turbulent Flow Chap. 6
Sec. 6.4 The Eddy Viscoslty

predomina tes in the core region and falls to zero at the wall. We also discover
that the maximum intensity of turbulence and turbulent shear stress is near
r/r0 = 0.9. We conclude that turbulence is generated near the wall and the
intensity falls off toward the center of the tube. In the central region, the
generating force (shear deformation) decreases and viscous forces tend to
reduce the intensity.

6.4 The Eddy Viscosity

An interesting intuitive description ofthe turbulent stress may be obtained


by proposing a constitutive equation for -t< 1> similar to Newton's law of
viscosity. In Chap. 5 the constitutive equation for an incompressible
Newtonian fluid was stated as
~ = 2,ud (6.4-1)
The time-average form is
- consider the steady (on the average), o
i = 2,ud (6.4-2) 6.4-1. If we assume that discrete "lum
By analogy (albeit poor), we might write one position to another and still retai
then the net rate of momentum transfe
(6.4-3)
x-momentum transferred
thus defining the turbulent or "eddy" viscosity, ,u<t>. The total time-averaged {across a y-surface
stress tensor is now given by
where the mass M of the two fluid pa
(6.4-4) incompressible. The turbulent shear s
which this transfer takes place, and the .
The eddy viscosity is not a very useful quantity for analysis, but it helps to
develop a qualitative feel for the nature of turbulent flow. The constitutive occurs. Thust,
force per) {momen
equation, Eq. 6.4-1, is a postulated relationship between the stress and the {unit area = unit tin
rate of deformation, and experiments show that ,u is indeed a constant for a and we write
large class of fluids. On the other hand, ,u<t> is nota constant but depends on
position and on the intensity of turbulence. In this section, we shall make use
of an empirical expression for ,u<t> to analyze the turbulent velocity profile in
a tube. where tl.t is the average time required f1
the average area associated with the t1
and tl.A are dependent on the structure
Prandtl's mixing length theory
tude and frequency of the velocity fluct
of a Taylor series expansion for v., give
Prandtl's mixing length theory 2 was one of the first attempts at a semi-
theoretical analysis of turbulent flow. The arguments leading to its develop-
ment are not very precise, but they are useful in developing a qualitative
v., ~+l
~v.,l
11

understanding of the mechanism of turbulent momentum transfer. We now t This expression follcws from the linear
2. L. Prandtl, "Investigations on Turbulent Flow," Zamm, 1925, S :136. tion is not obvious, and the student must wa!
is derived in Chap. 7 before he can understan
Turbulent Flow Chap. 6
Sec. 6.4 The Eddy Viscosity 201
to zero at the wall. We also discover
and turbulent shear stress is near
is generated near the wall and the
the tube. In the central region, the
decreases and viscous forces tend to

ofthe turbulent stress may be obtained


for i' 1l similar to Newton's law of
ve equation for an incompressible

Flg. 6.4-1. Turbulent momentum transfer.


2,ud (6.4-1)

- consider the steady (on the average), one-dimensional ftow illustrated in Fig.
2,ud (6.4-2)
6.4-1. Ifwe assume that discrete "lumps" or particles offtuid can move from
one position to another and still retain their momenta during this motion,
then the net rate of momentum transfer for the process shown in the figure is
(6.4-3)
x-momentum transferred} = Mv1 _ Mv1 (6.4-5)
{across a y-surface "' v+l "' 11

where the mass M of the two fluid particles must be the same if the ftow is
(6.4-4) incompressible. The turbulent shear stress, :;~~. will depend on the rate at
quantity for analysis, but it helps to which this transfer takes place, and the y-surface area over which the transfer
of turbulent flow. The constitutive occurs. Thust,
onship between the stress and the { umt
fo~ce per} =
area
{m~m~ntum tran~ferred per}
umt time per umt area .
(6.4-6)
that ,u is indeed a constant for a and we write
fL' 1l is nota constant but depends on
In this section, we shall make use -r-W M ( -~
11.,=-- v.,
- 1)
-v., (6.4-7)
the turbulent velocity profile in !l. t !l. A v+l 11

where !l.t is the average time required for the process to take place and !l.A is
the average area associated with the transfer. lt is easy to imagine that !l.t
and !l.A are dependent on the structure of the turbulence, i.e., on the magni-
tude and frequency of the velocity ftuctuations. Making use of the first term
of a Taylor series expansion for v., gives,
one of the first attempts at a semi-
arguments leading to its develop-
useful in developing a qualitative
v.,l v+l ~ v., + t dv)
(--= (6.4-8)
11 dy
Ient momentum transfer. We now
t This expression follcws from the linear momentum principie; however, the connec-
Flow," Zamm, 1925,5:136. tion is not obvious, and the student must wait until the macroscopic momentum balance
is derived in Chap. 7 before he can understand this result clearly.
202 Turbulent Flow Chap. 6 Sec. 6.5 Turbulent Flow in a Tube

and Eq. 6.4-7 becomes


6.5 Turbulent Flow in a Tube
.:<t> = Mt (dv.,) (6.4-9)
~t~A dy
11
"'
In this section, we will apply the
We can carry the analysis further by writing the mass M as problem of turbulent fl.ow in a tube. 1
gravity is neglected, the stress equatio
M=p~V (6.4-10)
and the volume ll V as .:<Tl
,. =_
~V= ct'~A (6.4-11)
Use of Eq. 6.4-4 to determine f~;> yie
where e is an unknown dimensionless constant. Substitution of Eqs. 6.4-10
and 6.4-11 into Eq. 6.4-9 gives

(6.4-12) lfwe examine the time-averaged rateo


(see Table 5.2-1) we find only two non
We can go no further, for there is no a priori knowledge of the constant e
or the time interval over which the transfer takes place. However, we do
know from experimental data that the turbulent shear stress tends to be large
when the velocity gradient is large, and small when the velocity gradient is and the eddy viscosity given by Eq. 6.
small. This fact, of course, does not hold at the wall where the velocity
gradient is a maximum and the turbulent shear stress is zero; yet the Prandtl
rnixing length theory in effect states

(~t) = ~:"' 1 1
(6.4-13) Substitution of Eqs. 6.5-3 and 6.5-4 in

and Eq. 6.4-12 takes the form

f~~~ = p?ll ~:"' (~:"') 1


(6.4-14) Following Prandtl, we let y be the dis

This equation defines the mixing length t in terms of experimentally deter-


y= l

rnined values of f~~ and (~;"'). The general tensor form of Eq. 6.4-14 is and let the mixing length be a linear 1

.:<t> = {2pt2.J id : <i}"d (6.4-15)


where a. will be chosen to give the b~
which leads to an eddy viscosity given by experimental data. The form chosen fi
however, it does satisfy the requireme
1-'(t) = pt2.Ji: d (6.4-16) zero at the wall. Noting that

If t'were a constant, or depended upon d in a simple manner, this result would r = r0 - y a


be very useful. However, neither is true, and the Prandtl mixing length theory
is of limited value. t See Eqs. 6.3-6 through 6.3-10.
Turbulent Flow Chap. 6 Sec. 6.5 Turbulent Flow in a Tube 203

6.5 Turbulent Flow in a Tube


(6.4-9)
In this section, we will apply the Prandtl rnixing length theory to the
writing the mass M as problem of turbulent fl.ow in a tube. If the time-averaged fl.ow is steady and
gravity is neglected, the stress equation of motion integrates tot
(6.4-10)
i'(T) = - (flp)~ (6.5-1)
rz L 2
(6.4-11)
Use of Eq. 6.4-4 to determine i'~;> yields
constant. Substitution of Eqs. 6.4-10

2(p + i'>) drz = - ( fli) ~ (6.5-2)

(6.4-12)
If we examine the time-averaged rate of strain tensor in cylindrical coordina tes
(see Table 5.2-1) we find only two nonzero components
a priori knowledge of the constant e
transfer takes place. However, we do
turbulent shear stress tends to be large
Jrz = Jzr = !(diJ.)
2 dr
(6.5-3)
small when the velocity gradient is
and the eddy viscosity given by Eq. 6.4-16 is
hold at the wall where the velocity
shear stress is zero; yet the Prandtl
#(t) = pfl J(~~) 2
(6.5-4)

(6.4-13) Substitution of Eqs. 6.5-3 and 6.5-4 into Eq. 6.5-2 yields

[# + pfl J(~~) ]~~ (!liH


2
= _ (6.5-5)
(6.4-14)
Following Prandtl, we let y be the distance measured from the wall,

t in terms of experimentally deter- y= r0 - r (6.5-6)

general tensor form of Eq. 6.4-14 is and let the mixing length be a linear function of y,

t= 1XY (6.5-7)
(6.4-15)
where IX will be chosen to give the best agreement between the analysis and
experimental data. The form eh osen for the mixing length is purely intuitive;
however, it does satisfy the requirement that the turbulent shear stress become
(6.4-16) zero at the wall. Noting that

d in a simple manner, this result would r = r0 - y and ~= - ~


and the Prandtl mixing length theory dr dy
t See Eqs. 6.3-6 through 6.3-10.
204 Turbulent Flow Chap. 6 Sec. 6.5 Turbulent Flow in a Tube

we may write Eq. 6.5-5 where r is a dummy variable of integr


integration C1 , we require that the veloci
(dv.) (dv.) !lp(r y)
2
0
2 2
..t- +pocy - -
=- -- - (6.5-8) entire region, O :::;; y :::;; r 0 Equating Eqs
dy dy L 2 us to evaluate the constant C1 , and the
which is a rather difficult, first-order, ordinary differential equation to solve, 6.5-12 and
requiring the use of either an analogue or a digital computer. However, we
may obtain a reasonably satisfactory solution by recognizing that the laminar
- ( Ap )1/2
V= - -
2Lpoc 2
f ~=U


(ro - r)1/2
r
dr
sublayer and turbulent core are two fairly distinct regions where the viscous ~=11

and turbulent stresses predominate, respectively. In what follows, we in


effect "shrink" the transition region to a cylindrical surface Jocated at
Let us review briefly our attack on th
y=y*. equations of motion, which for this case
We assume that there is a region close to the wall such that
a constitutive equation, Eq. 6.4-4, anda
p. ~ p,<tl for O :::;; y ::::;; y* (6.5-9) for the eddy viscosity. We obtained a
the velocity v., which contained a single
where y* ~ r 0 In this region, Eq. 6.5-8 becomes, approximately,
undetermined distance, y*.
P.(dv,) = !lfJ ~, (6.5-10)
A comparison of the solution to t
mental dat~ perhaps could be used to t
dy L 2
and to specify the parameters oc and y*;
Integration and application of the boundary condition eters it is quite likely that even an unsa
B.C. 1: v. =O, y=O (6.5-11) experimental data reasonably well.
To obtain a solution to this equati
yields the solution for the velocity profile near the wall.
effort, we need to evaluate the integral iJ
i5

= (Afi)'oY
L 2..t'
o::::;; y ::::;; y* (6.5-12)
cally or graphically for any value of y;

We now assume that


we can obtain an algebraic solution fe
for y* :::;; y ::::;; r0 (6.5-13)
experimental data. This approximation
which, of course, implies a discontinuity in the eddy viscosity at y =y*. the tube, but it is not a bad approxirr
Such a discontinuity is not allowed on physical principies; however, p,<tl does turbulent core. For example, if r = 0.3
change very rapidly in the transition region, and it is not too unsatisfactory to (ro- r)1/2 = r~/2(0.7)1
approximate this change by a discontinuity.
In the second region we write Eq. 6.5-8 as Since this value of r takes us well into l
solution to predict satisfactorily the ve
22
poc y dy
(di5,)2 = (AfJ)
L ('o- y)
-2- ' y* ::::;; y ::::;; ro (6.5-14a) turbulent core. On the basis of Eq. 6.5-

or (!lp)
v=
'
roY
L 2..t'
dv, __ [~(.cc__r)J1/2, y* ::::;; y ::::;; ro (6.5-14b)
!lp )112 (y ) (!lp)
dy 2Lpoc 2 l v. = ( 2Lpoc2 ln y* + L .
lntegration gives

- ( AfJ )1/2
v, = - -
2Lpoc 2
f ~ =V


(ro - r)1/2
r
dr + C, y* ::::;; y ::::;; r 0 (6.5-15)
We now define the wall shear stress as,
!lp ro
'To = - - =
~ =U L 2
Turbulent Flow Chap. 6
Sec. 6.5 Turbulent Flow in a Tube 205

where r is a dummy variable of integration. To determine the constant of


integration C, we require that the velocity be a continuous function over the
(6.5-8) entire region, O ~y ~ r0 Equating Eqs. 6.5-12 and 6.5-15 at y= y* allows
us to evaluate the constant C1 , and the velocity profile is then given by Eq.
ordinary differential equation to solve, 6.5-12 and
or a digital computer. However, we
by recognizing that the laminar
distinct regions where the viscous
- ( .p )1/2
v. = - -
2Lpt:~.
2
J ~=11


(ro - r)l/2
r
dr + (.ft)
-L -- roy*
2.
'
~=11
respectively. In what follows, we in
to a cylindrical surface located at (6.5-16)
Let us review briefly our attack on this problem. We started with the stress
to the wall such that equations of motion, which for this case reduce to Eq. 6.3-8. We introduced
a constitutive equation, Eq. 6.4-4, anda semitheoretical equation, Eq. 6.4-16,
(6.5-9) for the eddy viscosity. We obtained a differential equation, Eq. 6.5-8, for
beco mes, approximately, the velocity v,, which contained a single undetermined parameter IX, and the
undetermined distance, y*.
(6.5-10) A comparison of the solution to this differential equation with experi-
mental data perhaps could be used to test the Prandtl mixing-length theory
condition and to specify the parameters t:1. and y*; however, with two adjustable param-
eters it is quite likely that even an unsatisfactory theory could be made to fit
(6.5-11) experimental data reasonably well.
near the wall. To obtain a solution to this equation without expending a great deal of
effort, we need to evaluate the integral in Eq. 6.5-16. We can do so numeri-
o:;:;; y:;:;; y* (6.5-12) cally or graphically for any value of y; however, by assuming
r) 112 ~::::~ r~1 2
(r 0 - (6.5-17)
y* s y:;:;; r0 (6.5-13) we can obtain an algebraic solution for the velocity and compare it with
experimental data. This approximation cannot be valid near the center of
in the eddy viscosity at y = y*. the tube, but it is not a bad approximation over a limited portion of the
physical principies; however, .<t> does turbulent core. For example, if r = 0.3r0 , then
and it is not too unsatisfactory to
12
(r 0 - r) ' = r~' (0.7) 1 ' 2 = 0. 84r~ 12 ~::::~ r~12
2
nuity. (6.5-18)
.5-8 as Since this value of r takes us well into the turbulent core, we can expect the
solution to predict satisfactorily the velocity profile from the wall into the
y* :;:;; y :;:;; ro (6.5-14a) turbulent core. On the basis of Eq. 6.5-17, the velocity is given by

v, = (6.i) ~;, O ~y ~y* (6.5-19a)


y* :;:;; y :;:;; ro (6.5-14b)
In (L) + (.p)roy*, y* ~y ~ 0.3r
112
v = ( .ft 2) (6.5-19b)
2Lpt:~. y* L 2. 0

We now define the wall shear stress as,


y* :;:;; y :;:;; r0 (6.5-15)
7'O -_ .ft ~ = --r-<T>j
r (6.5-20)
L 2 r=ro
Sec. 6.6 Relative Magnitude of Molecular
206 Turbulent Flow Chap. 6

Thus, T 0 is the shear stress that the fluid exerts on the wall. Noting that
,.) To/ p has the units of velocity, and that v, the kinematic viscosity, has the
units oflength times velocity, we may define a dimensionless velocity, v+, as

(6.5-21)

anda dimensionless distance, y+, as

(6.5-22)

Equations 6.5-19a and b are now written in dimensionless formas


(6.5-23a)

y+
(6.5-23b)
where Fig. 6.5-1. Comparison ofPrandtl mix

Y=--p
yt data.

p
6.6 Relative Magnitude of Mol e
These results may be compared with the experimental work of Deissler
Eddy Viscosity
and Laufer4 for turbulent velocity distributions in tubes. The results are
shown in Fig. 6.5-1 for ex= 0.36 and yt = 10.0. We see that for y+< 5.0,
Eq. 6.5-23a represents the data quite well. This region is that of the laminar We may gain sorne insight into ti
sublayer where 1-' ~ .t<tl. The analysis also agrees well with the data for momentum transport by examining the
y+> 25. We may assume, therefore, that the transition zone occupies the is best done by starting with
region 5 < y+ < 25, and we did not expect to get good agreement in this
region. -(T) _
7 rz -
-
7 rz
+ -(t) _
Trz -
More extensive and detailed treatments of this problem are available
elsewhere.&-6 Most important is that at present there exists no theoretical and rearranging to obtain
treatment of turbulent flow which would provide a generally valid constitutive
equation relating the turbulent shear stress, i<tl, to the time-average rate of
strain tensor, d. For this reason, we must always make use of sorne experi-
mental information to solve turbulent flow problems.
3. R. O. Deissler, NACA Tech. Note 3016, 1953.
4. J. Laufer, NACA Tech. Note 2954, 1953. By Eqs. 6.3-10 and 6.5-20, the stress ma
5. R. B. Bird, W. E. Stewart, andE. N. Lightfoot, Transport Phenomena (New York:
John Wiley and Sons, lnc., 1960), Chap. 5.
6. H. Schlichting, Boundary Layer Theory, 4th ed. (New York: McGraw-Hill Book [-rz + -(t)] = -
7 Trz
!::.
L
Company, Inc., 1960), Chap. 20.
Turbulent Flow Chap. 6
Sec. 6.6 Relative Magnitude of Molecular and Eddy Viscosity 207

fluid exerts on the wall. Noting that


that v, the kinematic viscosity, has the 1
1
:
1
define a dimensionless velocity, v+, as
25r-~---1--~--~~r-~--~--~--~--+-~-~o0~~
0
1
~~
(6.5-21)
1
1
1--rurbulent
1 core
ooVc~
i o~~
20r-~---+--+-~~--+-~~-h~~~~--~-+--~~
1i
14-!.rons1't'1on ~>Y 11

l region: .,Y
15r-~----r--+--~~~~~~~--~L_~---+----~
v+ e _j~ '- Eq 6.5-23b
,.__Lammar,...Ji a..-"'..>''"

in dimensionless form as
(6.5-22) 10
sublayer

ligo 1/"

sr-,_--7r~e--+1-~--~--+---+-~~-+---+~
!
1

(6.5-23a)
1
_v: 1
"'-....Eq. 6 5-2~a
~ !
Y ~y+~ !:
0.3r0

V
p
(6.5-23b)
2 5 10
11
20 25 50
1

y+
100 200 500 1000 2000 5000

Fig. 6.5-1. Comparison of Prandtl mixing length theory with experimental

y*!: data.

V
6.6 Relative Magnitude of Molecular and
the experimental work of Deissler3
Eddy Viscosity
ons in tubes. The results are
y~ = 10.0. We see that for y+< 5.0,
. This region is that of the laminar We may gain sorne insight into the relative magnitude of turbulent
also agrees well with the data for momentum transport by examining the ratio /i(t) ffi for flow in a tu be. This
that the transition zone occupies the is best done by starting with
expect to get good agreement in this
-(T) _
Trz -
-
Trz
+ Trz-(t) _
-
[
fi
+ fi (t)] diJ.
-- (6.6-1)
of this problem are available dr
at present there exists no theoretical and rearranging to obtain
provide a generally valid constitutive
-t<t), to the time-average rate of J/(t)
r__ = _1 [f'rz + ;<t)]
rz _
1 (6.6-2)
always make use of sorne experi-
flow problems. fi fi e~)
1953.
By Eqs. 6.3-10 and 6.5-20, the stress may be written as

4th ed. (New York: McGraw-Hill Book


- + Trz
[Trz -W] -__ Aj!.. -_ _ To (!._) (6.6-3)
L 2 r0
208 Turbulent Flow Chap. 6 Problems

Changing variables from r to y and substituting Eq. 6.6-3 into Eq. 6.6-2, we
have PRO
p.<t> 1 -r0 [1 - (y/r 0 )]
-=- -1 (6.6-4) 6-1. Given a velocity
ft ft (~;) Vz =

lt will be helpful if we put this result in dimensionless form to make use of where b and e are constants ha
compute the time-average veloc
our previously derived results. range of values !lt may take (in t
p.<t> [1 -(y/r0)] make the statement that Vz is "e
-= -1 (6.6-5) this by comparing the average o
p. (dv+)
dy+
requiring that they be within l pe
Ans: !lt ~ [0.03(1 + bt + ct2)/c]
Equation 6.5-23b may be differentiated to give
6-2. Prove that
dv+ 1
-=--, y+ 10 (6.6-6)
dy+ 0.36y+
2/lt
We now choose y to be r 0 J4, so that Eq. 6.6-6 will be valid and the point at
which the eddy viscosity is evaluated is within the turbulent core. By Eq. by expanding the integrand into
6.5-22,
6-3. Prove that

y ~ (~)~
V
(6.6-7)
for incompressible flows by expa1
equal. The expression for the di
and we may write Eq. 6.6-5 as Note that the scalar components

~
6-4. Put Eq. 6.2-24 in dimensionless

#~) = ()(0.36)()~-
(exclusive of those occurring in
1 (6.6-8) u pon.

To examine this result for a particular case, we must make use of sorne infor- 6-5. Apply dimensional analysis to 1
shown in Fig. 6-5. Show that th
mation discussed more thoroughly in Chap. 8. There we find that the shear
angle oc. Thus one can expect the
stress at the wall is given approximately by depend upon both NRe and oc.
0.039p<v.)2 3 5
-r0 = for 2.1 X 10 :::::;: NRe :::::;: 10 (6.6-9)
114
N Re
Substituting Eq. 6.6-9 into Eq. 6.6-8 and carrying out the arithmetic, we find
(t)
!!:...... ~ NU! X 10-2 (6.6-10)
ft
F or Reynolds numbers inthe range from 103 to 105 , the ratio of eddy viscosity
to molecular viscosity ranges from 10 to 103. From this example, we learn
x=O
that molecular momentum transport is generally negligible compared to the
turbulent momentum transport in the core region. F
Turbulent Flow Chap. 6 Problems 209
substituting Eq. 6.6-3 into Eq. 6.6-2, we
PROBLEMS
(6.6-4)
6-1. Given a velocity

where b and e are constants having units of sec- 1 and sec- 2 respectively,
compute the time-average velocity by means of Eq. 6.1-1. Indicate what
range of values !1t may take (in terms of b and e) so that it is permissible to
make the statement that z is "essentially constant in the period 2M." Do
(6.6-5)
this by comparing the average of the average Vz with the average f. and
requiring that they be within 1 per cent of each other.
Ans: M ::::; [0.03( l + bt + et 2)fe]1/2
to give
6-2. Prove that
t+!.t
y+ 10 (6.6-6)

Eq. 6.6-6 will be valid and the point at


-f
1
-2M v' VV dt
t - !. t
=O

is within the turbulent core. By Eq.


by expanding the integrand into the scalar components and integrating.
6-3. Prove that
v Vv = V (vv)
(6.6-7)
for incompressible flows by expanding both sides and showing that they are
equal. The expression for the divergence of a tensor is given by Eq. 4.4-11.
Note that the scalar components of the tensor vv are v.,u.,, u.,v11 , etc.
6-4. Put Eq. 6.2-24 in dimensionless forro in order to determine what parameter
(6.6-8) (exclusive of those occurring in the boundary conditions) v should depend
u pon.
case, we must make use of sorne infor- 6-5. Apply dimensional analysis to the flow in the diverging circular conduit
Chap. 8. There we find that the shear shown in Fig. 6-5. Show that the velocity depends upon both NRe and the
by angle ex. Thus one can expect the transition from laminar to turbulent flow to
depend upon both NRe and ex.
(6.6-9)

and carrying out the arithmetic, we find

(6.6-10)

103 to 105, the ratio of eddy viscosity


10 to 103 From this example, we learn
is generally negligible compared to the
210 Turbulent Flow Chap. 6

6-6. Put Eq. 6.5-8 in dimensionless form in terms of v+ and y+ and integrate
numerically for values of ex ranging from 0.3 to 0.5. Compare the results with
the approximate solution given in the text.
Note: This problem requires sorne careful analysis, in addition to the use of a
digital computer.
6-7. Using the approximation Macroscopic B
r )1/7
Vz = Vz,max (1 - ;;;
lnert1
for water at 70F flowing in a smooth 2-in. diameter tube at an average
velocity of 6 ft/sec, compute
(a) The total shear stress at r/r0 = 0.9
(b) The fraction of this shear stress dueto the turbulent stress.
Note: Use Eq. 6.6-9 to compute the wall shear stress, T 0
6-8. Water at 70F is flowing in a 6-in. diameter pipe with an average velocity of
48 ft/sec. Find the thickness of the laminar sublayer.
6-9. Assuming Eq. 6.5-23b can be used to determine the velocity from the laminar
sublayer to the center of the tu be,

derive a relationship between the friction factor f and the Reynolds number
NRe Neglect the flow in the region O :::; y+ :::; y in carrying out this calcu-
PART
lation, and simplify your result for large values of r0 qpjv. Express the
friction factor as
f = 4T0 ftp ( v.) 2 In Chap. 5, we formulated th
6-10. Derive the time-averaged form of the continuity equation for compressible applied them to sorne simple lam
flow. Keep in mind that the density may be a function of both time and the turbulent flow, and it became appa
spatial coordinates. tions of the time-averaged equations
flows. In this chapter, we will form
momentum, and mechanical ener~
macroscopic equations will provide
average." Thus, we will be able to
while the pressure distribution and
ftow topology-will remain undet<
derive Bernoulli's equation giving
approximate solutions to complex f
After deriving these equations, 1
where inertial effects predominate
results. The application of the mac
effects must be considered will be co
able to split up our study of the ma
Turbulent Flow Chap. 6

in terms of v+ and y+ and integrate


0.3 to 0.5. Compare the results with
text.
analysis, in addition to the use of a

Macroscopic Balances:
Inertial Effects 7
2-in. diameter tube at an average

wall shear stress, T0

ldtm~:ter
pipe with an average velocity of
laminar sublayer.
determine the velocity from the laminar

factor f and the Reynolds number


O :::; y+ :::; yt in carrying out this calcu-
PART I-THEORY
large values of r0 qpjv. Express the

In Chap. 5, we formulated the differential equations of motion and


the continuity equation for compressible applied them to sorne simple laminar ftows. In Chap. 6, we investigated
may be a function of both time and the turbulent flow, and it became apparent that we could not obtain exact solu-
tions of the time-averaged equations of motion for even the simplest turbulent
flows. In this chapter, we will formulate the macroscopic ( or integral) mass,
momentum, and mechanical energy balances. An exact solution of these
macroscopic equations will provide us with results which are correct "on the
average." Thus, we will be able to compute total forces and total flow rates
while the pressure distribution and velocity profile-i.e., the details of the
ftow topology-will remain undetermined by the analysis. We shall also
derive Bernoulli's equation giving us another useful equation for obtaining
approximate solutions to complex ftow problems.
After deriving these equations, we shall apply them to a variety of flows
where inertial effects predominate but cannot completely explain observed
results. The application of the macroscopic balances to flows where viscous
effects must be considered will be covered in Chap. 8. It may seem unreason-
able to split up our study of the macroscopic balances in this way; however,
211
212 Macroscopic Balances-lnertial Effects Chap. 7 Sec:. 7.1 The Macroscopic Mass Balance

we encounter two distinct problems in applying these balances: determination


of ftuxes of mass, momentum, and energy at the entrances and exits of the
control volume; determination of the forces which act upon the surfaces of
the control volume. In Chap. 7, we deal with the first problem in addition to
treating pressure forces, while the detailed discussion of the second is left
until Chap. 8.
The student should be forewarned that the methods to be studied in this
chapter, and subsequent ones, are approximate; in general, there will be no
"right" answers. There will often, however, be a "best" answer, andas often
as possible we shall try to determine the best answer by comparing our results
with experiments. In attacking this chapter, we should remember that the
macroscopic balances are perhaps the most powerful too! the engineer pos-
sesses for solving the often ill-defined problems of everyday practice. Judi-
cious application of these equations comes only with experience and practice.
At best, the student can hope to understand the development of the equations
and gain sorne insight regarding the difficulties that may be encountered in
their application.

*7.1 The Macroscopic Mass Balance

To develop a completely general form of the macroscopic mass balance,


we will apply the principie of conservation of mass to the arbitrary volume
"''a(t) illustrated in Fig. 7.1-1. Remember that points on the surface of this Fl 1 7.1-1.
volume move with a velocity w which may be a function of the spatial co-
ordinates and time. The continuity equation, We now make use of the general tr
op in Eq. 3.4-16 be the density, p, whic
-+V (pv) =O (7.1-1)
ot
holds at every point in space; thus, we may integrate Eq. 7.1-1 o ver the volume ~t f pdV=
Y'".(1)
f
Y'".(1)
"''a(t) to obtain

f [::+V
Y'"o(l)
(pv)J dV =O (7.1-2)
1t should be clearly understood that
and is not related to the physical 1
volume integral of (opfot) in Eq. 7.1
This equation, in effect, represents the macroscopic mass balance, but it is to give
of little use to us in this form. The following steps are not aimless mathe-
matical manipulation but necessary steps in the development of an equation ~tf pdV=- f p
Y'".(l) .<:1.(1)
useful in solving problems.
Application of the divergence theorem allows us to write the second term Putting the two terms on the right-
under the integral as an area integral to obtain and rearranging the equation, we ge

(7.1-3) ~tf pdV+ f .


Y'".(l) ""
ic Balances-lnertial Effects Chap. 7 Sec. 7.1 The Macroscopic Mass Balance
213
applying these balances: determination w
at the entrances and exits of the
forces which act upon the surfaces of
with the first problem in addition to
discussion of the second is Ieft

that the methods to be studied in this


uJI.luutu;;; in general, there will be no
, be a "best" answer, and as often
best answer by comparing our results
, we should remember that the
most powerful too! the engineer pos-
problems of everyday practice. Judi-
only with experience and practice.
the development of the equations
,_, ..... ___ ,.,. that may be encountered in

of the macroscopic mass balance,


of mass to the arbitrary volume
that points on the surface of this Flg. 7.1-1. Arbitrary moving control volume.
may be a function of the spatial co-
We now make use of the general transport theorem by letting the scalar S
in Eq. 3.4-16 be the density, p, which gives
(7.1-1)

integrate Eq. 7.1-1 over the volume ~~ f p dV =


~.(1)
f
~.(1)
(~~) dV + f pw n dA
Jllf.(t}
(7.1-4)

lt should be clearly understood that Eq. 7.1-4 is a mathematical relationship


(pv)J dV= O (7.1-2)
and is not related to the physical principie of conservation of mass. The
volume integral of (opfot) in Eq. 7.1-3 may now be substituted into Eq. 7.1-4
macroscopic mass balance, but it is to give

~~f pdV=- JpvndA +JpwndA


tUIIIuwu"steps are not aimless mathe-
in the development of an equation (7.1-5)
~.(1) .rt!.(t) Jllf.(t)

allows us to write the second term Putting the two terms on the right-hand side under the same integral sign,
and rearranging the equation, we get

JpvndA =O (7.1-3) ~~ Jp dV +Jp(v - w) n dA = O (7.1-6)


~.(1) ~.(1) Jllf.(t)
214 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.1 The Macroscopic Mass Balance

It will be convenient to separate the area da(t) into three areas: where A.(t) has been replaced by
l. A.(t), the area of entrances and exits through which fluid may enter volume is fixed in space. Because
and leave the control volume; in this case, the total derivative ma
2. As(t), the area of solid moving surfaces; order of differentiation and integra
3. A., the area of solid fixed surfaces.
Since v - w = O on the solid moving surfaces, and v = w = O on the solid
fixed surfaces, Eq. 7.1-6 may be written in final formas
..,.J(ap)dv-
at
For incompressible flows, (apjat) =
~tf pdV+J p(v-w)ndA=O (7.1-7)
..Y0 (l) A,(t)

Although this process has been relatively simple, it is helpful in understanding


J
A,
pv n dA= O

the development to review the steps. The velocity w is arbitrary and


l. The physical principie is stated in Eq. 7.1-1. The area integral in Eq. 7.1-7 is th
2. Integration is performed to obtain a macroscopic equation, Eq. 7.1-2: -
3. The term V (pv) in Eq. 7.1-2 would be rather difficult to evaluate for
the problems we have in mind, so we applied the divergence theorem
f
~ pdV=
dt -r.(t)
to yield an area integral of pv n. This term represents a mass flux
However, the arbitrary volume "Y
and is very often precisely the quantity that is known or must be
Eq. 7.1-12 should be written
determined.
4. Similarly, the term (apat) in Eq. 7.1-2 would be difficult to evaluate. D
Application of the general transport theorem yielded another mass Dt
1
flux term and the time rate of change of the mass in control volume,
which is often the quantity to be calculated or is specified by the nature which is our original statement ofi
of the problem.
The point is that practica! reasons exist for arranging the macroscopic Removal of liquid from a cylind
balance in the form indicated by Eq. 7.1-7.
In words, Eq. 7.1-7 may be written as Asan illustration ofthe applica
shown in Fig. 7.1-2. A rod of dia1
time rate of change} {mass flux } {mass flux }
( of the mass of the - into the + out of the = O (7.1-8) hole of diameter D1 and length D
control volume control volume control volume and the flow is to be treated as in
which is the same statement used in Sec. 3.5 for the alternate derivation ofthe sion for the average velocity in
contin~ity equation. Equation 7.1-8 is valid for any arbitrary volume; thus, constant, Eq. 7.1-7 may be writte
it naturally applies to the differential cube illustrated in Fig. 3.5-2.
d"Ya(t) +
dt A
Special forms
The only entrance or exit is the cr
Very often we will want to apply the macroscopic mass balance to a At that surface, the outwardly dir
control volume fixed in space. Under this condition, w = O and we write
d
-
dt ,
f p dV + f pv n dA= O
Control volume
fixed in space (7.1-9)
and since w = O, we write
-v"
A, (v- w) n = v
ic Balances-lnertial Effects Chap. 7 Sec. 7.1 The Macroscopic Mass Balance 215

where A.(t) has been replaced by A. and "Ya(t) by "Y to indicate that the
volume is fixed in space. Because the spatial coordinates are held constant
in this case, the total derivative may be written as a parta! derivative and the
order of differentiation and integration changed to give

(7.1-10)
surfaces, and v = w = O on the solid
in final form as
For incompressible flows, (apjat) =O, and Eq. 7.1-10 reduces to
- w)ndA =O (7.1-7)
lncompressible flow,

y simple, it is helpful in understanding


f
.A.
pv n dA = O control volume
fixed in space
(7.1-11)

The velocity w is arbitrary and we may set it equal to the fluid velocity, v.
in Eq. 7.1-1. The area integral in Eq. 7.1-7 is then zero, and we obtain
a macroscopic equation, Eq. 7.1-2;-
be rather difficult to evaluate for ~ Jp dV = O for w = v (7.1-12)
so we applied the divergence theorem dt f.(l)
n. This term represents a mass flux However, the arbitrary volume "Ya(t) is now a material volume "Y m(t), and
quantity that is known or must be
Eq. 7.1-12 should be written
7.1-2 would be difficult to evaluate.
theorem yielded another mass
Q_
Dt
f p dV= O
fm(l)
(7.1-13)

change of the mass in control volume,


which is our original statement of the principie of conservation of mass.
calculated or is specified by the nature

exist for arranging the macroscopic Removal of liquid from a cylindrical hole
7.1-7.
Asan illustration ofthe application of Eq. 7.1-7 we will analyze the system
shown in Fig. 7.1-2. A rod of diameter D 0 is being forced into a cylindrical

} + {:s~:~xe
control volume
}= O (7.1-8) hole of diameter D 1 and length L. The velocity of the rod is constant at u0
and the flow is to be treated as incompressible. We wish to derive an expres-
3.5 for the alterna te derivation of the sion for the average velocity in the annular region. Since the density is
val id for any arbitrary volume; thus, constant, Eq. 7.1-7 may be written as

+ J(v -
illustrated in Fig. 3.5-2.
d"Ya(t) w) n dA = O (7.1-14)
dt .A.(I)

The only entrance or exit is the cross-sectional area of the annulus at z = O.


the macroscopic mass balance to a At that surface, the outwardly directed unit normal to the control volume is
this condition, w = O and we write
n= -k (7.1-15)
Control volume and since w = O, we write
o fixed in space (7.1-9)
(v- w) n = v (-k) = -v. at z =O (7.1-16)
216 Macroscopic Balances-lnertial Effects Chap. 7 S.C. 7.2 The ~acroscoplc Momentum Bal

7.2 The Macroscopic Mome1

In deriving the macroscopic m


begin with the general statement of tJ

.!!.... J
pvdV=
Dt 7'",.<t> -r
Representing the stress vector in tern
Reynolds transport theorem (see Eq.

J ~a
j'".. <t> t
(pv) dV + J
JI(..
pvv n
<t>
Note that the term vv n can be looke
a vector n, or a vector v multiplied b.
equivalent. From a physical point o
- - - - - z = u0 t pvv n dA as the momenturo per uni
rate, v n dA ; thus, this term represe
unit volume times volume per unit ti
flux tensor.
Applicatioil of the divergence th
7.2-2 allows us to put all the terms
readily obtain the stress equations o~

Flg. 7.1-2. Removal of fluid from a cylindrical hole. ~~ (pv) + V (p


Formation ofthe integral ofEq. 7.2-3
Substitution ofEq. 7.1-16 into Eq. 7.1-14 and expression ofthe area integral scopic momentum balance
in terms of the average velocity times the area give
J: (pv)dV+ J V(pvv)
(v.)[ (D~ -
1
d:;(t)- D~ J=O (7.1-17)
j'". (t)

As with the first step in the macro!


j'". (t)

equation are difficult to analyze from


Substitution of the expression for "Ya(t), this problem, we will use the diverge;
the momentum flux tensor and the s
"Ya(t) = !( 1T D~L- 1T D~u 0 t) (7.1-18) theorem to replace the first term in Eq
into Eq. 7.1-17 and rearrangement yield the average velocity in the annular to analysis.
region, Application of the divergence the<
gives
(7.1-19)
J ~/pv)
j'".(t)
dV +Jpvv n
Ji'/,(t)

t See Prob. 4-15 for the derivation of the


Balances-lnertial Effects Chap. 7 S.C. 7.2 The Aacroscopic Momentum Balance 217

7.2 The Macroscopic Momentum Balance

In deriving the macroscopic momentum balance, it is convenient to


begin with the general statement of the linear momentum principie:

~~ f pv dV =
Y'",.(t)
f pg dV +
Y'",.(1)
f t<a> dA
J1ff,.(1)
(7.2-1)

Representing the stress vector in terms of the stress tensor and applying the
Reynolds transport theorem (see Eq. 3.4-18) gives

J :t
Y'",.(1)
(pv) dV + J pvv n dA
J1ff,.(1)
= J pg dV
Y'",.(1)
+ J T. n dA
J1ff,.(1)
(7.2-2)

Note that the term vv n can be looked upon as either a tensor vv dotted with
a vector n, or a vector v multiplied by a scalar v n; both interpretations are
equivalent. From a physical point of view, it is best to interpret the term
- - - - - z = u0 t pvv n dA as the momentum per unit volume, pv, times the .volumetric flow
rate, v n dA; thus, this term represents a momentum flux (momentum per
unit volume times volume per unit time), and pvv is often called the momentum
flux tensor.
Applicatioh of the divergence theoremt to the two area integrals in Eq.
7.2-2 allows us to put all the terms under the same integral sign, and we
readily obtain the stress equations of motion.

:t (pv) + V (pvv) = pg + V T (7.2-3)

Formation ofthe integral ofEq. 7.2-3 over the volume "Ya(t) yields the macro-
4 and expression of the area integral scopic momentum balance

f :t
7"".(1)
(pv) dV + f V (pvv) dV =
7"".(1)
f pg dV
7"".(1)
+ f V. T dV
7"".(1)
(7.2-4)
(7.1-17)
As with the first step in the macroscopic mass balance, the terms in this
equation are difficult to analyze from a practica] point of view. In attacking
this problem, we will use the divergence theorem to obtain area integrals of
the momentum flux tensor and the stress vector, and the general transport
(7.1-18)
theorem to replace the first term in Eq. 7.2-4 with something more susceptible
the average velocity in the annular to analysis.
Application of the divergence theorem to the second and fourth integrals
gives
(7.1-19)
f
7"".(1)
~~ (pv) dV + f pvv n dA
.#.(1)
= f pg dV
7"".(1)
+ f T n dA
Jif.(t)
(7.2-5)

t See Prob. 4-15 for the derivation of the divergence theorem for a tensor.
218 Macroscopic Balances-lnertlal Effects Chap. 7 Sec. 7.2 The Macroscopic Momentum

We now employ the general transport theorem (see Eq. 3.4-16) to provide the When viscous effects are important,
mathematical relationship be determined on the basis of expe1
drag coeffi.cients. However, a num
~ J
dt 7'".<t> ,
pvdV= J ~(pv)dV+ J
7'".(1) at
pvwndA
.llf.<ll
(7.2-6) reasonably neglect viscous effects. S
changes of velocity in the directio
which we may use in conjunction with Eq. 7.2-5 to obtain equations of motion predominate.
contraction in a pipeline, or the im]
~~ Jpv dV +Jpv(v- w) n dA= Jpg dV +Jt(nl dA (7.2-7)
cases where viscous effects may
obtained. For such problems the stJ
7'".<1) .llf,.(t) 7'".<t> .llf.(1)
t(D)
Here, we have made use of and the macroscopic mass and mo1
Tn=t<n> (7.2-8) average velocities and pressures.
We may write the area integral of the momentum flux as
The differential stress equations
J
d
pv(v - w) n dA = J pv(v - w) n dA
.A,(t)
(7.2-9)
Asan application ofEq. 7.2-10, ~
0 (1)
stress equation of motion given p
and the final form of the macroscopic momentum balance results.t seem to be overly repetitious inas
Time rate of change Net flux of momentum ential equation to obtain the macro
of the momentum of leaving the control
the control volume volume of the differential equation by appli

~ Jpv dV +Jpv(v -
fixed in space provides an excellent <
w) n dA fluxes. Past experience has shown
dt 7'".(1) .A,(t)
fine test of the students' understanq
(7.2-10)
= f pg dV
7'"a(l)
+ f t<nl dA
.llf4 (1)
We consider a differential cube
Figs. 7.2-la and b. For a control
x-component of Eq. 7.2-10 reduces
Body force Surface force

An especially useful form of Eq. 7.2-10 is that for steady flow in a control ~ J pv.,dV+ J
volume fixed in space. Under these conditions, we have
ot 7'" .A

J
A,
(pv)v n dA = J
7'"
pg dV + J
.llf
t<nl dA
Steady flow,
control volume
fixed in space
(7.2-11) =j
1

In Fig. 7.2-1a, the surface stresses a:


For turbulent flows, we simply carry out the preceding analysis using the
shown as positive stresses. In Fig. 7.
time-averaged stress equations of motion. The final result is obtained by
sented by arrows the direction of w
replacing v and t<n> by the time-averaged quantities, v and t<n> where
presuming Vr, v11 , and v. are all pe
i<n> = -np + n [.:r + _:r<t>] (7.2-12) represents the convection of x-dirc
The body force term in Eq. 7.2-10 is easily evaluated, and the two arrow pointing out of the cube rep1
integrals on the left-hand side can usually be represented with a fair degree of tum out of the cube.
accuracy by means of average velocities, volumes, and cross-sectional areas. Returning to Fig. 3.5-2, which i
of a cube, we see that the momer
t Note that the phrase momentum of the control volume should properly be interpreted x-direction momentum per unit vo
as momentum of the fluid within the control volume.
balance involves the convection of
Chap. 7 Sec. 7.2 The Macroscopic Momentum Balance 219

the1ore1m (see Eq. 3.4-16) to provide the When viscous effects are important, the area integral of the stress vector must
be determined on the basis of experimental or empirical friction factors or
drag coefficients. However, a number of problems exist for which we may
dV +Jpvw n dA (7.2-6) reasonably neglect viscous effects. Such problems arise when there are large
""(1) changes of velocity in the direction of flow and the inertial terms in the
Eq. 7.2-5 to obtain equations of motion predominate. Flow through a sudden expansion or
contraction in a pipeline, or the impinging of a liquid jet on a flat plate are
dA= J
j"
pg dV +
.<11
J"".<11
t<n> dA (7.2-7) cases where viscous effects may be neglected and reasonable results still
obtained. For such problems the stress vector is written as
t(n) = -np (7.2-13)
and the macroscopic mass and momentum balances are solved in terms of
(7.2-8)
average velocities and pressures.

The differential stress equations of motion


J pv(v - w) n dA
A.(l)
(7.2-9)
Asan application ofEq. 7.2-10, we will rederive the x-direction differential
stress equation of motion given previously in Chap. 4. This example may
seem to be overly repetitious inasmuch as we have just integrated the differ-
ential equation to obtain the macroscopic balance. However, the derivation
of the differential equation by application of Eq. 7.2-10 toa differential cube
fixed in space provides an excellent opportunity to compute sorne momentum
- w) odA
fluxes. Past experience has shown that this particular example provides a
(7.2-10) fine test of the students' understanding of Eq. 7.2-10.
We consider a differential cube, fixed in space, such as those shown in
Figs. 7.2-1a and b. For a control volume fixed in space, w =O and the
x-component of Eq. 7.2-10 reduces to
Surface force

is that for steady flow in a control


we have
q_
ot r
J pv.,dV+ J
A.
pv.,vndA
(7.2-14)
Steady fiow,
control volume (7.2-11) = J pg., dV J
+ t<nl~~: dA
fixed in space r ""
out the preceding analysis using the In Fig. 7.2-la, the surface stresses are all drawn in the x-direction, and are all
The final result is obtained by shown as positive stresses. In Fig. 7.2-1 b, the momentum flux terms are repre-
quantities, v and t<nl where sented by arrows the direction of which indicates the direction of mass ftow,
presuming vz, v11, and v. are all positive. An arrow pointing into the cube
n . [.:r + .:r<t>] (7.2-12) represents the convection of x.direction momentum into the cube and an
O is easily evaluated, and the two arrow pointing out of the cube represents convection of x-direction momen-
be represented with a fair degree of tum out of the cube.
volumes, and cross-sectional areas. Returning to Fig. 3.5-2, which indicates the mass convected into and out
of a cube, we see that the momentum balance involves the convection of
volume should properly be interpreted
x-direction momentum per unit volume pv., in the same way that the mass
balance involves the convection of mass per unit volume, p. We shall now
220 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.3 The Macroscopic Mechanical E

----,......::;..-+ Tzx 1 r + l:.r Ifwe substitute Eqs. 7.2-15 into Eq.


~--------~----~
q_ (pv ) + (pv~
at ~

(al
Lx
= pg.,

pv.X v.
y y +fJ.y

Taking the limits !:ix-O, !::.y-


motion in the x-direction,

a a a
at (pv~) + ax (pv.,v.,) + ay (pv.,vll
(b)
Fl1. 7.1-1. x-direction surface forces.
which we can obtain readily from Ec
analyze Eq. 7.2-14 term by term, starting with the first term on the left-hand
side.
i. In applying Eq. 7.2-10 to more co
af a
at -r pv., dV = at (pv., !l.x !l. y !l.z) (7.2-lSa)
careful in evaluating the surface inteJ
flux. The majority of errors that occ
momentum balance are caused by
- pv.,v.,., !l. y !:iz + pv.,v.,l:r:+A:r: !::.y !l.z Momentum llux
acrots x-aurfacca surface force acting on sorne portio!

f
~
pv.,v n dA = - pv~v1 1 1 !l.x !:iz + pv.,v 1v+A 11 11 !:ix !l.z
Momentum llux
acroso y-auaces (7.2-lSb)
*7.3 The Macroscopic Meche
- pv.,v,!. !l.x !l. y + pv.,v,!z+A !:ix !l. y Momentum llux
acr011 z-auaa:s
Energy Balance
J pg., dV = pg., !l.x !l. y !l.z (7.2-lSc)
Before we apply the macroscop
-r
Forc:cs on tbe ftows, it will be advantageous to de1
-T=I.,!:iy !l.z + T.,.,.,+A.,t::.y !:iz .x-auaces
balance so that we may compare t
T..,l 11 !l.x !l.z + T.~lv+Av !l.x !l.z
..,f ta):e dA =
Forces on tbe
(7.2-lSd) methods. We should keep in min<
- y-auac:cs
involved in this derivation. It is sil
F orc:cs on tbe viding us with an extremely useful e
-auac:cs
Chap. 7 Sec. 7.3 The Macroscopic Mechanical Energy Balance 221

lfwe substitute Eqs. 7.2-15 into Eq. 7.2-14 and divide by ~x ~y ~z. we obtain

~( v ) + (pv.,v.Jz+.1z - pv.,v.,l.,)
at P"' ~x

+ (pv.,v11 1 11H~; pv.,v~~l")


+ (pv.,v.lz+~; pv.,v.l.)
(7.2-16)
= pg., + (T"'"''"'H~: T.,.,,.,)
+ c~z~H4~; Ty.,lv)

+ (TzxlzH~; T..,l.)

Taking the Iimits ~x---+ O, ~y---+ O, ~z---+ O yields the stress equation of
motion in the x-direction,

aat (pv.,) + ax
a (pv.,v.,) +ay
a (pv.,v,J + az
a (pv.,v.)
(7.2-17)
_
-pg., + [ar.,.,
-- + - aTy.,
-+- ar-.,J
ax ay az
with the first term on the Ieft-hand which we can obtain readily from Eq. 7.2-3 by forming the dot product with
i. In applying Eq. 7.2-10 to more -complex control volumes, we must be very
(7.2-15a) careful in evaluating the surface integrals of the surface force and momentum
flux. The majority of errors that occur in the application of the macroscopic
Momontum llux momentum balance are caused by neglecting the momentum flux or the
x-surfaces
aaotl
surface force acting on sorne portion of the control surface.
Momontum llux
across y-audaces (7.2-15b)

Momcntum l!uJ< *7.3 The Macroscopic Mechanical


across z-surfaces
Energy Balance
(7.2-15c)
Before we apply the macroscopic momentum balance to sorne simple
Forc:es on tbc flows, it will be advantageous to derive the macroscopic mechanical energy
x-suaces
balance so that we may compare the results from these two approximate
Fon:et on tbc
y-suac:es (7.2-15d) methods. We should keep in mind that no new physical principies are
involved in this derivation. It is simply a mathematical development pro-
Forc:es on tbe
z-surfaces viding us with an extremely useful equation based on the linear momentum

,1
222 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.3 The Macroscopic Mechanical

principie. The development is somewhat longer than those presented in the where the restriction to incompre
previous two sections, for we must first derive the differential mechanical V p = O and V v = O. Analysis o
energy equation. The development will be restricted to incompressible flows. notation. We first note that
Starting with the stress equations of motion,
v(V
p (avat + vv) = pg +- V . T
V (7.3-1)
and then examine the term (ofox;)
we make use of the material derivative to obtain a
;- (v;T;;) =
UX;
Dv (7.3-2)
p-=pg+VT In Gibbs notation, the tensor (ov;/o
Dt
takes the form
Forming the scalar product of Eq. 7.3-2 with the velocity vector v gives V (v T) =
Dv The double dot (or scalar) product
pv - = v pg + v (V T) (7.3-3) index notation. For example,
Dt
There is no easy way of treating the left-hand side of Eq. 7.3-3, so we must A : B = A;;B,; = +
expand the scalar product and regroup the terms as follows:
+
Dv ) ( Dv., . Dv11 k Dv.) If both of the tensors A and B we
pv - = p(1v., + JV11 + kv. 1 - + J - + -
Dt Dt Dt Dt more careful in defining the double
stress tensor T is symmetric and we
= pv.,-
Dv.,
+ pv11 -Dv11 + pv.-
Dv. By Eq. 7.3-11, the second term o
Dt Dt Dt written
(7.3-4) V. (V. T) = ~
= !!__ (lpv;) + !!__ (tpv!) + !!__ (lpv;) which we may simplify by further a
Dt Dt Dt
side may be expressed in terms of ti
= !!__ (lpv2) Vv:T='
Dt
Switching to index notation,
where v2 = v2a: + v112 + v 2z The density has been treated as a constant in
accordance with the restriction to incompressible flows. In treating the
gravitational term, we note that any constant vector may be written as the ( ~V;) T;; = ~
UX; (

gradient of a scalar. For example, if we represent g as we note that


g = ig., + jgl/ + kg. (7.3-5) av) b;;P
(ox:
where g.,, g11 , and g. are constant, we may define a scalar cfo as
'
for incompressible flows. Incorp01
cP = -(xg., + ygll + zg.) (7.3-6) and defining the dissipation function
so that <1>=
g = -Vcfo (7.3-7) allows us to write
V. (V. T) =
We may now write the first term on the right-hand side of Eq. 7.3-3 as
It may not be apparent now (but it '
v pg = -v V(pcfo) =-V (pcfov) (7.3-8) rate at which mechanical energy is
ic Balances-lnertial Effects Chap. 7
Sec. 7.3 The Macroscopic Mechanical Energy Balance 223
1onger than those presented in the
where the restriction to incompressible flow has been imposed so that
derive the differentia1 mechanical
V p = O and V v = O. Ana1ysis of the Iast term is best done in index
be restricted to incompressib1e flows. notation. We first note that
motion,
cv. T) = v;(ar,i)
v. (7.3-9)
=pg+VT (7.3-1) ax,
and then examine the term (ajax,) (v;T to obtain1
;)

to obtain
- a (v.T..) = v. -
ar.. (av.)T. .
ax, " 'a x," + -' ax, " (7.3-10)
(7.3-2)
In Gibbs notation, the tensor (av;jax,) is represented by Vv, and Eq. 7.3-10
takes the form
with the velocity vector v gives
V (v T) = v (V T) + Vv : T (7.3-11)
The doub1e dot (or scalar) product between two tensors is clearly defined by
+V (V. T) (7.3-3)
index notation. For examp1e,

side of Eq. 7.3-3, so we must AuBu + A12B12 + A1aB13


the terms as follows: A : B = A;,B,, = + A21B21 + A22B22 + A23B23
+ AaiB31 + Aa2Ba2 + AaaBaa
lf both of the tensors A and B were skew symmetric, we wou1d have to be
more carefu1 in defining the doub1e-dot product. However, in our case the
Dv. stress tensor T is symmetric and we need not worry about this prob1em.
+ pv.-
Dt
By Eq. 7.3-11, the second term on the right-hand side of Eq. 7.3-3 may be
written
(7.3-4)
D v (V T) =V (v T)- Vv: T (7.3-12)
+ Dt (!pv;)
which we .m ay simp1ify by further ana1ysis. The 1ast term on the right-hand
side may be expressed in terms of the pressure and the viscous stress tensor.
Vv: T = Vv: ( -lp + 't) (7.3-13)
has been treated as a constant in Switching to index notation,
flows. In treating the av;) T,; = -a
(-a av; (-uuP + T,; .i )
(7.3-14)
vector may be written as the x, x,
we note that
(7.3-5) av)
b,ip = (av
(-ax, -.)p = o (7.3-15)
define a sca1ar cp as
-....._
ax;
for incompressib1e flows. lncorporating this simp1ification m Eq. 7.3-12
ygv + zg.) (7.3-6) and defining the dissipation function, <1>, as
<1> = Vv : 't (7.3-16)
(7.3-7) allows us to write
right-hand side of Eq. 7.3-3 as V (V . T) = V . (V T) - <1> (7.3-17)
1t may not be apparent now (but it will be in Chap. 10) that <1> represents the
(7.3-8)
rate at which mechanical energy is converted to thermal energy via viscous
Sec. 7.3 The Macroscopic Mechanical E
224 Macroscopic Balances-lnertial Effects Chap. 7

forces. We can understand this concept clearly only when the principie of Here, we have made use of
conservation of energy is analyzed in detail.t Substitution of Eqs. 7.3-4, D 2
- (~ pv) =-
o
7.3-8, and 7.3-17 into Eq. 7.3-3 yields the differential mechanical energy
equation for incompressible flows.
Dt ot
Time rate of change Rate of work done o
of k.inetic energy
per unit rume
by surfacc forces
per unit rume ot
since V v = O. Application of th
.Q_ (tpv 2
) = -V (pfv) +V (v T)- <1> (7.3-18) integral in Eq. 7.3-19, and of the e

Dt t t Ta
Rate of work done Rate of conversion
hy gravity per unit to thermal energy THE DISSIPATION FUN
volume per unit volume

This derivation has been somewhat more abstract than our previous efforts Rectangular coordinates x, y, z
and it may be helpful to discuss this result before going on to the derivation
of the macroscopic balance.
We started with the stress equations of motion, which representa balance
between the time rate of change of momentum per unit volume and the forces
<I> = 2"' [( av.)'
ax + (av)'
a; +
per unit volume. Forming the scalar product of the velocity and the time av.
rate of change of momentum gives rise to the time rate of change of kinetic +p. [( ox
energy, while the scalar product between the velocity anda force (force times
distance per unit time) naturally gives rise to the rate of work. The body Cylindrical Coordinates r, O, z
force, pg, gave rise to a term representing the rate of work done by gravity,

[( ov,)
while the surface force V T gave rise to two rate of work terms. The second 2
(~ ove ~
term represents the irreversible rate of surface work, which is converted to ll>= 2P. a:+ roO+,
thermal energy.
If the rate of surface work is negligible, Eq. 7.3-18 indicates a balance
+P. [ r-a (ve
between kinetic energy and potential energy. The student may recall experi- or -r
ments in beginning physics courses designed to demonstrate that kinetic and
potential energy were conserved provided "frictional" effects were small.
av, ov,
+P. (; + or
Expressions for <1> in rectangular, cylindrical, and spherical coordinates are
given in Table 7.3-1 for Newtonian fluids. There, we note that <1> is composed
Spherical Coordinates r, O,</>
entirely of squared terms and is therefore always positive.
Forming the integral of Eq. 7.3-18 over the volume "Ya(t) yields the
macroscopic balance,
ll> = 2P- [( av,)' (~ove ~)
or + r ao + r
:t (! pv ) dV +
2
f 2
V (t pv v) dV =
1'" 0 (t)
- f V (pfv) dV
r.(t)

+f f (7.3-19)
V (v T) dV- <1> dV sin O a (
r.(t) r.(t) +P. [ -,-a '
t This analysis is given in Secs. 10.1 and 10.2.
ic Balances-lnertial Effects Chap. 7 Sec. 7.3 The Macroscopic Mechanical Energy Balance 225

clearly only when the principie of Here, we have made use of


detaiJ.t Substitution of Eqs. 7.3-4,
the differential mechanical energy
(7.3-20)
Rate of work done
by surface forces
u rume
per ni t

since V v = O. Application of the general transport theorem to the first


+ V (v T) - <l> (7.3-18) integral in Eq. 7.3-19, and of the divergence theorem to the second, third,

done
unit
i
Rate of conversion
to thermal encrgy
Table 7.3-1
per unit volume THE DISSIPATION FUNCTION el> FOR NEWTONIAN FLUIDS

abstract than our previous efforts


Rectangular coordina tes x, y, z
t before going on to the derivation

of motion, which represent a balance cl>=2.u [(-


ox
2
ov,)+ -
oy
ov,)
ov.)+ -
oz
(
2
(
2
]

per unit volume and the forces


product of the velocity and the time
to the time rate of change of kinetic ov. ov.) ov, ov.) ov, ov.)']
2
(
2
(
+ .u [( ox + oy + oy + oz + ox + oz
the velocity and a force (force times
rise to the rate of work. The body Cylindrical Coordinates r, O, z
ng the rate of work done by gravity,
two rate of work terms. The second
surface work, which is converted to cl>=2.u
0Vr)
[ ( a;+
2
( 1 ov
~ar+;:-
0 ov,) ]
Vr) +a;
2
(
2

Eq. 7.3-18 indicates a balance


. The student may recall experi- +.u [ ra,.o (v-; )
8 1(oOovr)~J +.u [1~ (ov,)
+~
2

oO + ( a;-
ov 0 )]
2

to demonstrate that kinetic and


"frictional" effects were small. ovr ov,)"
rica!, and spherical coordinates are +.u ( oz +a;
There, we note that <l> is composed
always positive. Spherical Coordinates r, O, </>

over the volume "f/a(t) yields the


2
el> - 2.u -
ovr)" + (-
1 ov0
- + Vr)" + ( -1- --
- ovq, + -Vr + -
v0 cot
- 0) ]
- [ ( or r oO r r sin o orf> r r
dV=- JV (pr/Jv) dV
'f'".(t)

dV- I <l> dV
'f'".(t)
(7.3-19)
sin O o ( vq, ) 1 ov8]
+ .U [ -r- oO sin O + r sin Oaj;
226 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.3 The Macroscopic Mechanical E

and fourth integrals, gives

~t J (t pv ) dV +
2
J(t 2
pv ) (v - w) n dA (
r .w .ot.w
(7.3-21)
= J (v T) n dA - J pcfov n dA - J <1> dV
p._j_t..J~
.ot.w .ot.w r.w 1
1
Once again, there are sorne very definite reasons for transforming the volume L

integrals to area integrals. For example, if the gravitation term were left as
it first appeared in Eq. 7.3-3, the macroscopic balance would contain an
integral of the type

f v pg dV
r.(t)
Fig. 7.31. Movi1
Evaluation of this integral would require knowledge of v everywhere in the
system. However, expressing the gravity vector as the divergence of a scalar
and applying the divergence theorem yield the area integral in Eq. 7.3-21, In attacking the first term on the r
which we may evaluate with reasonable accuracy in terms of distance above remember that t(o) is the force per un
a reference plane and average velocities.t the system. Thus, the term
Following our development of the macroscopic mass and momentum
balances, we again wish to separate the total area daCt) into the area of (v T) n = v
entrances and exits, fixed solid surfaces, and moving solid surfaces. Thus,
represents the rate at which the surro
A.(t) Area of entrances and exits If we define Was the rate at which tn
da(t) = + A.(t) Area of solid moving surfaces (7.3-22) on the control volume, t then
+A. Area of so lid fixed surfaces

An illustrative control volume "Ya(t) is shown in Fig. 7.3-1, and we now .A,I
wish toexamine each term inEq. 7.3-21 to see what kind of simplifications or
special interpretations we can make in terms of the three different areas. The and the total area integral takes the f
first term (from left to right) remains unchanged. In the second term, we note
that
v = w =O on A.
J( v T) n dA '
:;,t.w
v - w = O on A.(t)
and we write Here, we have made use of the fact tt
of viscous dissipation to interna! ener
J(tpv )(v -
2
w) n dA = J(tpv )(v -
2
w) n dA (7.3-23)
..ot.w A.(t) E=11

t If we sought an exact solution, we would ha ve to know v everywhere, in which case


there would be no point in using the macroscopic balance and no advantage in using the t The dot over this term is used as a r
area integral. However, we seek approximate solutions to a great many problems, and in per second or sorne equivalent.
those cases the area integral has great preference.
Chap. 7 Sec. 7.3 The Macroscopic Mechanical Energy Balance 227

(
(7.3-21)
Jprf>v n dA - J <1> dV

reasons for transforming the volume


if the gravitation term were Ieft as
balance would contain an
Doshed line indicotes ~(f)

knowledge of v everywhere in the Fig. 7.3-1. Moving control volume,


vector as the divergence of a scalar
yield the area integral in Eq. 7.3-21,
In attacking the first term on the right-hand side of Eq. 7.3-21, we must
accuracy in terms of distance above
remember that t<a> is the force per unit area that the surroundings exert on
t the system. Thus, the term
macroscopic mass and momentum
total area da(t) into the area of (v T) n = v (T n) = v t(a) (7.3-24)
and moving solid surfaces. Thus,
represents the rate at which the surroundings do work on the control volume.
of entrances and exits
If we define W as the rate at which the solid moving surfaces are doing work
of so lid moving surfaces (7.3-22) on the control volume, t then
of solid fixed surfaces

is shown in Fig. 7.3-1, and we now


W= J V fa)
A,(t)
dA (7.3-25)
to see what kind of simplifications or
terms of the three different areas. The and the total area integral takes the form
In the second term, we note

on A, J (v T) n dA =
:W.(t)
J v t<a> dA
A,( t)
+W (7.3-26)

on A 8(t)
Here, we have made use of the fact that v = O on A . Defining the total rate
of viscous dissipation to interna! energy as
J(tpv )(v -
2
w) n dA (7.3-23)

ha veto know v everywhere, in which case


balance and no advantage in using the
solutions to a great many problems, and in t The dot over this term is used as a reminder that the units are foot pounds force
per second or sorne equivalent.
228 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.3 The Macroscopic Mechanical Ene

and incorporating Eqs. 7.3-23 and 7.3-26, we can write Eq. 7.3-21 and invoke the order of magnitude a
Time rate of change Net flux of kinetic Rate of work done on
of kinetic energy of
the control volume
energy leaving the
control volume
control volume at
entrances and exits
(v'l
--,
~t J(! pv 2
) dV + Jl pv\v- w) n dA J
= v t<"> dA v' T
f' a(t) A , (t) A , ( t) Eq. 7.3-28 reduces to the same form

-Jp4>v n dA
Rate of work done
on the control volume
by gravity
(7.3-27)

s1.(t)

Rate of work done


on the control volume Thus, the macroscopic mechanical ene
+W by solid moving
surfaces flows provided the terms are interpre
Rate of viscous Laminar flow
dissipation to
interna! energy
V

Turbulent flows V

(o)

In treating turbulent flows, we wish to obtain a form of Eq. 7.3-27


containing the appropriate time-averaged terms. This task is most easily
done by examining the time-averaged differential mechanical energy equation
and then reinterpreting the terms in Eq. 7.3-27. We may express each term
as the sum of time-averaged and fluctuating components as follows :
= v + v'
v
v 2
= v v = v2 + 2v v' + (v') 2 An important special case of Eq
T=T+T' control volume. Under these conditic

Substitution of these quantities into Eqs. 7.3-18 and 7.3-16, and taking the
time average, yieldt
J }pv2y n dA= Jv t<> dA- f P4
q_ [}pv 2
+ }p(v' )2 ] + v V[}pv 2 + } p(v') 2 ] If we compare this result with th
ot the same conditions (see Eq. 7.2-11
= -V. (p1>v) +V. (v. 'f<T>)- (~+~')+V. (v'. T') (7.3-28) product with the velocity vector the
boundaries has been eliminated. Ha
where f<T> = -Ip + i + i<t> dissipation term Ev, which must be ev<
T' = -Ip' + 't'' and mechanical energy balances, det
present two fairly distinct approachc
and we obtain <1>' by substituting v~, v~ , and v~ into the formulas for <1>. Ifwe
In general, we make use of the me
define the total viscous dissipation function as
surfaces are required, and of the mee
(i)<T> = {i) + {i)' (7.3-29) and velocities are required at entran
t The route to Eq. 7.3-28 follows the development in Chap. 6 and is straightforward, sol ved by both balances; they are in te
although lengthy. Here, again, we have expanded the material derivative as indicated by son between the two solutions and e:
Eq. 7.3-20. the validity of the assumptions made

Chap. 7 Sec. 7.3 The Macroscopic Mechanical Energy Balance 229
we can write Eq. 7.3-21 and invoke the order of magnitude arguments
Rate of work done on
control volume at (v')2 ~
entrances and exits 62

- w) n dA = f V t<> dA v' T' ~ v T<T>


.A,(t)
Eq. 7.3-28 reduces to the same forro as Eq. 7.3-18
Rate of work done (7.3-27)
dA on the control volume
by gravity (7.3-30)

Thus, the macroscopic mechanical energy balance may be applied to turbulent


ftows provided the terms are interpreted as follows:
Laminar flow Turbulent flow

V v
V v
- -(T)
t(n) t<n> = n T
to obtain a forro of Eq. 7.3-27
terms. This task is most easily
fferential mechanical energy equation
w f v t<> dA

A,(t)
. 7.3-27. We may express each term
components as follows: v f + [<i>
r.<t>
<i>'] dV

An important special case of Eq. 7.3-27 is for steady ftow and a fixed
control volume. Under these conditions, we obtain

s. 7.3-18 and 7.3-16, and taking the


f !pv~ n dA = J V t<n> dA - J pcf>v n dA - Ev
For a fixed control
volume and
(7.3-31)
A, steady flow

If we compare this result with the macroscopic momentum balance for


the same conditions (see Eq. 7.2-11), we see that by forming the scalar
_ (il> + <i>') + V (v' T') (7.3-28) product with the velocity vector the necessity of evaluating terms at solid
boundaries has been eliminated. However, it does give rise to the viscous
dissipation term .,, which must be evaluated experimentally. The momentum
and mechanical energy balances, derived from the same physica/ principie,
and v; into the formulas for <1>. lf we present two fairly distinct approaches to obtaining approximate solutions.
In general, we make use of the mornentum balance when forces on solid
surfaces are required, and of the mechanical energy balance when pressures
(7.3-29) and velocities are required at entrances and exits. Many problems can be
in Chap. 6 and is straightforward, solved by both balances; they are interesting cases to study, for the compari-
the material derivative as indicated by son between the two solutions and experimental data gives an indication of
the validity of the assumptions made.
230 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.-4 Bernoulli's Equation

In setting up macroscopic balance problems, the placement or choice of in a suitable fashion. The scalar
control volume deserves considerable attention. Often the same result will
be obtained for two different control volumes simply because the approxi- dr (pv Vv) =
ds
p(v., OJ
mations made to solve the problem hide the errors of the solution. Proper
placement of the control volume may not change the result, but it should alert
the engineer to the assumptions so often made tacitly. Two rules generally
= p(v., ~01
should be followed.
l. The control volume, whether it is fixed in space or moving, should be + p( v.,a;01
set up so that surface forces are easily calculated.
2. The surface of the control volume should be placed where the ftow is
one-dimensional, thus allowing the flux terms to be calculated easily.
Making use of Eq. 3.6-10 in Chap. 3

*7.4 Bernoulli's Equation v11 dz


v. dx
Before studying the application of the macroscopic balances, we will v., dy
derive Bernoulli's equation along a streamline. We do so by forming the
scalar product of the unit tangent vector to a streamline and the equations of which we may use to arrange Eq. 7.
motion, thus yielding the scalar component of the equations of motion in
the direction of the streamline. We are already familiar with the scalar -dr (pv Vv) = p [ v., (ov.,)
-
components in the coordinate directions (listed in Tables 5.1-1, 5.1-2, 5.1-3), ds ox
and the important aspect of the scalar component along a streamline is that
the inertial terms take a particularly simply form. + p[v~~(~~)~
The principal value of Bernoulli's equation is that it provides an extremely

+ p[v.(~:)d
simple relationship between velocity, pressure, and height above sorne refer-
ence plane for steady ftow when viscous effects may be neglected. Under
these conditions, the equations of motion reduce to
We have contrived to arrange the ind
pv Vv = -Vp + pg (7.4-1) as derivatives of the square of the ve

The unit tangent vector A may be expressed as

dr v
A=-=- (7.4-2) Making this change in each term an,
ds v
get
In forming the scalar product ofA with Eq. 7.4-1, we write
-dr (pv Vv) = p[1-
ds 2
dr dr
- (pv Vv) V(p + pcp) (7.4-3)
+ p[~
=--
ds ds
where the ftow is taken to be incompressible and the gravity vector has been
replaced by - Vcp. In treating the left-hand side of Eq. 7.4-3, we must
expand this term, form the scalar product, and regroup the individual terms
+PG
roscopic Balances-lnertial Effects Chap. 7 Sec. 7.-4 Bernoulli's Equation 231

problems, the phicement or choice of in a suitable fashion. The scalar product is given by
attention. Often the same result will
volumes simply because the approxi- dr. (pv. Vv) = p(v., avax
ds
+VIl
ay
+ v. av) (dr)
av
az ds
hide the errors of the solution. Proper
not change the result, but it should alert
= p(v., av., + V av., + Vz av.,) (dx)
made tacitly. Two rules generally ax ay 11
az ds
(7.4-4)
is fixed in space or moving, should be ( ., av
+ pv- ax v-
11+ av + vav!l)
ay az- -
11
(dy)
11

ds
easily calculated.
should be placed where the flow is
the flux terms to be calculated easily. ( ., av.
+p ax
av. -
v+- v
ay v
+ av.) (dz)
az- -ds11

Making use of Eq. 3.6-10 in Chap. 3, we obtain

v dz = v. dy
11
along a streamline (7.4-5a)
v. dx = v., dz along a streamline (7.4-5b)
the macroscopic balances, we will
streamline. We do so by forming the
v.,dy = v dx 11
along a streamline (7.4-5c)
to a streamline and the equations of which we may use to arrange Eq. 7.4-4 in the form
of the equations of motion in
are already familiar with the scalar
(listed in Tables 5.1-1, 5.1-2, 5.1-3),
dr.
ds
(pv. Vv) = p[v.,(av.,)dx + v.,(av.,)dy + v.,(av.,)dz]
ax ds ay ds az ds
component along a streamline is that
form. + p[v~~(av11 )dx + v (av )dy + v (av )dz] (7.4-6)
11 11

is that it provides an extremely ax ds ay ds az ds


11 11

fnr,,oonr<> and height above sorne refer-


+ p[v.(av.)dx + v.(av.)dy + v.(av.)dz]
effects may be neglected. Under ax ds ay ds az ds
reduce to
We have contrived to arrange the individual terms so that we may write them
(7.4-1) as derivatives of the square of the velocity component-i.e.,

v (av.,) =
., ax .2! ax~ (v2)., (7.4-7)

(7.4-2) Making this change in each term and summing the columns in Eq. 7.4-6, we
V
get
Eq. 7.4-1, we write
drds (pv VV) = p[1l aax ( 2 +V., V11
2
+ Vz) dx]
2
ds
(7.4-3)
+ p[.!2 ~
ay (v2., + v2 + v2) dsdy]11

(7.4-8)
and the gravity vector has been
side of Eq. 7.4-3, we must + p[.!2 az
~ (v2 + v2 + v2) dz]
., ds 11

and regroup the individual terms
232 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.4 Bernoulli's Equation

Noting that the differential operator dfds can be expanded in the form point as O and any other arbitrary po

~ = dx ~ + dy ~ + dz ?.._ lp(v~ - v~) + (p1 - J


(7.4-9)
ds ds OX ds oy ds oz Another variation of this develop
and that of many flows. We may obtain it by
(7.4-10) with the unit normal to the stream
we can write Eq. 7.4-8 normal to the streamline. The analys
obtaining the tangent scalar compon

(dr)
ds
.(pv. Vv) = P!2 dv2
ds
= ~ (t pv2)
ds
(7.4-11) will resort to an intuitive constructio
cates a streamline for steady flow u
provided the variations in the density are negligible. The right-hand side of
Eq. 7.4-3 is treated in an analogous manner to yield

(dsdr) .V(p + pcf>) = ~ds (p + pe/>) (7.4-12)

Note that the quantity A Vis often called the "directional derivative," for it
is a measure of the gradient in the A-direction by the formula

A. V = (dr)
ds
.V = dxds ~ox + dyds oy~ + dsdz oz?.._ = ~ds (7.4-13)

Substitution of Eqs. 7.4-11 and 7.4-12 into Eq. 7.4-3 yields

~ (!pv2 + p + pcf>) = O (7.4-14)


ds
which may be integrated to give
lpv 2 + p + pe/> = c. (7.4-15)
where c. is the constant of integration. The subscript s has been used to Fig. 7.4-1. Strear
indicate that the value of the constant may be a function of the streamline.
The assumptions for Eq. 7.4-15 should be kept in mind-i.e., steady, incom- curvature at two points. At each of
pressible flow and negligible viscous effects. The simplicity of this result the equations of motion normal to ti
makes it especially valuable for rapid qualitative evaluation of many flows. equation of motion in cylindrical e
In general, the coordinate system is taken so that the positive z-axis is coordinate system were Iocated as ir
oppositely directed to the gravity vector; thus, development is that for each point
e/> =gz (7.4-16) cylindrical coordinate system having
the streamline. Referring now to
and Bernoulli's equation takes the form r-component equation of motion

- -v:r + V, OVr)
lpv 2 +p + pgz = C, (7.4-17) OVr + Ve OVr
p
( Vr- -- -_
v
If the fluid is at rest, = O and the result derived in Sec. 2.2 for static fluids or r }(J oz
is obtained. In order to evaluate C,, we need to know the pressure and the
magnitude of the velocity vector at sorne point in space. If we designate this
+ pgr + .t [-aor (1-r a
- (rt
or
ic Balances-lnertial Effecu Chap. 7 Sec. 7.4 Bernoulli's Equation 233

can be expanded in the form point as O and any other arbitrary point as 1, Eq. 7.4-17 reduces to
!p(vi - v~) + (p 1 - p0 ) + pg(z1 - z0 ) = O (7.4-18)
(7.4-9)
Another variation of this development is useful in the qualitative analysis
of many flows. We may obtain it by forming the scalar product of Eq. 7.4-1
(7.4-10) with the unit normal to the streamline to determine pressure variations
normal to the streamline. The analysis is a good deal more complicated than
obtaining the tangent scalar component of the equations of motion, and we
1 dv 2 d will resort to an intuitive construction of the final result. Figure 7.4-1 indi-
- - = -(! pv)2 (7.4-11)
2 ds ds cates a streamline for steady flow using two tangent circles to describe the
are negligible. The right-hand side of
to yield

(7.4-12)

the "directional derivative," for it


~"-""v" by the formula

+ dy _q_ + dz _q_ = ~ (7.4-13)


ds oy ds oz ds
into Eq. 7.4-3 yields

(7.4-14)

pcf> =c. (7.4-15)


The subscript s has been used to
Fig. 7.4-1. Streamline for steady flow.
may be a function of the S!reamline.
be kept in mind-i.e., steady, incom-
curvature at two points. At each of those points, the scalar component of
effects. The simplicity of this result
the equations of motion normal to the streamline would be the r-component
ualitative evaluation of many flows.
equation of motion in cylindrical coordinates, provided the origin of the
taken so that the positive z-axis is
coordinate system were located as in Fig. 7.4-1. The key assumption in this
; thus,
development is that for each point on the streamline we can construct a
(7.4-16) cylindrical coordinate system having one of its coordinate curves tangent to
the streamline. Referring now to Table 5.4-2, we write the normal or
r-component equation of motion
(7.4-17)
derived in Sec. 2.2 for static fluids
p (v. OVr
or
+ ~ ovr -
roO
~ + v. OVr) = -
r oz
op
or
need to know the pressure and the
point in space. If we designa te this + pgr
, [oOr (1; oOr (rv,) ) + ~1 06o v,
T f-l
2

2 -
2 ov8
r2 ao
2
o v,]
+ OZ 2
7 4 19
( - )
2H Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.5 Sudden Expanslon in a Pipeline

For the special coordinate systems under consideration, vr = Vz =O at negligible. We can gain sorne insight i
r = r0 , where r0 is the radius of curvature of the streamline. Once again we dimensionless form of the equations of ti
neglect viscous effects, and Eq. 7.4-19 reduces to
a-+UVU=-
u
v: op
-p- = - - + pgr (7.4-20) o0
ro or
Most often, this result is expressed When the Reynolds number becomes
viscous terms would be negligible, but
v2
-p- = - op
- + pgn (7.4-21) example, Eq. 7.4-26 for steady flow in a
ro on
0=-V9"~
where n is the distance measured normally to the curve and gn represents the
scalar component of the gravity vector in the normal direction. Since v6 is
the only nonzero component of the velocity, it may be represented at every and the viscous terms are equal to the p.
point as the magnitude of the velocity, v. The point of this intuitive develop- Reynolds number might be. Thus, a
ment is to show that pressure variations across streamlines are hydrostatic necessarily mean negligible viscous effi
only if the streamlines are straight (r0 ---+ oo). If the streamlines are curved, viscous effects can be neglected when
the pressure variation will be different from hydrostatic according to the
term - pv2fr 0 Because of this the surfaces of control volumes will always be UVU~
placed where the streamlines are straight, so that the surface forces may be
calculated easily. If the Reynolds number is large and
UVU=
Turbulence
the inequality will be satisfied. This latt
In applying Bernoulli's equation to steady turbulent flows, we again change of velocity in the direction of flov
neglect viscous effects and write Eq. 7.4-17 ast of velocity perpendicular to the flow.
regarding the importance of viscous e
lp[(v + v') (v + v')] + (p + p') + pg(z + z') =
o e, (7.4-22)
compare the results with experiment, f
Forming the scalar product in the first term and taking the time average, we guide us in making these approximatio
find
(7.4-23)

Once, again, we invoke the order of magnitude argument that PART ll-APE
(V')2 ~ jj2 (7.4-24)
and the time-averaged form of Bernoulli's equation becomes 7.5 Sudden Expansion in a Pipe
lpv 2
+ p + pgz = c. (7.4-25)
In this example, we wish to calcul
Viscolis effects sudden expansion in a pipeline, such as
line indicates the surface of the control v
Before applying Bernoulli's equation, we need to consider under what macroscopic mass, momentum, and me
conditions we might expect the viscous terms in the equations of motion to be a and b indica te the segment of the cen
Bernoulli's equation. The flow is stead
t Here we are making use of the fact that the time average of ov'fot is zero.
Balances-lnertial Effects Chap. 7 Sec. 7.5 Sudden Expansion in a Pipeline 235

under consideration, vr = vz =O at negligible. We can gain sorne insight into this problem by examining the
of the streamline. Once again we dimensionless form of the equations of motion derived in Sec. 5.5.
reduces to
op au +u. vu = -V&J + - 1-V 2U (7.4-26)
- or + pgr (7.4-20) o0 Nne
When the Reynolds number becomes very large, it might seem that the
viscous terms would be negligible, but this is not necessarily so. As an
(7.4-21) example, Eq. 7.4-26 for steady fiow in a pipe reduces to

to the curve and g n represents the o= -V&J + - 1-V 2U (7.4-27)


in the normal direction. Since v8 is Nne
, it may be represented at every and the viscous terms are equal to the pressure terms regardless of what the
v. The point of this intuitive develop- Reynolds number might be. Thus, a high Reynolds number does not
across streamlines are hydrostatic necessarily mean negligible viscous effects. In general, we will find that
--+ ex:>). If the streamlines are curved, viscous effects can be neglected when
from hydrostatic according to the
of control volumes will always be (7.4-28)
so that the surface forces may be
If the Reynolds number is large and

U VU = 0(V2U) (7.4-29)

the inequality will be satisfied. This latter condition requires that the rate of
to steady turbulent flows, we again change of velocity in the direction of fiow is comparable to the rate of change
.4-17 ast of velocity perpendicular to the fiow. The best way to gain sorne insight
+ p') + pg(z + z') = c. (7.4-22) regarding the importance of viscous effects is to solve sorne problems and
compare the results with experiment, for there are no hard and fast rules to
term and taking the time average, we guide us in making these approximations.

(7.4-23)
tude argument that PART 11-APPLICATIONS
(7.4-24)
's equation becomes
7.5 Sudden Expansion in a Pipeline
pgz =e, (7.4-25)
In this example, we wish to calculate the pressure change caused by a
sudden expansion in a pipeline, such as that shown in Fig. 7.5-1. The dashed
line indicates the surface of the control volume to which we wish to apply the
we need to consider under what macroscopic mass, momentum, and mechanical energy balances. The points
terms in the equations of motion to be a and b indica te the segment of the center streamline to which we will apply
the time average of ov'fot is zero. Bernoulli's equation. The flow is steady, turbulent and incompressible.
236 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.5 Sudden Expansion in a Pipeline

"O" "1" "2" vector is given by


1
1

In all these examples, we will dispense


should be used with turbulent flows, a
adjustments necessary to account pro
tution of Eq. 7.5-4 into Eq. 7.5-3, and
the unit vector k, yield

J pv,v odA=
A,

where we have made use of the fact th1


placed far enough up.stream and d
dimensional at these pomts; however, t
be independent of the placement of Oa
Fig. 7.5-1. Sudden expansion in a pipeline.
correct location is in sorne respects a
the disturbance propagate upstream
Mass balance one go before the flow is on~.-dime~si
"What is the flow topology? Askmg
For a fixed control volume and steady, incompressible flow, we start with assumptions made in solving the protl
Eq. 7.1-11 2 is very large,t the assumption of neg
J
A,
pv n dA= O (7.5-1)
The area integral on the left-hand s
at the entrance and exit of the control
right-hand side must be evaluated
which quickly reduces to
8 k = O along the walls of the pipe, a

(7.5-2) area averages yields

or 2
- p(v,) 0 4D~ + p(v,)s-
11 2 ?:

1TD2 1T
= (P)o7 + (p)4
Momentum balance

There is a very important point


We start with Eq. 7.2-11 ( 11f4)(D:- ~), for experience has s

J
A,
(pv)v n dA= J
r
pg dV + J
d
t<n> dA (7.5-3)
this term. The reason appears to be a
the analysis of macroscopic balance
entrances and exits and completely f4
and assume that viscous and turbulent stresses are small,t so that the stress ever the symbol .91 appears on an int
each and every portion of the control
t The assumption of negligible viscous effects is to be made throughout this chapter.
Turbulent stresses are identically zero at solid surfaces and are generally considered to be entrances or exits.
small compared to the pressure at entrances and exits. t The distance between O and 1 is negligi
Sec. 7.5 Sudden Expansion in a Pipeline 237

"2"
vector is given by
1
1 t(n) = -np (7.5-4)

In all these examples, we will dispense with the time-average symbol, which
should be used with turbulent flows, and require the student to remember the
adjustments necessary to account properly for the time averaging. Substi-
tution of Eq. 7.5-4 into Eq. 7.5-3, and formation of the scalar product with
the unit vector k, yield

J
A,
pv.v n dA = - J
.!#
pk n dA (7.5-5)

where we have made use of the fact that k g = O. The planes at O and 2 are
placed far enough upstream and downstream so that the flow is one-
dimensional at these points; however, the answer obtained in this example will
be independent of the placement of Oand 2. The purpose in placing them in a
correct location is in sorne respects a device for forcing the questions: "Does
the disturbance propagate upstream at all ?" "How far downstream must
one go before the flow is one-dimensional ?" or, in the language of Sec. 1.3,
"What is the ftow topology ?" Asking these questions helps to emphasize the
y, incompressible flow, we start with assumptions made in solving the problem, for if the distance between 1 and
2 is very large,t the assumption of negligible viscous forces is unsatisfactory.
dA =0 The area integral on the left-hand side of Eq. 7.5-5 is to be evaluated only
(7.5-1)
at the entrance and exit of the control volume, while the area integral on the
right-hand side must be evaluated over the entire surface. Noting that
n. k = O along the walls of the pipe, and expressing the integrals in terms of
(7.5-2) area averages yields

(7.5-6)

There is a very important point to be made regarding the term (p)1


('1T/4)(D:- Dl:), for experience has shown that students will often neglect

dV + f
.o/
t<> dA (7.5-3)
this term. The reason appears to be associated with a general error made in
the analysis of macroscopic balance problems-i.e., focusing attention on
entrances and exits and completely forgetting about other surfaces. When-
stresses are small,t so that the stress ever the symbol .91 appears on an integral, the student is urged to examine
each and every portion of the control surface, especially those which are not
is to be made throughout this chapter.
surfaces and are generally considered to be entrances or exits.
exits. t The distance between O and 1 is negligible except for ftows where Nxe < 1.
Sec. 7.5 Sudden Expansion in a Pipeline
238 Macroscopic Balances-lnertial Effects Chap. 7

If we wish to develop an expression for the pressure differencet The assumptions for this result follow.

Ap = (p )o - (p )2 (7.5-7) l. Viscous effects were neglected-


considered negligible.
we must eliminate (p )1 from Eq. 7.5-6. If the distance between O and 1 is
2. The average pressure on the annu
negligible and the pressure variation over the annular region is hydrostatic,
the average pressure at point O.
it is reasonable to make the approximation
3. The velocity profiles were assumec
(p)l = (p )o (7.5-8)
Next, we will attack this problem using t
and Eq. 7.5-6 can be rearranged in the forro will allow us to obtain a solution for AP

p(v~) 2 p(v~)o(~:)
2
different from those listed above.
Ap = - (7.5-9)

F or turbulent flows, the velocity profile is nearly flat; thus, the average of the Mechanical energy balance
square may be approximated by the square of the average. We shall investi-
gate this approximation in more detaillater; for now, we simply write,
We start this analysis with the stead
(7.5-10) the mechanical energy balance
and make use of the macroscopic mass balance to arrange Eq. 7.5-9 in the
forro f!
A,
2
pv v n dA = f
A,
v t<nl

(7.5-11)
In this case, neglecting viscous effects re
We note that
E=
(~~ - 1 ~o
2
V

Since the flow is one-dimensional at the


and the pressure therefore rises from O to 2. It often seems incongruous that
the pressure would rise in the direction offlow. An intuitive explanation may
be helpful. As a particle of fluid moves from O to 2, its velocity, and hence its
V (nl = (kv.) t<n> =
momentum, must decrease. F or the momentum of a fluid partide to decrease,
a force must be applied in the direction opposite to the flow. In the sudden But ov.foz is zero for one-dimensional f
expansion this force is the pressure rise.
In comparing this result with experimental data, it is convenient to define
V (n) =
a dimensionless pressure difference as indicating that the viscous effects at the e
AP=~ (7.5-12)
The potential energy flux term may be \
!p(v.)~
where the factor of! in the denominator is simply a matter of tradition. The
dimensionless pressure difference given by the momentum equation now
f
A,
p<fov n dA= -p(cpv.)

takes the forro


Since the velocity proflles are taken to
(7.5-13)

t The symbol tlp will always be used to denote upstream pressure minus downstream
f
A,
pcpv n dA = - p(v.)o 7!.
pressure.
Chap. 7 Sec. 7.5 Sudden Expansion in a Pipeline 239

for the pressure differencet The assumptions for this result follow.

- (p)2 (7.5-7) l. Viscous effects were neglected-i.e., viscous surface forces were
If the distance between O and 1 is considered negligible.
2. The average pressure on the annular surface was taken to be equal to
the average pressure at point O. '
3. The velocity profiles were assumed to be flat atO and 2.
(7.5-8)
Next, we will attack this problem using the mechanical energy balance, which
form
will allow us to obtain a solution for ~Pwhile making simplifying assumptions

p(v!>o(~:)
2
different from those listed above.
- (7.5-9)

is nearly flat; thus, the average of the Mechanical energy balance


square of the average. We shall investi-
later; for now, we simply write,
We start this analysis with the steady flow, fixed control volume form of
(7.5-10) the mechanical energy balance
balance to arrange Eq. 7.5-9 in the
(7.5-14)

(7.5-11)
In this case, neglecting viscous effects requires that we assume

-1:::;:0 (7.5-15)

Since the flow is one-dimensional at the entrance and exit, we may write
O to 2. lt often seems incongruous that
offlow. An intuitive explanation may
from O to 2, its velocity, and hence its V t(n) = (kv.) t<n> = v.(=f p 2.t ~;) (7.5-16)
of a fluid particle to decrease,
opposite to the flow. In the sudden But av.faz is zero for one-dimensional flows and Eq. 7.5-16 becomes

data, it is convenient to define (7.5-17)

indicating that the viscous effects at the entrance and exit are identically zero.
The potential energy flux term may be written as
(7.5-12)
n~0
is simply a matter of tradition. The
f pcfov n dA = - p(cfov.) 0 - + p(cfov.)2 -n~2 (7.5-18)
by the momentum equation now
4 4
A,

Since the velocity proflles are taken to be flat , and 1/> = gy , this term is zero ,
(7.5-13)
(7.5-19)
denote upstream pressure minus downstream
Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.5 Sudden Expansion in a Pipeline

Incorporating this result into Eq. 7.5-14 along with Eqs. 7.5-15 and 7.5-17, points a and b to obtain
and expressing the integrals in terms of averages and areas, we get
tpv! + Pb + pcf>b =
- p v3)o1r-D~
t (z - ( 3) 1r D~
+ t pv2--
4 z 4 Since cf>b = cf>a, the pressure difference
(7.5-20)
= - (pv.)2 4
1rDi
+ (pv. )o 41rD~ Pb- Pa =
To proceed with the ana1ysis and make
We again assume that the velocity profiles are flat in order to write we must assume that the velocity pro
(v~) = (v.) 3
and (pv.) = (p)(v.) (7.5-21) write
and apply the mass balance to Eq. 7.5-20 to obtain the dimensionless pressure Va =
difference

(7.5-22) Because the pressure variation is line2

The assumptions for this result follow. Pa =


l. Viscous effects were neglected-i.e., viscous dissipation was considered pb=
negligible. Substitution of Eqs. 7.5-27 and 7.5-28
2. The velocity profiles were assumed to be flat at O and 2.
(p)2 - (P)o = !~
Aside from the tlat velocity profiles, the assumptions made to obtain a
solution with the mechanical energy balance were very different from those Use of the macroscopic mass balance
made in applying the momentum balance. That the results differ simply pressure difference give
indica tes sorne of the assumptions must be in error. In using the mechanical
energy balance, no assumption regarding the pressure acting on the annular ~p = [ (3
region was necessary. The assumption of negligible viscous effects leads to
two different assumptions when applied to the momentum and mechanical which is the same result obtained '
energy balances. In the former case, it leads to the neglect of viscous surface Because the two equations were derive
forces assumptions made in solving this pro!
over the surface of the resu1t to be similar if not identical
the control volume (7.5-23)
Very often the final result can b
while in the latter case it leads to the neglect of viscous dissipation Bernoulli's equation instead of the 1
the mechanical energy balance is deri'
over the volume of
Vv: T ~O (7.5-24) macroscopic ( or integral) problems, ar
the control volume
Later, we shall encounter sorne prc
Before comparing these two solutions with experimental data, let us apply equation; thus, it is necessary to beco
Bernoulli's equation to this problem. Although Bernoulli's equation is not The results obtained from the me
particularly suited for analyzing problems of this type, it will help us gain
energy balance are compared with exf
sorne practice in its application.
From this comparison, we learn th
tum balance and experiment is reas01
Bernoulli's equation along a streamline the mechanical energy balance becomc
In this particular instance, the viscous
We assume the tlow to be steady and neglect viscous effects. Under these in the momentum balance, whereas
conditions, we may apply Bernoulli's equation along the streamline between mechanical energy balance may not 1:
Chap. 7 Sec. 7.5 Sudden Expansion in a Pipeline 241

.5-14 along with Eqs. 7.5-15 and 7.5-17, points a and b to obtain
of averages and areas, we get
ipv: + Pb + PtPb = !pv! + Pa + PtPa (7.5-25)
a) D~
1 (
2P v. 24 1r
Since tPb = tPa the pressure difference between points a and b is
(7.5-20)
Pb- Pa = ip(v!- v~) (7.5-26)
To proceed with the analysis and make use of the macroscopic mass balance,
are flat in order to write
we must assume that the velocity profiles are flat atO and 2. We may then
(7.5-21) write
to obtain the dimensionless pressure Va= (v.)o (7.5-27a)
vb = (v.)2 (7.5-27b)
(7.5-22) Because the pressure variation is linear across the pipe, we may also write
Pa = (p )o (7.5-28a)
., viscous dissipation was considered Pb = (p )2 (7.5-28b)

Substitution of Eqs. 7.5-27 and 7.5-28 into Eq. 7.5-26 gives


to be flat at O and 2.
the assumptions made to obtain a (p)2 - (P)o = !p[(v.)~ - (v.)~] (7.5-29)
balance were very different from those Use of the macroscopic mass balance and the definition of the dimensionless
balance. That the results differ simply pressure difference give
be in error. In using the mechanical
the pressure acting on the annular (7.5-30)
of negligible viscous effects leads to
ied to the momentum and mechanical which is the same result obtained with the mechanical energy balance.
it leads to the neglect of viscous surface Because the two equations were derived in nearly the same manner and the
assumptions made in solving this problem were the same, we should expect
over the surface of the result to be similar if not identical.
the control volume (7.5-23)
Very often the final result can be obtained with less effort by using
neglect ofviscous dissipation Bernoulli's equation instead of the mechanical energy balance; however,
over the vo/ume of the mechanical energy balance is derived for the express purpose of solving
the control volume (7.5-24) macroscopic (or integral) problems, and it is best to use it whenever possible.
Later, we shall encounter sorne problems best treated with Bernoulli's
with experimental data, Jet us apply
equation; thus, it is necessary to beco me skillful in applying both methods .
. Although Bernoulli's equation is not
The results obtained from the momentum balance and the mechanical
of this type, it will help us gain
energy balance are compared with experiments in Table 7.5-1.
From this comparison, we learn that the agreement between the momen-
tum balance and experiment is reasonably good, while the agreement with
the mechanical energy balance becomes progressively worse as (D 0 /D 2 )-+ O.
In this particular instance, the viscous surface forces are justifiably neglected
and neglect viscous effects. Under these in the momentum balance, whereas the viscous dissipation term in the
equation along the streamline between mechanical energy balance may not be neglected. lf the dissipation term is
242 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.5 Sudden Expansion in a Pipeline

Table 7.5-1 Table


COMPARISON OF CALCULATED AND MEASURED PRESSURE DEPENDENCE OF TH1
RISE AT A SUDDEN EXPANSION IN A PIPELINE" THE REYNOLDS

Ratio of Experimental Momentum Mechanicl Energy


Diameters Value Balance, Eq. 7.5-13 Balance, Eq. 7.5-22 n = _______..,6
D0 /D 2 l!..Pexp l!..Pmom l!..Pmech eng
NRe = 4 X 10"
0.843 -0.33 -0.41 -0.49
0.695 -0.42 -0.50 -0.77
The, average velocity in a pipe is g
0.590 -0.45 -0.45 -0.88
0.394
0.328
-0.24
-0.18
-0.26
-0.19

a Data obtained from W. H . Archer. Trans. ASCE. 1913. 76: 999.


-0.98
-0.99

(v.) =
f~ r:. max( 1 -

:.o_..:o___2~--,---
~rn
included in the analysis, Eq. 7.5-22 becomes
4
J f r dr dO
fl.P = [(Do) _ 1] + _8..,.--::.Ev_ (7.5-31) o o
D2 1r D~p(v.)~
Inasmuch as the dissipation term is always positive, Eq. 7.5-31 indicates that Similarly, we can determine the averag
the pressure difference determined by the mechanical energy balance would
be smaller (and therefore in better agreement with the experimental data) if
the viscous dissipation term were included.
We should make an important point regarding these results. Namely,
the momentum and mechanical energy balance may give results which are (v~) = v~.max-
not very satisfactory. lf they both give essentially the same result, there is a
good chance that the result is satisfactory. However, if they give appreciably Using these results, we can test
different results, the engineer must beware. Another point to be made is
7.5-32 for various values of n. The re
that it is not at al! obvious, a priori, that it was reasonable to neglect viscous
that even though the velocity profil
surface forces and unreasonable to neglect viscous dissipation. It takes more
(v.)fvz,max = 1.0, the approximations J
experience than a student can possibly gain in one or two courses in fluid
The values for the parabolic profile w
mechanics to make sound judgments regarding such assumptions.

Tabl
Velocity profiles
DEVIATIONS FROM THE

For future reference, we wish to investigate the validity of the approxi- Turbulent
mations 6 7 8
n
(v~) = (v.) 2 (7.5-32a)
(v~) = (v.) 3
(7.5-32b) (v.)
-- 0.79 0.82 0.84
v%,max
Experimental measurements of the time-averaged velocity for turbulent flow
in a pipe indicate that the velocity profile is given approximately by <v!> 1.03 1.02 1.02
r )1/n <v.>"
Vz = Vz,max ( 1 - ;:: (7.5-33)
(v:)
1.08 1.06 1.05
where n is given as a function of the Reynolds number in Table 7.5-2. (v.)'
Sec. 7.5 Sudden Expansion in a Pipeline 243

Table 7.5-2
DEPENOENCE OF THE COEFFICIENT, n, ON
THE REYNOLOS NUMBER, NRe
MechanicaJ Energy
Balance, Eq. 7.5-22
llPmech eng
NRe =
n-~
4 X
6
1()3
1 1-10
1 X
7
10 3 X 106
-0.41 -0.49
-0.50 - 0.77
-0.45 -0.88
The?average velocity in a pipe is given by
-0.26 -0.98
211 ro
-0.19

1913. 76: 999.


-0.99
JJ
o o
v.max(1- :Y'nrdrdO
0
2n 2
(v ) - -v ------
z,max (1 + n)(1 + 2n)
f
- 211 r -

+ 8Ev
Jr dr d()
1] (7.5-31) o o (7.5-34a)
7T D~p(v.)~

ys positive, Eq. 7.5-31 indicates that Similarly, we can determine the average ofthe square and cube ofthe velocity.
the mechanical energy balance would
iYrPPrYIPnt with the experimental data) if
n2
(v2) - v2 (7.5-34b)
z - z,max (1 + n)(l + n)
regarding these results. Namely,
balance may give results which are 3 3 2n2
(v ) - v (7.5-34c)
essentially the same result, there is a z - z,max (3 + n)(3 + 2n)
However, if they give appreciably
Using these results, we can test the approximations indicated by Eqs.
Another point to be made is
7.5-32 for various values of n. The results appear in Table 7.5-3 and indicate
it was reasonable to neglect viscous
that even though the velocity profile is certainly not flat, for which case
viscous dissipation. It takes more
(v.)Jv.max = 1.0, the approximations given by Eqs. 7.5-32 are very reasonable.
gain in one or two courses in fluid
The values for the parabolic profile which occurs when the flow is laminar are
regarding such assumptions.

Table 7.5-3
DEVJATIONS FROM THE FLAT YELOCITY PROFILE

the validity of the approxi- Turbulent Laminar

n 6 7 8 9 10 Parabolic Profile
(7.5-32a)
(7.5-32b) (v.)
-- 0.79 0.82 0.84 0.85 0.87 0.5
V.z , max

<v!> 1.03 1.02 1.02 1.01 1.01 1.33


(v.>2
(7.5-33)
<v!> 1.08 1.06 1.05 1.04 1.03 2.0
number in Table 7.5-2. (v.)"
244 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.6 Sudden Contraction in a Pipeline

also given for comparson, and we see that thc terms in the macroscopc In this case, our intuition does not eno
balances must be treated wth caution f the flow is laminar. does it seem reasonable to equate (p) 1 2
sion for the pressure difference betw
must be eliminated from Eq. 7.6-l. F
7.6 Sudden Contraction in a Pipeline
(p ) 0 and assume the velocity profiles

In ths example, we again wish to apply the momentum and mechanical


energy balances to a particular system and then compare the results with (P)o - (p)2 = p(v
experimental data. We will not apply Bernoulli's equation inasmuch as it
In this case, we define the dimensionl
"O" "1'' "2"
1 1
1 1
1 1
1
1
1
the denominator of which differs fro
1 The macroscopic mass balance may E
1 and 7.6-3 to yield
1

Because (D 2 /D 0 ) 2 is always less than


in the direction of flow. Since the m1
to 2, a force in that direction must
supplied by the pressure drop.

Mechanical energy equation

Fl. 7.6-1. Sudden contraction in a pipeline. Starting with Eq. 7.3-31, we assurr
negligible and note from the previous
obviously gives the same result as the mechancal energy balance. The is zero. We obtain
sudden contraction and the control volume to be used are illustrated in
Fig. 7.6-1.

Momentum balance Application of this equation to the s


and the result is
Starting with Eq. 7.2-11, we neglect the viscous stresses and apply it to
this system to obtain

2) 7TD~ 2 7TD~
- p(v, o4 + p(v,)2 4 Once again, the results from the two I1
(7.6-1) The comparison with experiments is
'7T D~
= ( P) o 4 - (
P)t
'7T ( 2 2
Do - D2) - (p)2 4Di
'7T balance predicts a pressure drop ti
energy balance predicts a pressure dr'
Balances-lnertial Effects Chap. 7 Sec. 7.6 Sudden Contraction in a Pipeline 245

that the terms in the macroscopic In this case, our intuition does not encourage us to set (p) 1 equal to (p) 0 , nor
if the flow is laminar. does it seem reasonable to equate (p )1 and (p )2 However, to derive an expres-
sion for the pressure difference between points O and 2, the pressure at 1
Pipeline must be eliminated from Eq. 7.6-1. For this reason we will set (p) 1 equal to
(p) 0 and assume the velocity profiles are flat atO and 2. We obtain
apply the momentum and mechanical 2
and then compare the results with (P)o - (p)2 = p(v.)2
2- p(v.)o2(Do)
D (7.6-2}
2
Bernoulli's equation inasmuch as it
In this case, we define the dimensionless pressure difference as
"2"
1
1 fl.p = (P)o - (p)2 (7.6-3)
1
1 !p(v.)~
1
1 the denominator of which differs from that used for the sudden expansion.
1 The macroscopic mass balance may be used in conjunction with Eqs. 7.6-2
1 and 7.6-3 to yield
1

(7.6-4)

Because (D 2/ D0} 2 is always less than unity, Eq. 7.6-4 predicts a pressure drop
in the direction of flow. Since the momentum of the fluid increases from O
to 2, a force in that direction must be applied to the fluid and this force is
supplied by the pressure drop.

Mechanical energy equation

in a pipeline. Starting with Eq. 7.3-31, we assume that the rate of viscous dissipation is
negligible and note from the previous example that the potential energy term
mechanical energy balance. The is zero. We obtain
to be used are illustrated in
f<tpv~vndA = - JpvndA (7.6-5)
A.

Application of this equation to the sudden contraction is straightforward,


and the result is
the viscous stresses and apply it to
(7.6-6)

Once again, the results from the two macroscopic balances are very different.
(7.6-1) The comparison with experiments is given in Table 7.6-1. The momentum
balance predicts a pressure drop that is too large and the mechanical
energy balance predicts a pressure drop that is too small.
246 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.7 The Nozzle and the Borda

Table 7.6-1 is always negative, oP must be


COMPARISON OF CALCULATED AND MEASURED PRESSURE DIFFERENCES experimental data. Thus, the pressu
AT A SUDDEN CONTRACTION IN A PIPELINE& O. We may make an attempt to
Bernoulli's equation along a stream
Experimental Momentum Mechanical Energy
Balance, Eq. 7.6-6
mechanical energy balance, we
Value Balance, Eq. 7.6-4
D./Do t:lPexp t:lPmom t:lPmech eng
negligible; however, it is probable t
the region between 1 and 2. Neglecti
0.91 0.35 0.34 0.31 effects, we find that Bernoulli's
0.71 0.98 0.99 0.75
0.50 1.30 1.50 0.94
0.33 1.41 1.78 0.99 If we assume
0.20 1.45 1.92 1.00 op =
0.10 1.46 1.98 1.00
0.0 1.47 2.00 1.00 (v.)~ =
we may write Eq. 7.6-10 as
a Data are taken from H. W. King and E. F. Brater, Handbook of Hydraullcs (New York:
McGraw-Hii!BookCompany, Inc.,l963), pp. 6-21. Experimental valuesare for Reynoldsnumbers
on the order of 10.
oP=
Since we do not know the flow field at
The error in the momentum balance must arise either from the assumption
however, it seems logical that vb >
that (p )1 is equal to (p )0 , or from the neglect of viscous surface forces. Let
our previous deduction that oP must
us return to Eq. 7.6-1 and modify it to include a viscous surface force, F,,
In comparing the mechanical
which must act in the direction opposite to the flow and also express the
we can conclude easily that the
average pressure at point 1 as
of disagreement between the ... a, ... u,au..
(p ) = (p )o + op (7.6-7) is included iri the analysis, Eq. 7.6-6
Equation 7.6-1 now becomes

+ p(v.)2 4 =
2 7T v: (P)o 4D~
7T

(7.6-8) Because E, is positive, inclusion of vi

- [(p)0
2v: F,
+ op] (Do- D2)- (p)24-
7T 2 7T
necessarily improve the result.
Once again, the momentum and
noulli's equation provide only approxi
Assuming the velocity profiles to be flat, and making use of the macroscopic expansion in a pipeline, the mome
mass balance, we can write the dimensionless pressure drop, agreement with experiment while the
poor results. For the sudden contract
(7.6-9) error as (D 2/ D 0 ) - O. If nothing else
the student to view approximate se
where
section, we shall continue this comp
oP = __jp_ solutions and experimental data for
ip(v.>:
We see from this result that the inclusion of viscous forces would make L'l.P
7.7 The Nozzle and the Bord
even larger, thus increasing the disagreement between calculation and
experiment. lnasmuch as the term Figure 7.7-1 illustrates a large tar
the bottom. If the nozzle is properly
parallel as the fluid issues into the su1
Chap. 7 Sec. 7.7 The Nozzle and the Borda Mouthpiece 247

is always negative, P must be negative for Eq. 7.6-9 to agree with the
experimental data. Thus, the pressure at l must be less than the pressure at
O. We may make an attempt to calculate the pressure at 1 by applying
Bernoulli's equation along a streamline from a to b. From the results of the
Mechanical Energy
Balance, Eq. 7.6-6 mechanical energy balance, we already know that viscous dissipation is not
t:.Pmech eng negligible; however, it is probable that most of the dissipation takes place in
the region between 1 and 2. Neglecting viscous dissipation and gravitational
0.34 0.31 effects, we find that Bernoulli's equation gives us
0.99 0.75
1.50 0.94 Pb- Pa = !p(v~- v~) (7.6-10)
1.78 0.99 If we assume
1.92 1.00 p = Pb- Pa
1.98 1.00
2.00 1.00 ( v.)~ = v!
we may write Eq. 7.6-10 as
. Brater, Handbook of Hydrau/lcs (New York:
l. Experimental values are for Reynolds numbers
(7.6-11)

must arise either from the assumption Since we do not know the flow field at l, we can carry this analysis no further;
neglect of viscous surface forces. Let however, it seems logical that vb > (v.)0 and Bernoulli's equation confirms
include a viscous surface force, Fv, our previous deduction that (Jp must be negative.
to the flow and also express the In comparing the mechanical energy balance with the experimental data,
we can conclude easily that the neglect of viscous dissipation is the cause
of disagreement between the calculated and experimental values. lf this term
(7.6-7) is included iri. the analysis, Eq. 7.6-6 takes the form

f:.p = [t - (D2)4] +
D0
8v
p( v.)~1T D~
{7.6-12)

(7.6-8) Because v is positive, inclusion of viscous dissipation in the analysis would


necessarily improve the result.
{D~- DD- (p)2 1T:~- Fv Once again, the momentum and mechanical energy balances and Ber-
nouJli's equation provide only approximate answers. In the case ofthe sudden
and making use of the macroscopic expansion in a pipeline, the momentum balance gave reasonably good
pressure drop, agreement with experiment while the mechanical energy balance gave rather
poor results. For the sudden contraction, both balances showed considerable
(7.6-9) error as ( D 2/ D0 ) --+ O. lf nothing el se, these two examples should encourage
the student to view approximate solutions with skepticism. In the next
section, we shall continue this comparison between the three approximate
_jp_ solutions and experimental data for the nozzle and the Borda mouthpiece.
!p(v.)~
of viscous forces would make t:.P
between calculation and
7.7 The Nozzle and the Borda Mouthpiece
Figure 7.7-1 illustrates a large tank with a smooth rounded nozzle near
the bottom. lf the nozzle is properly shaped, the streamlines will be nearly
parallel as the fluid issues into the surrounding atmosphere and the pressure
Sec. 7.7 The Nozzle and the Borda Mout
248 Macroscopic Balances-lnertial Effects Chap. 7

We also assume that viscous effe


product of Eq. 7.7-1 with the unit ve~

z J pv.,(v - w) n

L, A,(t)

The momentum flux integral is zero

!
Grovity
it takes on the value
Jpv.,(v -
A,(t)
w)

where A 0 equals the nozzle are~. In


that the velocity profile at the ex1t of
everywhere in the tank to be hydros
becomes
line indicotes the control
volume for mocroscopic balances
- Jpn i dA = (11
.<l.(t) Op
Flg. 7.7-1. Nozzle in a large tank.
Substituting Eqs. 7.7-4 and 7.7-5 into
in the jet will be atmospheric. Since the top surface, at z = h, is moving, we and volumetric fl.ow rate, yield
must be cautious when applying the macroscopic balances.
Vo =
Momentum balance Q-
We start with the complete momentum balance given previously as The following assumptions contribu
Eq. 7.2-10. l. The fl.ow is quasi-steady.

~~ f pv dV
r.<t>
+ f pv(v- w) n dA
A,<t>
= f pg dV
r.<tl
+ f t<> dA
.!ll.(t)
(7.7-1)
2. Viscous forces are negligible.
3. The velocity profile at the noz
proper nozzle design).
We assume first that the leve! in the tank is changed slowly, and we make the 4. The pressure everywhere in th
approximation

q_J pvdV ~O (7.7-2) Mechanical energy balance


dt
r.(t)
This type of assumption classifies the analysis as "quasi-steady," for we are We start with the complete me<
treating an inherently transient system with the steady-state form of the W = O for th.
7.3-27. Noting that
momentum balance. This kind of approximation is a common tool of E,= O, we obtain
engineering analysis, and it should always be accompanied by the question,
"Is this a reasonable approximation ?" This question is rather difficult to
~ J(tpv 2
) dV +
dt r.<t> :
answer; however, it would appear to be a reasonable approximation if the
tank were lar~ and the drainage rate small.t
t After the solution for the flow rate is obtained one may use it to estimate the accelera-
= J V t(o) dA -
A.(t)
tion term, and thus determine the validity of Eq. 7.7-2.
Sec. 7.7 The Nozzle and the Borda Mouthpiece 249

We also assume that viscous effects are negligible and take the scalar
~-~-=-.,..-~--- ---- -z=h product of Eq. 7.7-1 with the unit vector i to obtain

for
z J pv:r;(v- w) n dA = - J pn idA (7.7-3)

L. A , (t) d

The momentum flux integral is zero everywhere, except at the nozzle where
it takes on the value
0 (t)

J
A ,(t)
pv:r;(v - w) - n dA = pvgA 0 (7.7-4)

where A 0 equals the nozzle area. In obtaining this result, we have assumed
that the velocity pro file at the exit of the nozzle is flat. lf we take the pressure
everywhere in the tank to be hydrostatic, the right-hand side of Eq. 7.7-3
beco mes
-Jd
pn idA= (p 0
(1)
+ pgh)A 0 -
Opposite the nozzle
PoAo
At the
(7. 7-5)
0
nozzle exit

Substituting Eqs. 7.7-4 and 7.7-5 into Eq. 7.7-3, and solving for the velocity
top surface, at z = h, is moving, we
and volumetric flow rate, yield
acrosc:oo"tc balances.
Vo = Jih (7.7-6a)

Q = AoJih (7.7-6b)

balance given previously as The following assumptions contributed to this result.


l. The flow is quasi-steady.
dA = J pg dV
r.(l)
+J t<n> dA
d . (t)
(7.7-1)
2. Viscous forces are negligible.
3. The velocity profile at the nozzle is flat (very nearly accomplished by
proper nozzle design).
is changed slowly, and we make the 4. The pressure everywhere in the tank is the hydrostatic pressure.

(7.7-2) Mechanical energy balance

analysis as "quasi-steady," for we are We start with the complete mechanical energy balance given by Eq.
with the steady-state form of the 7.3-27. Noting that W = O for this control volume, and assuming that
is a common too! of Ev = O, we obtain
~t J(tpv
be accompanied by the question,
' This question is rather difficult to
2
) dV + J(tpv )(v- w) n dA
2

be a reasonable approximation if the r.(t) A,<t>


(7.7-7)
small.t
one may use it to estimate the accelera-
= Jv t(n) dA - J p~v n dA
A , (t) .r/0 (1)
Eq. 7.7-2.
250 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.7 The Nozzle and the Borda

Once again we restrict ourselves to quasi-steady flows and neglect viscous We note that the assumption of
forces to obtain rate of change of the kinetic energy

J(ipv~(v - w) n dA = - J pv n dA - J pgzv n dA (7.7-8)


assume that the free surface velocity, v
by Eq. 7.7-11 would not be zero, and 1
A . (l) A . (t) .o>'0 (1) and volumetric flow rate would have
where we have written will apply Bernoulli's equation alo
t<a> = -op 7.6-1 and take v, to be negligible t
cP = gz mechanical energy balance.
The kinetic energy flux term is zero everywhere except at the nozzle where it
has the value Bernoulli's equation

J (tpv2)(v - w) o dA
A.(t)
= lpv~A 0 (7.7-9) Assuming quasi-steady flow and
7.4-18 toa streamline between the fr
The first area integral on the right-hand side of Eq. 7. 7-8 is finite only at the
nozzle and at the free surface where v = w. The velocity at the free surface, ip(v~- v~) +(Po-
v., may be obtained by making a mass balance on the system, which gives The assumption that the cross-section
(7.7-10) the cross-sectional area of the nozzle

where AT is the cross-sectional area of the tank. Evaluation of the area Ao = ~ ~ 1,


integral indicates that it is identically zero. AT Vo

-f A
pv n dA= p0 v,AT- p0 v0 A 0
(t) At the At the
(7.7-11) and we may solve for the velocity an
111
free nozzle Vo =
surface

The second integral on the right-hand side of Eq. 7.7-8 is given by Q=.

J
$
pgzv o dA= -pghv,AT
(1) At the
+ g(z )0v0 A0 =
At the
-pghv,AT (7.7-12)
which is the same result obtained wi1
of Bernoulli's equation leads us to the
0
free nozzle the mechanical energy balance requi
surfacc Note that in Bernoulli's equation the
beca use
it is small compared to the nozzle ve),
(z)0 = O (7.7-13)
balance, the term involving th~ sud
Substitution of Eqs. 7.7-9 and 7.7-12 into Eq. 7.7-8, and solving for the neglected. While the derivation of th
velocity and volumetric flow rate, yield the mechanical energy balance and
(in the first, we formed the scalar pro<
Vo = .J2ih (7.7-14a) one equation was integrated over a
Q = Ao.J2ih (7.7-14b) along a streamline; therefore, we m
different.
The following assumptions contributed to this result. The results obtained by the momer
l. The flow is quasi-steady. balance (and Bernoulli's equation) d
2. Viscous dissipation is negligible. l. H. W. King and E. F. Brater, Handb
3. The velocity profile at the nozzle is flat. Book Company, 1963), p. 4-20.
Chap. 7
Sec. 7.7 The Nozzle and the Borda Mouthpiece 251

We note that the assumption of quasi.:steady ftow led us to neglect the time
rate of change of the kinetic energy in the control volume, but we did not
pv n dA - J
~.(t)
pgzv n dA (7.7-8)
assume that the free surface velocity, v., was zero. Ifwe had, the integral given
by Eq. 7.7-11 would not be zero, anda rather different answer for the velocity
and volumetric ftow rate would have been obtained. In the next section, we
will apply Bernoulli's equation along the streamline indicated in Figure
7.6-1 and take v, to be negligible to obtain the same result given by the
mechanical energy balance.
except at the nozzle where it
Bernoulli's equation

(7.7-9)
Assuming quasi-steady ftow and neglecting viscous effects, we apply Eq.
7.4-18 toa streamline between the free surface and the nozzle, which gives us
side of Eq. 7.7-8 is finite only at the
w. The velocity at the free surface, ip(v~- v~) + (Po - Po) + pg(O- h) = O (7.7-15)
balance on the system, which gives
The assumption that the cross-sectional area of the tank is much larger than
(7.7-10) the cross-sectional area of the nozzle gives

A0 V8
- = - ~1. or v.~ v0
AT Vo

(7.7-11) and we may solve for the velocity and volumetric ftow rate to obtain
Atthe
nozzle Vo .J2ih
= (7.7-16a)

of Eq. 7.7-8 is given by Q = AoJlih (7.7-16b)


which is the same result obtained with the mechanical energy balance. Use
(7.7-12)
of Bernoulli's equation leads us to the final result in just a few steps, whereas
Atthe
nozzle the mechanical energy balance requires a somewhat more careful analysis.
Note that in Bernoulli's equation the surface velocity v, is dropped because
it is small compared to the nozzle velocity v0 , while in the mechanical energy
(7.7-13)
balance, the term involving th~ surface v.elocity very definitely cannot be
into Eq. 7.7-8, and solving for the neglected. While the derivation of the two differential equations that led to
the mechanical energy balance and Bernoulli's equation were very similar
(in the first, we formed the scalar product with v and in the second, with vfv),
(7.7-14a)
one equation was integrated over a volume while the other was integrated
(7.7-14b) along a streamline; therefore, we may expect the results to be somewhat
different.
The results obtained by the momentum balance and the mechanical energy
balance (and Bernoulli's equation) differ by a factor of /2. Experiments1
l. H. W. King andE. F. Brater, Handbook of Hydraulics (New York: McGraw-Hill
is ftat. Book Company, 1963), p. 4-20.
252 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.7 The Nozzle and the Borda Mout

for water under conditions such that inertial effects predominate indicate
Pressure is equal to
v0 ~ 0.98J2gh (7.7-17a) pressure in this regi

Q ~ 0.98 A 0J2gh (7.7-17b)


The excellent agreement between the mechanical energy balance and Pressure is not equal
experiments leads us to believe that either the assumptions were valid, or pressure 1n this regia
there are compensating errors in the solution. This latter possibility is not
considered as often as it should be; in this case, however, assumptions l and
3 are satisfactory, and we conclude that viscous dissipation is indeed negligible
for the reported experiments.
A question naturally arises from this example : "Why is the momentum
equation in error by approximately 40 per cent ?" The answer is that the
pressure in the tank is not everywhere hydrostatic, and an analysis of the
Borda mouthpiece should help illustrate this point.

The Borda mouthpiece

The nozzle in the tank illustrated in Fig. 7.7-1 may be replaced with a
Borda mouthpiece shown in Fig. 7.7-2. The momentum and mechanical
energy balances are applied to the Borda mouthpiece in exactly the same
manner as for the nozzle, and the results are
1 2
A 0) / momentum balance (7.7-18)
Vo = ( Ac Vgii
Fig. 7.7-2. Flow nozzle
mechanical energy balance (7.7-19)
Experimental resultst indicate that
On the basis of the excellent result obtained with the mechanical energy
balan..ce for the velocity v0 at the nozzle, we expect the velocity at the Borda Ae
mouthpiece to be given accurately by Eq. 7.7-19. Note that we choose the Ao
control volume for the Borda mouthpiece to cut the jet at a place where the
streamlines are parallel, thus assuring us that the flow is one-dimensional at The good agreement between calcul
that point and the pressure in the jet is atmospheric. Although the velocity indicates that the assumptions used in
is determined by Eq. 7.7-19, we cannot compute the volumetric flow rate balance are satisfactory for the Bord
because the area of the jet where the flow becomes parallel, A e, is unknown. be that the region over which the pre
lf the momentum balance, contrary to the result for the nozzle, provides smaller by the Borda mouthpiece,
a satisfactory result for the Borda mouthpiece, we may use Eqs. 7.7-18 and assumption.
7.7-19 to calculate Ac and thus the volumetric flow rate. We obtain In Sec. 7.5, 7.6, and 7.7 we have
energy balances, along with Bernoull

and
e:)= 0.50 (7.7-20a)
flow configurations. As expected, t
Bernoulli's equation gave the same re
of interest was made among the mom
Q = 0.50A 0 J2gh (7.7-20b) t See Ref. 1, p. 4-19.
Chap. 7
Sec. 7.7 The Nozzle and the Borda Mouthpiece 253
1

Pressure is equal to the hydrostotic--J


(7.7-17a) pressure in this region
1
(7. 7-17b) 1
1

the mechanical energy balance and 1


Pressure is not equal to the hydrostotic 1

'"""'"" '""'';'" \ -----j


either the assumptions were valid, or
solution. This latter possibility is not
this case, however, assumptions 1 and
viscous dissipation is indeed negligible

example: "Why is the momentum


per cent ?" The answer is that the
hydrostatic, and an analysis of the
this point.

in Fig. 7.7-1 may be replaced with a


. The momentum and mechanical
mouthpiece in exactly the same

(7.7-18)
Flg. 7.7-2. Flow nozzle and Borda mouthpiece.
(7.7-19)
vuau11.-u with the mechanical energy Experimental resultst indicate that
we expect the velocity at the Borda Ae
Eq. 7.7-19. Note that we choose the - = 0.52
Ao
to cut the jet at a place where the
that the flow is one-dimensional at The good agreement between calculated and experimental values of A c
atmospheric. Although the velocity indica tes that the assumptions used in obtaining a solution to the momentum
compute the volumetric flow rate balance are satisfactory for the Borda mouthpiece. The reason must surely
becomes parallel, Ac, is unknown. be that the region over which the pressure is not hydrostatic is made much
to the result for the nozzle, provides smaller by the Borda mouthpiece, thus reducing the error caused by this
;._.-, we may use Eqs. 7.7-18 and
u ... .., .. assumption.
flow rate. We obtain In Sec. 7.5, 7.6, and 7.7 we have used the momentum and mechanical
energy balances, along with Bernoulli's equation, to analyze four different
(7.7-20a) flow configurations. As expected, the mechanical energy balance and
Bernoulli's equation gave the same result for each case and the comparison
of interest was made among the momentum balance, the mechanical energy
(7.7-20b) t See Ref. 1, p. 4-19.
254 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.8 Applications of the Momentum 1

balance, and experimental data. For the sudden expansion in the pipeline, z
the momentum balance gave the superior result, while both balances gave
unsatisfactory results for the sudden contraction. For flow through a nozzle
the mechanical energy balance gave an excellent answer, while the momentum
balance was in error by 40 per cent. Both equations gave valid results for the
L, r--------
1
1
Borda mouthpiece, and both had to be used simultaneously for the volumetric 1
1
flow rate to be determined. A disadvantage of the momentum balance in 1
1
analysing these flows is the necessity of specifying the stress at solid surfaces, 1
and the assumptions made regarding the pressure for the sudden contraction 1
u,n
and the nozzle appear to be the main source of error in the momentum 1!
balance. The mechanical energy balance, on the other hand, does not require
evaluation of the stress at solid surfaces where the scalar product v t<> is
1 =v,=:
~,
zero. A further advantage of the mechanical energy balance is that neglect
!
of viscous effects produces an error of known sign; thus, we usually know
a priori the sign of the error involved in the approximate analysis.
We cannot always choose to use either the momentum or mechanical
energy balance; if forces on fixed solid surfaces are desired, we must use
the momentum balance. In the following section, we shall study sorne
problems of this type. 1
Grovity
L-------
Fig. 7.8-1. Plane jet im .
7.8 Applications of the Momentum Balance
We see from that equation that the
Plane jet impinging on a flat plate
n 't' = i[n.,T.,., -t
Figure 7.8-1 illustrates a two-dimensional plane jet impinging on a flat +i[n.,TZII
plate of width b. We would like to determine the force that the liquid exerts
+k[n.,T.,
on the plate. For steady, incompressible flow and a fixed control volume, we
make use of Eq. 7.2-11, forming the scalar product with the unit vector i, to At the surface of the plate,
obtain

J
A,
pv.,v n dA = J
.,;
i t(n) dA (7.8-1)

and the stress vector becomes


The stress vector is represented as

t<> = - np + n 't' (7.8-2)


or
and if the streamlines are straight and the velocity profiles flat wherever the
i t(n) = -p
jets are cut by the control surface, the stress vector becomes
(7.8-3) For a Newtonian fluid this componen

everywhere except on the plate. In examining the stress vector at the surface
of the plate in contact with the fluid, we must refer to Eq . 4.2-14 in Chap. 4. T.,.,=:
Chap. 7 Sec. 7.8 Applications of the Momentum Balance 255

sudden expansion in the pipeline,


result, while both balances gave
n. For flow through a nozzle
answer, while the momentum
equations gave valid results for the
L. 1
1
"2"
r----------- -1
1
1
simultaneously for the volumetric 1 11--.....- - - - - - - . -
of the momentum balance in 1 1
specifying the stress at solid surfaces, 1 1
1 1
pressure for the sudden contraction 1 1
1"1" 1
source of error in the momentum 1
1
on the other hand, does not require 1
where the scalar product v t<> is 1 L
1
anical energy balance is that neglect 1
known sign; thus, we usually know 1
1 1
1
the approximate analysis. 1 1
1
the momentum or mechanical 1
1
1

!
surfaces are desired, we must use 1
1 lt-'------------'-
section, we shall study sorne 1 1
1 1
L-----------
"3"
.J

Grovity
Fig. 7.8-1. Plane jet irnpinging on a flat plate.

We see from that equation that the term n -r may be expanded as follows:

n 't" = + nvrv., + n.r.,]


i[n.,r.,.,
plane jet impinging on a flat +j[n.,r.,v + nvTvv + n.r.v] (7.8-4)
the force that the liquid exerts
flow and a fixed control volume, we +k[n.,r.,. + nvrv + n.r ]
product with the unit vector i, to At the surface of the plate,
n., = 1

(7.8-1)
nv =O
n. =O
and the stress vector becomes

(7.8-2) t(n) = -np +ir.,.,+ jr.,v + kr.,. at the surface ofthe plate, (7.8-5)
or
velocity profiles flat wherever the
vector becomes i t<n> = -p + r.,., at the surface of the plate (7.8-6)

(7.8-3) For a Newtonian fluid this component of the viscous stress tensor is given by

(av.,)
the stress vector at the surface
must refer to Eq. 4.2-14 in Chap. 4. r.,.,=2. ax (7.8-7)
Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.8 Applications of the Momentu
256

The continuity equation and the fact that the velocity is zero at the solid volume, the stress vector becomes
surface may be used to demonstrate that

(~~) = O at the surface of the plate (7.8-8) tcnl = -np 0 +t


and where t(';,> is the stress in the solid in
i ten> = - p, at the surfaceof the plate (7.8-9) ambient pressure.
The stress produced by atmos
Making use of Eqs. 7.8-3 and 7.8-9, and assuming that v., is zero at points to the area integral on the right-ha
2 and 3, we evaluate the terms in Eq . 7.8-1 to obtain balance yields
- p(v.,)ibhl = PobL - ( p ) pJate bL (7 .8-10) - pvib
Because the velocity profile at point 1 is specified as flat, the force F., that the where AP is the cross-sectional area
fluid exerts on the plate is Sin ce t{';,A P is the force exerted on t
and opposite to the force exerted o
(7.8-11) and Eq. 7.8-14 is equivalent to Eq.
We note that the atmosphere acts on the back side of the plate producing a the proper choice of a control vol u
force of -p 0bL, and the net force exerted on the plate becomes for in this case viscous effects need

F x,net = pvi bh1 (7.8-12)


The effect of ambient pressure
A much simpler method of obtaining this result is to apply the momentum
balance to the control volume illustrated in Fig. 7.8-2. For that control We saw in the Iast example tha
made no contribution to the area int
balance. Since we will encounter tH
"2" is worthwhile now to draw sorne ge
,----- - -- -- -- - - - - - - - -,
1 1
1 1
is a constant p 0 , the stress vector m
1
1
1 t(n) =
1
1 where t,!> will be referred to as the n
1
1 written in terms of the total pressUJ
"1" 1
t(n) =

1 L and the net stress vector is then gi


------1 Cross sect1onol
1
1
oreo= Ap * = -n
t(n)
1 1 where
1
1
1
1
p
1 1
1 1 the gauge pressure. In the absence o
1 1
1 terms of the gauge pressure
1
1
1 1
*
t(o)

l
- _______ J
L-------- --
"3" The area integral of the stress vect
Eq. 7.8-15 to give

Grovity
Fig. 7.8-2. Plane jet impinging on a flat plate.
roscopic Balances-lnertial Effects Chap. 7 Sec. 7.8 Applications of the Momentum Balance 257

that the velocity is zero at the solid volume, the stress vector becomes
that everywhere except at
ttn) = -npo, the solid surface (7.8-13a)
surface of the plate (7.8-8) t 10 = -np 0 + t 10* at the solid surface (7.8-l3b)
where t;';.> is the stress in the solid in excess of that arising from the constant
the surface of the plate (7.8-9) ambient pressure.
The stress produced by atmospheric pressure will contribute nothing
and assuming that u., is zero at points
to the area integral on the right-hand side of Eq. 7.8-1, and the momentum
7.8-1 to obtain balance yields
bL - (p)piatebL (7.8-10) - pvibh = i t~n>Av (7.8-14)
is specified as flat, the force F., that the where AP is the cross-sectional area of the solid section supporting the plate.
Since t;';.>AP is the force exerted on the plate by the solid support, it is equal
and opposite to the force exerted on the plate by the impinging jet of liq u id,
= p0 bL + pvibh (7.8-11) and Eq. 7.8-14 is equivalent to Eq. 7.8-12. The point of this analysis is that
the back side of the plate producing a the proper choice of a control volume can greatly simplify the computation,
on the plate becomes for in this case viscous effects need never be considered.

= pvibh1 (7.8-12)
The effect of ambient pressure
ng this result is to apply the momentum
in Fig. 7.8-2. For that control We saw in the last example that the ambient pressure, being a constant,
made no contribution to the area integral ofthe stress vector in the momentum
balance. Since we will encounter this problem in subsequent applications, it
-------, is worthwhile now to draw sorne general conclusions. Ifthe ambient pressure
1
1 is a constant p 0 , the stress vector may be written as
t(n) = - np0 + t 1: (7.8-15)
where t1!> will be referred toas the net stress tector. The stress vector may be
written in terms of the total pressure p and the viscous stress tensor, 't.
t<n> = - np + n 't (7.8-16)
and the net stress vector is then given by
Cross sect1onol
oreo= Ap ft:> = - n(p - Po) + n 't (7.8-17)
where
P- Po = Pu
the gauge pressure. In the absence of viscous stresses, the net stress is given in
terms of the gauge pressure
- - - _ _ _ _ .J
* = - npu
t(n) (7 .8-18)
The area integral of the stress vector in Eq. 7.2-10 may be written in terms of
Eq. 7.8-15 to give

J ftnl dA = - J np0 dA + Jt~ dA (7.8-19)


mpinging on a flat plate. d.(t) .-1.(1) <~.<t)
258 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.8 Applications of the Momentum

We may use the divergence theorem for a scalar to show that the integral Substituting Eq. 7.8-22 into Eq. 7.8-2
containing the ambient pressure is identically zero. Thus, and j, respectively, we get

f np 0 dA=
.<O
fr .m
Vp 0 dV == O (7.8-20)
-pvgA 0 +p
pv~A 1 si
On the basis ofthis result, we may replace t<> by te!> in Eqs. 7.2-10 and 7.2-11, The macroscopic mass balance gives
and we will neglect the constant ambient pressure in further discussions of the
area integral of the stress vector. pvoAo
e
lf we assume that the angle and the
Force exerted by a liquid jet on a curved vane of the impinging jet are given, we are

The use of jets of Iiquid to m ove turbine blades is of considerable impor-


tance in the theory of turbomachinery. In this example, we will examine the Since we have only the two momentt
forces exerted on a curved vane by a jet of liquid, as illustrated in Fig. 7.8-3. need one more equation if the probl
always, the mechanical energy equati
y If we neglect viscous dissipation and

L. r- -- -----------
1
1
1
1
1
Evaluating the individual terms give1
1
-tpvgA 0 + lpv~
Because the fluid is incompressible, 1
by the mass balance. Application o

Thus, we have our fourth equation l

Fig. 7.8-3. Jet impinging on a curved vane.


F.,= -pv

Fv = pvgA
Making the usual assumptions of steady, incompressible flow and flat
velocity profiles, the momentum equation is Plane jet impinging on an incline

f
A,
(pv)v n dA = Jt~.,
.d
dA (7.8-21) An analysis of the flow illustrate<
example of the use of Bernoulli's eq
The area integral of the net stress vector gives intuitive knowledge of the flow fielc
the jet exerts on the plate and the am
J t~., dA = F (7.8-22) on the plate. We start with the me
neglect gravitational effects, and for
.o/
Chap. 7 Sec. 7.8 Applications of the Momentum Balance 259

for a scalar to show that the integral Substituting Eq. 7.8-22 into Eq. 7.8-21 and taking the scalar product with i
zero. Thus, and j, respectively, we get

(7.8-20) -pv~A 0 + pv:A1 cos (} = FID (7.8-23)


pv;A 1 sin () = F11 (7.8-24)
t<> by t,:> in Eqs. 7.2-10 and 7.2-11, The macroscopic mass balance gives
pressure in further discussions of the
pv0 A 0 = pv1A 1 (7.8-25)
If we assume that the angle () and the density, velocity, and area p, v0 , and A 0
of the impinging jet are given, we are left with four unknowns,

rbine blades is of considerable impor-


In this example, we will examine the
Since we have only the two momentum equations and the mass balance, we
of liquid, as illustrated in Fig. 7.8-3.
need one more equation if the problem is to be solved. Sometimes, but not
always, the mechanical energy equation will supply the needed information.
If we neglect viscous dissipation and gravitational effects, Eq: 7.3-31 becomes

J
A,
2
(ipv )v n dA = - J
A,
pv n dA (7.8-26)

Evaluating the individual terms gives

-ipvgAo + ipv~Al = PoVoAo- PoV1A1 (7.8-27)


Because the fluid is incompressible, the right-hand side of Eq. 7.8-27 is zero
by the mass balance. Application of Eq. 7.8-25 to the left-hand side gives

Vo = V1 (7.8-28)
Thus, we have our fourth equation and the final result is

on a curved vane. F., = - pv~A 0(1 - cos fJ) (7.8-29)


F11 = pv~A 0 sin () (7.8-30)
steady, incompressible ftow and ftat
!S
Plane jet impinging on an inclined plate

= Jt~.,
d
dA (7.8-21)
An analysis of the flow illustrated in Fig. 7.8-4 will provide an interesting
example of the use of Bernoulli's equation and how nicely it fits in with our
intuitive knowledge of the flow field. We wish to determine the forces that
the jet exerts on the plate and the amount of fluid that flows in each direction
(7.8-22) on the plate. We start with the momentum balance given by Eq. 7.2-11,
neglect gravitational effects, and form the scalar product with the unit normal
Sec. 7.9 Moving Control Volumes and U
260 Macroscopic Balances-lnertial Effects Chap. 7

We still require one more equati


we have ex hausted the tools at our di
r------------------ equation always gave the same res
1 however, in this case the difference
1
'' our aid. lf we apply Bernoulli's eqm
1 Streomlines '' lines indicated in Fig. 7.8-4, we obta
1
1 ''
.. ,.. ' '> !p(v~ - v~)+ (p 2 -
/
!p(v~- v~) + (Pa-
/
//
/
//
If we neglect gravitational effects an
/
/ is atmospheric, Eqs. 7.8-36 yield
/
V=

This result is consistent with Eq. 7.


/
/ information by using Bernoulli's e
/
/
/ everywhere, and Eqs. 7.8-32 and 7.8
/

/
/
pv~(A 2 - A 1
/
' V / A=
Fig. 7.8-<C. Plane jet impinging on an inclined plane. The areas and volumetric flow rates a
by Eqs. 7.8-38 and 7.8-39,
and tangential vectors n and A to obtain
A2 = A 1 (1 + cos 0),
2
pv~A 1 sin fJ = Fn (7.8-31)
- pv~A 1 cos () + pv~A 2 - pv~A 3 = Fe (7.8-32) A3 = A 1 (1 - cos 0),
2
where A1 , A 2 , and A3 are the cross-sectional areas of the Iiquid streams at We see that the smaller the angle O
points 1, 2, and 3, respectively. We may assume that the density, velocity, plate, and Q 3 --+- O as O--+- O.
and area (p, V, A1) of the impinging jet are given; thus, the normal force Fn Our treatment of this problem h:
is immediately determined by Eq. 7.8-31. is a way in which the mechanical en
In keeping with the context of this chapter, we set Fe = O, for tangential result given by Eq. 7.8-37. lt invo
forces can only result from viscous effects. We are left with four unknowns- volumes, based on the fact that th
v2, v 3 , A 2 and A3-and only one equation (Eq. 7.8-32). The macroscopic
streams of fluid. The student is en
mass balance yields one more relationship, thought, for it will provide valuable 1
v1 A 1 = v2 A 2 + v3A3 (7.8-33) volumes and will illustrate the necessi
of the streamlines for any system to
and the mechanical energy balance may be applied to give

-tpv~A 1 + !pv~A 2 + !pv~A 3 (7.8-34)


7.9 Moving Control Volumes
= (v1A 1 - V2A 2 - VaAa)Po Flow Problems
The right-hand side of Eq. 7.8-34 is zero by the mass balance, and we
rearrange this equation to obtain Sorne of the aspects of moving c<
analysis of flow through the nozzle
(7.8-35)
Chap. 7 Sec. 7.9 Moving Control Volumes and Unsteady Flow Problems 261

We still require one more equation if we are to solve the problem; yet
we have ex ha usted the tools at our disposal. In previous examples, Bernoulli's
equation always gave the same result as the mechanical energy balance;
however, in this case the difference between these two equations comes to
our aid. If we apply Bernoulli's equation along the top and bottom stream-
lines indicated in Fig. 7.8-4, we obtain
ip(v~- vD + (p2 - p 1) + pg(z2 - z 1) =O (7.8-36a)
(7.8-36b)
If we neglect gravitational effects and note that the pressure in the free jets
is atmospheric, Eqs. 7.8-36 yield
V1 = v2 = v3 (7.8-37)
This result is consistent with Eq. 7.8-35, and we have 'obtained additional
information by using Bernoulli's equation. The velocity is now specified
everywhere, and Eqs. 7.8-32 and 7.8-33 will be simplified to
pv;(A 2 - A1 cos O - A3) = O (7.8-38)
A1 = A2 +A 3 (7.8-39)
on an inclined plane. The areas and volumetric flow rates at points 2 and 3 may now be determined
by Eqs. 7.8-38 and 7.8-39,
A
A2 = -...! (1 + cos 0), Q2 = Q1 (1 + cos O) (7.8-40a)
(7.8-31) 2 2
A
2 - pviA 3 = F1 (7.8-32) A3 = {1 - cos 0),
-...! Q3 = Q 1 (1 - cos O) (7.8-40b)
2 2
areas of the liquid streams at
We see that the smaller the angle O, the less fluid flows backwards on the
assume that the density, velocity,
plate, and Q 3 -+O as O-+ O.
are given; thus, the normal force Fn
Our treatment of this problem has not been altogether honest, for there
is a way in which the mechanical energy balance can be used tq obtain the
, we set F1 = O, for tangential
We are left with four unknowns- result given by Eq. 7.8-37. lt involves a rather subtle choice of control
volumes, based on the fact that there is a streamline separating the two
(Eq. 7.8-32). The macroscopic
streams of fluid. The student is encouraged to give this problem further
thought, for it will provide valuable knowledge in the art ofchoosing control
(7.8-33) volumes and will illustrate the necessity of having a qualitative understanding
of the streamlines for any system to be analyzed.

(7.8-34)
7.9 Moving Control Volumes and Unsteady
Flow Problems
by the mass balance, and we
Sorne of the aspects of moving control volumes were encountered in the
(7.8-35) analysis of flow through the nozzle and the Borda mouthpiece in Sec. 7.7.
Sec. 7.9 Moving Control Volumes and Un!
262 Macroscopic Balances-lnertial Effects Chap. 7

In this section, we shall investigate this type of problem more thoroughly. where we have removed w from the di
Before analyzing sorne specific problems, it will be helpful to examine the is constant. We simplify this result ~
macroscopic balances in terms of a coordinate frame moving at a constant integrals on the left-hand side sum to
velocity, for it is often convenient to view a system as if we were an observer and we obtain
moving with a velocity w rather than someone fixed in space.
In Sec. 5.4 we saw that the continuity equation and the equations of
motion took the form
f
d t pv.. dV
d ,r
+ f
A,
(pv..)v.. n

dp + V (pv..) = O (7.9-1) Because w is constant, the control


dt space, does not change size or shape,
p(~;r + v.. Vv,) = -Vp + pg + fl'V v.. 2
(7.9-2) fixed in space. For this reason, the 1'
written as though the control volume
where In attacking the macroscopic me
V, = V- W (7.9-3)
with the stress equations of motion fa
for an observer moving ata constant velocity w. In a manner similar to that
discussed in Sec. 3.6, we may define streamlines in terms of the moving dv.. Vv
coordinate frame. They are lines viewed by a moving observer that are P( -+v
dt r '
always tangent to v,. The unit tangent vector to such a streamline is given by
and form the scalar product with v.. t
A
r
=.!! = dr, (7.9-4)
v.. ds,
where r .. is the position vector and s.. is the distance along the streamline. If
the flow is steady, as viewed by an observer moving at constant velocity, and if
viscous effects are neglected, the scalar product between A, and Eq. 7.9-2 where the density is taken as consta
may be formed to yield Bernoulli's equation for a moving observer. gradient of a scalar, and noting that

(!pv~ + p + pgz) = c. (7.9-5) Vl


for incompressible flows, we may wri1
The macroscopic balances

The macroscopic mass, momentum, and mechanical energy balances were ~ (!pv~) +V (!pv;v,) =
dt
formulated for arbitrary moving control volumes; they therefore pro vide a
rather general attack on macroscopic problems. When the velocity of the Following Eq. 7.3-12 we write
control vo/ume is a constant, these balances take on especially simple forros.
For the mass balance, we start with Eq. 7.1-7 and substitute Eq. 7.9-3 to v.. (V T) =V
obtain However,
~t f p dV
,r0 (t)
+ f
A,(t)
pv.. n dA = O (7.9-6)
since Vw is zero, and Eq. 7.9-13 becc
Vv..

Substitution of Eq. 7.9-3 into the momentum balance given by Eq. 7.2-10 v.. (V T) =V
yields
~t f pv.. dV
,r.(t)
+ w ~t f p dV
,r0 (t)
+ f (pv..)v,. n dA
A,(t)
Substitution ofEq. 7.9-15 into Eq. 7.9
dissipation yield
(7.9-7)

+w f pv.. n dA
A .(t)
= f pg dV
,r0 (1)
+ f t<> dA
s1'0 (t)
Balances-lnertial Effects Chap. 7 Sec. 7.9 Moving Control Volumes and Unsteady Flow Problems 263

type of problem more thoroughly. where we have removed w from the differential and integral signs because it
it will be helpful to examine the is constant. We simplify this result by noting that the second and fourth
frame moving at a constant integrals on the Ieft-hand side sum to zero on the basis of the mass balance,
a system as if we were an observer and we obtain
\SOJne<me fixed in space.
equation and the equations of (7.9-8)

(pv,) =O (7.9-1) Because w is constant, the control volume "Ya(t), while moving through
space, does not change size or shape, and to the moving observer it appears
(7.9-2) fixed in space. For this reason, the Iimits on the integrals in Eq. 7.9-8 are
written as though the control volume were fixed in space.
(7.9-3) In attacking the macroscopic mechanical energy balance, we must start
with the stress equations of motion for a moving observer,
ocity w. In a manner similar to that
streamlines in terms of the moving
by a moving observer that are p(~; + v, Vv,) = pg +V T (7.9-9)
vector to such a streamline is given by
and form the scalar product with v, to obtain
(7.9-4)
ds,
the distance along the streamline. If ~
dt
(!pv~) + v, V(!pv~) = pg v, + v, (V T) (7.9-10)
moving at constant velocity, and if
product between A, and Eq. 7.9-2 where the density is taken as constant. Expressing the gravity vector as the
for a moving observer. gradient of a scalar, and noting that
pgz) = C8 (7.9-5) V Vr =O (7.9-11)

for incompressible flows, we may write Eq. 7.9-10 as

and mechanical energy balances were ~ (tpv~) +V (!pv;vr) = -V (pcf>v,) + Vr (V T) (7.9-12)


volumes; they therefore provide a dt
problems. When the velocity of the Following Eq. 7.3-12 we write
take on especially simple forms.
7.1-7 and substitute Eq. 7.9-3 to vr (V T) =V (vr T)- Vvr: T (7.9-13)
However,
(7.9-6) Vvr = Vv (7.9-14)
since Vw is zero, and Eq. 7.9-13 becomes
"'"''"u"' balance given by Eq. 7.2-10
vr (V T) =V (vr T)- Vv: T (7.9-15)
dV + J (pv,)v, n dA Substitution ofEq. 7.9-15 into Eq. 7.9-12 and use of<l> to represent the viscous
,(t) dissipation yield
(7.9-7)
pg d V
.(t)
+ J t<n> dA
.d.(t)
~ (!pv;) +V (tpv;vr) =
dt
-V (pcf>v,) +V (vr T)- <l> (7.9-16)
264 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.9 Moving Control Volumes and Uns

Integration over the volume "Ya(t) and application of the divergence theorem
yield

f
"Y0 (t)
~t (ipv;) dV + f (! pv;)vr n dA= -
.#0 (t)
f
0
prf>vr n dA
(t)
Fuel

(7.9-17)

+f Vr ~n) dA - Ev
/
a(t) /
"O" Compres
We ha ve so far only made use of the definition given by Eq. 7.9-3, and we 1
1 /
l. ~-

:~~:-Jr~m
have not yet restricted ourselves to a control volume moving at a constant
velocity. If we now make the restriction that "Ya(t) moves at the constant
velocity w, we may make use of a special form of the general transport _v...::...o- r
theorem referred to in Prob. 3-14 to write 1
\' __Che
f
1'". (t)
~dt (! pv;) dV = ~
dt
f (tpv;) dV
..y.<t)
(7.9-18) ''
' '-..,
' .......

Because all the solid surfaces move at the velocity w, vr is zero at the solid ---
Fig. 7.9-1. T
surfaces and Eq. 7.9-17 reduces to

~t f (!pv;) dV
"Y 0 (1 )
f
+ (~pv;)vr n dA
A , (t)
through the turbine driving the co
nozzle. lt leaves the nozzle close to
velocity than the entering air. If we
(7.9-19)

= f
A, (t)
Vr t(n) dA - f prf>vr n dA -
A , (t)
Ev
balance in the x-direction may be writ

i f (pvr)vr n d
A,
We note that there is no work term in this equation inasmuch as a moving
Assuming that the pressure at the sur
observer sees no moving surfaces. The actual work done is given by
pheric, the net stress t~> results only 1

W= f W~n>dA
A,Ct )
(7.9-20)
(presumably the fue! line would not be
the velocity profiles at O and 1 to be ft:

For this special type of control volume, W is most easily determined by -pov~Ao +1
application of the momentum balance. We notice there is no contribution to t
added to the system, because the x-c01
zero at the surface of the control volm
Analysis of a turbojet engine
PoVoAo +m
Figure 7.9-1 shows a turbojet engine moving through still air ata constant where m1 is the mass ftow rate of the f
velocity, v0 The problem may also be represented as the engine fixed in us to write the force F., as
space and the air entering the diffuser with a velocity v 0 As the air enters the F., = p0v0 A 0 [v 11
diffuser it slows down and the pressure rises. The compressor further where
increases the pressure as the air is forced into the combustion chamber where j

fue! is added and combustion takes place. The combustion mixture passes cx=-
Po
Balances-lnertial Effects Chap. 7 Sec. 7.9 Moving Control Volumes and Unsteady Flow Problems 265

L,
application of the divergence theorem

n dA =- Jpcf>v,
.st.(t)
n dA
Fu el
(7.9-17)

l
/
/
"O" 1 " Compressor
definition given by Eq. 7.9-3, and we
volume moving at a constant (
1 ----
'
j ..... . . . . . ----.. . '
that fa(t) moves at the constant vo J Combustion '
form of the general transport - - - 1 01ffuser
1
1
1
----- ..... __
Chomber

= q_
dt
J(!pv~)
r.(t)
dV (7.9-18)
\ ....

the velocity w, v, is zero at the solid


Fig. 7.9-1. Turbojet engine.

~pv~)v, n dA through the turbine driving the compressor and expands in the exhaust
nozzle. lt leaves the nozzle close to atmospheric pressure and at a higher
velocity than the entering air. If we assume steady flow, the momentum
(7.9-19)
balance in the x-direction may be written
pcf>v, n dA - Ev
i J(pv,)v,
Ae
n dA =Ji t~l
d
dA (7.9-21)
this equation inasmuch as a moving
actual work done is given by Assuming that the pressure at the surface of the control volume is atmos-
pheric, the net stress t~l results only from the stress on the motor mount
(7.9-20) (presumably the fuelline would not be subjected to any stresses). If we take
the velocity profiles atO and 1 to be flat, Eq. 7.9-21 reduces to

W is most easily determined by -p0 v~A 0


=F., + p viA
(7.9-22)
1 1

We notice there is no contribution to the momentum equation from the fuel


added to the system, because the x-component of the fuel velocity vector is
zero at the surface of the control volume. The mass balance requires that
PoVoAo + m, = (7.9-23) PtVtAt
moving through still air at a constant where m1 is the mass flow rate of the fuel. Making use of Eq. 7.9-23 allows
represented as the engine fixed in us to write the force F., as
a velocity v0 . As the air enters the F., = p0 v0 A 0 [v 1(1 + ex) - v0 ]
rises. The compressor further where
into the combustion chamber where m,
ace. The combustion mixture passes cx=-- (7.9-24)
PoVoAo
266 Macroscopic Balances-lnertial Effecu Chap. 7 Sec. 7.9 Moving Control Volumes and U

Because of the addition of fuel and the combustion, the exhaust velocity v1 from the point of view of an observer
is always much greater than the entrance velocity v0 and F,. is positive-i.e., macroscopic mass, momentum and
the force acts in the positive x-direction. We remember that the definition of point of view of an observer moving
the stress vector required that it represent the force per unit area exerted Eqs. 7.9-6, 7.9-8, and 7.9-19. We shall
by the surroundings on the control volume, thus, the force that the engine to discuss the first to gain sorne
exerts on the air frame is the negative of F:c, and we see that the "thrust" of treating moving control volume proble
the engine is in the proper direction. Starting with Eq. 7.2-10,
At this point, the problem has not yet been solved because we must
determine v1 to obtain a relationship between the thrust and velocity. In
practice then, the engineer must make use of the principie of conservation of
~t Jpv dV + J pv(v - w) n
r.(t) A,<t>
energy and knowledge of the combustion process to predict the temperature
ofthe exhaust gases. Once he knows the temperature, he may use an equation we neglect gravitational effects and no
of state to determine the density p1 and Eq. 7.9-23 to evaluate v1 volume is constant to obtain

jet impinging on a moving vane J pv(v- w)


A,(t)
We now wish to extend the problem examined in the previous section to
the case where the curved vane is moving with a constant velocity. This In solving this problem, we will wish to
system is illustrated in Fig. 7.9-2.t We may attack this problem in two ways: the force F; thus, we might first take tl:l
z unit vector i to obtain

Lx V

A W e can now start to see sorne of ti


momentum entering the control volm
r----------------
1 "1" 1 terms describing the momentum leavil
1 1 note first that the fluid velocity vector
1 1
1 that the angle {3 be calculated. Todo 1
1 1
1 1 vr, of the fluid leaving the jet. If we ca
1 1 very shortly that the task of computi
1 1
Ao 1 1 beco mes cumbersome; however if the
1 1
I"Q" 1
a direction either parallel or perpendic
1 difficulty in calculating the momentu
1
.,_wo
- l_ _
If we now switch our attack on 1
moving with the control volume, we s
tational effects to obtain

Fig. 7.9-2. Jet impinging on a moving vane. Making use of Bernoulli's equation al
t Note carefully that the jet leaving the vane is drawn as it would be seen by a moving system, Eq. 7.9-5, we find that the m
observer. Here we have made use of the intuitive idea that Vr should be tangent to the vane. vr is a constant at points O and l. 1
Chap. 7 Sec. 7.9 Moving Control Volumes and Unsteady Flow Problems 267

combustion, the exhaust velocity v1 from the point of view of an observer fixed in space and applying the general
velocity v0 and F., is positive-i.e., macroscopic mass, momentum and mechanical energy balances; or from the
We remember that the definition of point of view of an observer moving with the control volume and applying
the force per unit area exerted Eqs. 7.9-6, 7.9-8, and 7.9-19. We shall use the second approach, but we need
thus, the force that the engine to discuss the first to gain sorne insight into the difficulties encountered in
of F"'' and we see that the "thrust" of treating moving control volume problems.
Starting with Eq. 7.2-10,
yet been solved because we must
between the thrust and velocity. In
use of the principie of conservation of
~t f pv dV
r.<t>
+ f pv(v - w) n dA =
A.(t)
f pg dV
r.<t>
+ f t7n> dA
.JIJ1.(t)
(7.9-25)
process to predict the temperature
temperature, he may use an equation we neglect gravitational effects and note that the momentum of the control
Eq. 7.9-23 to evaluate v1 volume is constant to obtain

f pv(v - w) n dA =F (7.9-26)
A.(t)
examined in the previous section to
with a constant velocity. This In solving this problem, we will wish to determine the x- and z-omponents of
may attack this problem in two ways: the force F; thus, we might first take the scalar product of Eq. 7.9-26 with the
unit vector i to obtain
V
-pv0(v 0 - w0)A 0 + p(v 1 cos {J)(v1 sin {3)( 1 ) =F., (7.9-27)
sm e
We can now start to see sorne of the difficulties in this approach. The
momentum entering the control volume is easily calculated; however, the
terms describing the momentum leaving are not so easy to work with. We
note first that the fluid velocity vector v is not tangent to the vane, requiring
that the angle {3 be calculated. Todo so we must know the relative velocity,
vr, of the fluid leaving the jet. If we carry this problem much further, we see
very shortly that the task of computing the momentum leaving the system
beco mes cumbersome; however if the fluid stream left the control volume in
a direction either parallel or perpendicular to the vector w, we would ha ve no
difficulty in calculating the momentum flux.
If we now switch our attack on this problem to a frame of reference
moving with the control volume, we start with Eq. 7.9-8 and neglect gravi-
tational effects to obtain

f
Ae
(pYr)vr n dA =F (7.9-28)

on a moving vane. Making use of Bernoulli's equation along a streamline in a moving coordina te
is drawn as it would be seen by a moving system, Eq. 7.9-5, we find that the magnitude of the relative velocity vector
idea that vr should be tangent to the vane. vr is a constant at points O and l. Using this fact and taking the scalar
268 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.9 Moving Control Volumes and Unste

product of Eq. 7.9-28 with i and k, we get


- pv~A 0 + p( v, cos O)v,A 1 = F z (7.9-29)
p (v, sin O) v,A 1 = Fz (7.9-30)
where A 1 is the cross-sectional area of the jet and not the area of the exit,
r-------------------
which is A1/sin O. The mass balance gives 1
1

1
1
Ao =Al (7.9-31) 1~~--------~--~
Fx---~
and if we write the magnitude of the relative velocity vector as
(7.9-32)
the components of the force exerted by the vane on the jet are
p(vo _ Wo)2Ao ( 1 _cosO)= -Fz Force exerted by the ( 7 .9 _33)
vane on the jet
Force exerted by the
vane on the jet
(7.9-34)

The power - W generated by the action of the jet on the vane is the scalar 1/1///////////////l//ll///l///llll/!/1~
product of the force -F times the velocity w. Fl. 7.9-3. Movin
2
power = p(v0 - w0) woA 0(1 -cosO) (7.9-35)
Note that if O = O, no force is exerted on the vane and no power is generated. the mass of the control volume results
If O = 1r-i.e., the vane directs the jet back upon itself-the power generated .tbus, Eq. 7.9-36 yields
is a maximum.
d
If we compare Eqs. 7.9-27 and 7.9-35, we can see that a rather large -(pVn 1 o)-
amount of algebraic effort must be expended to reduce Eq. 7.9-27 to the dt
simple form given by Eq. 7.9-35. lt is apparent that visualizing the problem or
rate at which
in terms of a moving coordinate system leads to an easy solution.
water is taken
{
onto the train
Moving water scoop
Forming the scalar product of the mac
unit vector i gives
Figure 7.9-3 illustrates a water scoop used by a fast-moving train to pick
up water from a trough set between the tracks. We would like to know the
rate at which the train takes on water and the drag on the train resulting from ti_
dt
f pvz dV
1"'.(t)
+ f pvz(V -
.A.(t)
this process. The width of the scoop is b and the thickness of the water film
is h. In this example, we will view the problem as observers fixed in space and The velocity of the water in the movin
apply Eqs. 7.1-7 and 7.2-10. In applying the macroscopic mass balance enter the tank at a velocity greater tha
the tank must be w0 , and the first ter
~t Jp dV + Jp(v- w) n dA = O (7.9-36) becomes
1"'0 (1) A.(t)

we note that the fluid velocity v is equal to the velocity of the control volume
w everywhere except in the water, where v = O. The time rate of change of
ic Balances-lnertial Effects Chap. 7 Sec. 7.9 Moving Control Volumes and Unsteady Flow Problems 269

(7.9-29)
(7.9-30)
the jet and not the area of the exit,
,---------------------------------,
L.
(7.9-31)
Fx-----1~
velocity vector as
(7.9-32)
the vane on the jet are
Force exerted by the
(7.9-33)
vane on the jet
Force exerted by the
vane on the jet
(7.9-34) Water

of the jet on the vane is the scalar 1/11111/11111111////111111111111111111111/11//l11111/1//l/1111///1/i1//111//,


Fl1. 7.9-3. Moving water scoop.
1 -cosO) (7.9-35)
the vane and no power is generated. the mass of the control volume results only from the accumulation of water;
upon itself-the power generated -thus, Eq. 7.9-36 yields

we can see that a rather large d


-(pVa o)- pw0 hb =O (7.9-37)
lvnnt111 to reduce Eq. 7.9-27 to the dt
~n,..,.nt that visualizing the problem or
leads to an easy solution. rate at which }
water is tak~n = pwJzb
{
onto the tram

Forming the scalar product of the macroscopic momentum balance with the
unit vector i gives
used by a fast-moving train to pick
tracks. We would like to know the
the drag on the train resulting from
b and the thickness of the water film
ti_
dt
J pv., dV
7".(1)
+ J pv.,(v - w) n dA =
..t.(t)
Ji ~~>dA
d.(t)
(7.9-38)

blem as observers fixed in space and


The velocity of the water in the moving tank is not constant because it must
the macroscopic mass balance
enter the tank ata velocity greater than w0 ; however, the average velocity in
the tank must be w0 , and the first term on the left-hand side of Eq. 7.9-38
- w) ndA =O (7.9-36) becomes

to the velocity of the control volume -d


dt
J pv., dV = -d (p Va 1 oW0)
7".(1) dt
(7.9-39)
v = O. The time rate of change of
270 Macroscopic Balances-lnertial Effects Chap. 7 Sec. 7.9 Moving Control Volumes and U

The momentum flux term is identically zero, since either


V=W
or
v., =o
everywhere on the control surface. Ifwe neglect the small increase in pressure
in the water film, the net stress acting on the system is the result of only the ,------------------
1
1
applied force F.,, and Eq. 7.9-38 yields 1
: Compress
d 1
- (p vH.oWo) = F"' (7.9-40)
dt
Because w0 is a constant, we may use the mass balance to obtain the final
result,t
(7.9-41)
This force must be equal and opposite to the force the water exerts on the
scoop. It is exerted on the curved portion of the scoop where the water is
suddenly accelerated from rest to a velocity somewhat greater than w0 : For
Fig. 7.9-4. Accelen
a train moving ata high velocity, inertial effects probably predominate, and
Eq. 7.9-41 is a reasonable solution. On the other hand, velocity gradients at
the leading edge of the scoop will be large and viscous effects could be We assume:
important if a very accurate answer is desired. l. No friction occurs between the
2. The velocity profile in the emerg1
The accelerating water tank The mass of the control volume, exclud
the water in the control volume at
Another application of the mass and momentum balances for moving macroscopic mass balance gives
control volumes is the accelerating water tank shown in Fig. 7.9-4. Inasmuch p(v- w) nln
as the jet leaves the tank in a direction parallel to the velocity of the control
volume, we will have no difficulty in determining the momentum flux term; where A 0 is the area of the nozzle.
however, if the jet issued from the tank in sorne direction other than that momentum balance with the unit vecto
shown, the problem would be a great deal more complicated. effects in the surrounding air, we get
The tank contains a supply of compressed air regulated so that the flow
rate leaving the tank through the nozzle is constant; thus, ~t f
r.(tJ
pv., dV
A,<t>
+ f pv.,

~ (pVH.o) = -m (7.9-42) We now assume that the velocity e


dt is w. This is not strictly correct beca
where nozzle; however, if the amount of fiui
m= a positive constant through the nozzle is small, it may be ~
VH = the volume of water in the tan k use of Eq. 7.9-43 and the above assum
0
t This result can be deduced intuitively from the linear momentum principie since the d
force Fz is just equal to the time rate of change of momentum; i.e., (pw 0 )(w0hb) = momen- - (Mw 0
dt
+ pVH oWo 2
tum per unit volume x volume per unit time.
Sec. 7.9 Moving Control Volumes and Unsteady Flow Problems 271

zero, since either

neglect the small increase in pressure


on the system is the result of only the --------------------------------,1
1
1 1
1 1
1 1
: Compressed oir 1
1
(7.9-40)

the mass balance to obtain the final i - - - - wo(fl

(7.9-41)

of the scoop where the water is


somewhat greater than w0 ." For
Fig. 7.9-4. Accelerating water tank.
effects probably predominate, and
the other hand, velocity gradients at
be large and viscous effects could be We assume:
desired. l. No friction occurs between the wheels and the solid surface.
2. The velocity profile in the emerging jet is flat .
The mass of the control volume, excluding the water, is M, and the volume of
the water in the control volume at zero time is Vo. Application of the
and momentum balances for moving macroscopic mass balance gives
tank shown in Fig. 7.9-4. lnasmuch p(v - w) . nJnozzle Ao = m (7.9-43)
parallel to the velocity of the control
determining the momentum flux ter m; where A 0 is the area of the nozzle. Forming the scalar product of the
in sorne direction other than that momentum balance with the unit vector i and neglecting viscous and inertial
deal more complicated. effects in the surrounding air, we get
lJJauit:~~c:u air regulated so that the flow
is constant; thus, ~t f f
pvx dV + pvx(v - w) n dA = O
r .(t) A,<tl
(7.9-44)

(7.9-42) We now assume that the velocity everywhere inside the control volume
is w. This is not strictly correct because the fluid is moving towards the
nozzle; however, if the amount of fluid in the tan k is large and the flow rate
through the nozzle is small, it may be a reasonable approximation. Making
use of Eq. 7.9-43 and the above assumption allows us to write
from the linear momentum principie since the
ofmomentum; i.e., (pw 0)(w0hb) = momen- (7.9-45)
Macroscopic Balances-lnertial Effecu Chap. 7 Sec. 7.1 O Differentiai-Macroscopic Balances

Carrying out the differentiati~n gives


7.1 O Differentiai-Macroscopic Ba
(M
0
+ p Vulo) :: + Wo :t (p Vulo) + v:~:lnozzle m = o (7.9-46)
The title of this section seems som
which may be simplified by the use of Eq. 7.9-42 to yield point we have only examined differenti
macroscopic equations (always referred
. d~ 1 . (7.9-47) them from differential equations). In ob
(M+ pVa1o)- + (v"'- W0) nozzle m= O
dt integrated the differential equations ove
Since n = -i at the nozzle, we may use Eq. 7.9-43 to simplify this result volume-averaged versions of the differe1
further possible, and it is very practica! in certai
over an area A a(t) and thus obtain are
(M+ pValo)dwo- (m)2 =O (7.9-48) equations. For practica! purposes it is
dt pA 0 equations by applying the macroscopi<
We must now solve for Vu 10 as a function oftime, and we therefore integrate l!.x and then Ietting ~x ~ O.
Eq. 7.9-42 Once the method has been outlined
pVa 1o = -mt + C1 (7.9-49) in formulating these balances for speci
have difficulty in recognizing when
and impose the boundary condition should be applied instead of the
B.C. 1: t=O (7.9-50) followed in this case follows:

to obtain If sorne quantity such as the pre


pVu 10 = (pV0 - mt) (7.9-51) continuously in the x-direction (for ex
this variation, the macroscopic bal
Substitution of Eq. 7.9-51 into Eq. 7.9-48 and separation of the variables the x-direction.
give
(7.9-52) We will apply this idea to the rec
7.10-1. The end of this channel is close
where material that allows fluid to be draw
Mo=M+pV0 technique or another the fluid can b
which integrates to
Porous surface through
w0 = - ( m ) ln (Mo- mt) + C2 (7.9-53) which uniform distributio
pAo takes place
The boundary condition, or "initial condition" as it is sometimes called,
will be specified as
B.C. 2: w0 =O, t=O (7.9-54)
and the final result is
Oo----
w0 = (..!!:.__) In ( Mo ) (7.9-55)
pA0 M0 - mt
This solution is meaningful only for instances such that mt < p V0 , beca use f.-.- - -- - L - -
for Iarger times there is no water left in the tank according to Eq. 7.9-51.
Fig. 7.10-1. Flow in
Sec. 7.10 Differentiai-Macroscopic Balances 273

7.1 O Differentiai-Macroscopic Balances


(pVn.o) + v.,lnozzle m= O (7.9-46)
The title of this section seems somewhat contradictory, for up to this
Eq. 7.9-42 to yield point we have only examined differential equations (in Chaps. 2 and 5) and
macroscopic equations {always referred to as "balances" to help distinguish
(v., - Wo)lnozzle m= 0 (7.9-47) them from differential equations). In obtaining the macroscopic balances, we
integrated the differential equations over a volume i '4 (t) , and thus obtained
use Eq. 7.9-43 to simply this result vo/ume-averaged versions of the differential equations. 1t would have been
possible, and it is very practica! in certain problems, to integra te the equations
(m)2 over an area A 4 (t) and thus obtain area-averaged versions of the differential
--=0 (7.9-48) equations. For practica! purposes it is easier to obtain the area-averaged
pAo
equations by applying the macroscopic balances to a volume of thickness
of time, and we therefore integra te ~x and then Ietting ~x --+ O.
Once the method has been outlined, the student will have no difficulty
(7.9-49) in formulating these balances for special problems; however, students often
have difficulty in recognizing when the differential-macroscopic balance
should be applied instead of the macroscopic balance. The rule to be
t=O (7.9-50) followed in this case follows:

If sorne quantity such as the pressure or average velocity varies


(7.9-51) continuously in the x-direction (for example), and we wish to determine
.9-48 and separation of the variables this variation, the macroscopic balance must be made differential in
the x-direction.

(7.9-52) We will apply this idea to the rectangular channel illustrated in Fig.
7.10- l. The end of this channel is el o sed, and one si de is made of a porous
material that allows fluid to be drawn off. We will assume that by one
technique or another the fluid can be removed uniformly. Under these

(7.9-53) Porous surfoce through


which uniform distribution
tokes place
as it is sometimes called,

t=O (7.9-54)

z
(7.9-55)

ps1cances such that mt < p V0 , beca use


in the tank according to Eq. 7.9-51.
L:',
Fig. 7.10-1. Flow in a porous channel.
274 Macroscopic Balances-lnertial Effects Chap. 7 Problems

conditions, the volumetric flow rate in the channel is given by where

Q(x) = Q0 (1- i) (7.10-1)

and the volumetric flow rate per unit Iength through the porous surface is Use of Eq. 7.10-8 allows us to write

Qo q = (7.10-2)
2p~.,)~( 1
L
We will assume that the flow is turbulent so that the flat velocity profile which is easily integrated to obtain
assumption can be made; viscous effects will be neglected. We now wish to (p)., = (p)o +2
determine the pressure as a function
of x, and we therefore need to develop This result indicates that the pressure
a differential-macroscopic balance in entrance ofthe channel to (p )0 + p(v.,
the x-direction. We do so by applying macroscopic balances present no seri~
Eq. 7.2-11 to the control volume when they must be used.
shown in Fig. 7.10-2. Neglecting
viscous effects and dotting Eq. 7.2-11
PRO
with i, we get

f
A,
pv,v n dA =-
.9/
f n ip dA
(7.10-3)
7-1. Make use of the continuity equat1
the stress equations of motion gi
by Eq . 4.4-13.
Application to the control volume
gives 7-2. lf

Fig. 7.10-2. Differential control volume. -p(v!)AJ., + p(v!)AJ.,+a.,


= (p)AJ., - (p)AJ.,+a., (7.10-4)
f
A,!t)

and the flow is incompressible, sh


Because the area is constant we may divide by A tl.x to obtain
balance may be written as

(7.10-5) ~ fG
r.u l
pv
2
) dV +
Letting tl.x - O yields an equation that is macroscopic in the y- and z-
directions, but differential in the x-direction. =f V

A , (t)
ttn) dA -

p !!..._ (<v!>) = - d(p) (7.10-6)


dx dx 7-3. A hydraulic brake that consists
slightly larger cylinder is show:
Assuming flat velocity profiles gives
specified as u0 and we wish to 1
motion. Neglect viscous effects a
2 p(v.,) d(v.,) = _ d(p) (7.10-7) cal energy balances to obtain t\-1
dx dx
that must be made in order to ol
and Eq. 7.10-1 can be used to obtain Ans: The mechanical energy bah

(v.,) = (v.,) 0 ( 1 - i) (7.10-8) F=


Problems 275

the channel is given by where

(7.10-1) (v ) = Qo
"'o A
length through the porous surface is Use of Eq. 7.10-8 allows us to write Eq. 7.10-7 in the form

2p(v.,)~( 1 _ ~) = d(p) (7.10-9)


(7.10-2)
L L dx
so that the flat velocity profile which is easily integrated to obtain

<P>., = <P>o + 2p(v.,>~[z- !(zn


will be neglected. We now wish to
(7 .10-1 O)
determine the pressure as a function
of x, and we therefore need to develop
This result indicates that the pressure increases continuously from (p ) 0 at the
a differential-macroscopic balance in
the x-direction. We do so by applying
entrance ofthe channel to (p ) 0 +
p ( v., )~ at the end ofthe channel. Differential-
macroscopic balances present no serious difficulties provided we can recognize
Eq. 7.2-11 to the control volume
when they must be used.
shown in Fig. 7.10-2. Neglecting
viscous effects and dotting Eq. 7.2-11
with i, we get PROBLEMS

f pv.,v
A,
n dA = -
.1'1
f n ip dA
(7.10-3)
o 7-1. Make use of the continuity equation and the material derivative to show that
the stress equations of motion given by Eq. 7.2-3 are identical to those given
by Eq. 4.4-13.
Application to the control volume
gives 7-2. If

-p(v!>A. + p(v!>Aiz+&z
= (p)AI.,- (p)AJ.,+&., (7.10-4)
f
A.,(t)
v ndA = O

divide by A llx to obtain and the flow is incompressible, show that the macroscopic mechanical energy
balance may be written as
(p)Jz+&z (p)J.,
(7.10-5)
~ fG
fa(t)
pv
2
) dV + JG
A , (l)
pv
2
) (v - w). n dA

m the y- and z-

= J V t(n) dA - J prpV n dA + W- v
A , (t) d"(l)
(7.10-6)
7-3. A hydraulic brake that consists of a cylindrical ram displacing fluid from a
slightly larger cylinder is shown in Fig. 7-3. The velocity of the ram is
specified as u0 and we wish to know the force F required to maintain this
=- d(p) (7.10-7)
motion. Neglect viscous effects and use both the momentum and the mechani-
dx cal energy balances to obtain two answers. State carefully the assumptions
that must be made in order to obtain a solution.
Ans: The mechanical energy balance gives
(7.10-8) pu~( 1T Di/4)
F = 2[(D 1 / D0) 2 - 1F
276 Macroscopic Balances-lnertial Effects Chap. 7 Problems

Fi

Flg. 7-l. The hydraulic ram. 7-6. A very simple device for pumping
7-6. The fluid to be pumped is
(instead of the force supplied b
7-4. A jet of water issues from a 1-in. I.D. nozzle at a velocity of 30 ft/sec and sec 0 ndary fluid issuing through th
strikes a flat plate moving away from the jet at 17 ft/sec, as illustrated in Fig. place va viscous effects, and the
7-4. What force does the water jet exert on the plate?
"1''

17ft/sec

l
Gravity
Fig

energy balance cannot be neglectec


Fig. 7-4 be neglected if the flow is turbule
pump can be obtained with the mo1
the pressure rise between points 1
7-5. Water issues from a pipe imbedded in a concrete wall as shown in Fig. 7-5. nozzle area A 0 while maintainin
lfthe velocity leaving the pipe is 20 ft/sec and the cross-sectional area is 1 in. 2, Assume that the secondary fluid is
what are the components of the resultant force exerted on the pipe by the 7-7. The Egyptians reportedly used wa
wall? Assume that the pressure drop from the wall to the end of the pipe is 7-7. The radius r 0 of the circular
negligible. the bottom of the bowl. Determin'
quired if the depth of the liquid is
Ans: F.,= 0.72Ib1 , F11 = 2.70 lb1
F

Fig. 7-5

hydraulic ram. 7-6. A very simple device for pumping fluids is the ejector pump illustrated in Fig.
7-6. The fluid to be pumped is called the primary fluid, and momentum
(instead of the force supplied by mechanical pumps) is supplied by the
I.D. nozzle at a velocity of 30 ft/sec and secondary fluid issuing through the nozzle. The momentum exchange takes
the jet at 17 ft/sec, as illustrated in Fig. place va viscous effects, and the viscous dissipation term in the mechanical
exert on the plate?
"f'

01
r
Area= A
:
---- --1--

17ft/sec

l
Grcvity
Fig. 7-6

energy balance cannot be neglected. However, the viscous surface forces can
7-4 be neglected if the flow is turbulent and a satisfactory design of an ejector
pump can be obtained with the momentum balance. Derive an expression for
the pressure rise between points 1 and 2, and note the effect of reducing the
in a concrete wall as shown in Fig. 7-5. nozzle area A 0 while maintaining the secondary flow rate Q0 constant.
ft/sec and the cross-sectional area is 1 in. 2, Assume that the secondary fluid is the same as the primary fluid.
force exerted on the pipe by the 7-7. The Egyptians reportedly u sed water clocks similar to that illustrated in Fig.
from the wall to the end of the pipe is 7-7. The radius r 0 of the circular bowl is a function of z, the distance from
the bottom of the bowl. Determine the functional dependence of r0 on z re-
quired if the depth of the liquid is to be a linear function of time.
278 Macroscopic Balances-lnertial Effects Chap. 7 Problems

7-8. Ifthe valve on the pipe shown in Fi


elbow will decrease suddenly. The 1
for very short times when the velo
effects are negligible. Determine p
Hint: Two scalar components of ti
Ans: p* = p0 + pgL 1 L 2/(L 1 + L 2)
7-9. A standard method for determinin
of measuring the pressure differenc
Fig. 7-9. Derive an expression for

1"""""'""'-'-'-~"'"'~"
Areo = A1

'

Flg. 7-7

Fi

and 'the pressure difference measu


y the relative merits of the momentu

Lx
is much superior to the other for t
7-10. What is the resultant force at the C<
the pipe in Fig. 7-10, because ofth
L1 velocity in the 2-in. l. D. section is
7 lbm, and the interior volume is :
7-11. The device illustrated in Fig. 7-11
p mine the resultant horizontal forc
leaving fluid streams.
Given:
1 PA
1 PB
-f--- gauge pressure at A = 15 psig
gauge pressure at B = 20 psig
gauge pressure at e = o
L2
The average velocity atA is 10 ft/s
Flg. 78 Ans: Fx = -28 lb1
7-8. Ifthe valve on the pipe shown in Fig. 7-8 is opened rapidly, the pressure at the
elbow will decrease suddenly. The problem is a complex one but can be solved
for very short times when the velocity may be set equal to zero and viscous
etfects are negligible. Determine p* for short times after the valve is opened.
Hint : Two scalar components of the momentum equation must be used.
Ans: p* = p0 + pgL 1 L 2 /(L 1 + L 2)
t 7-9. A standard method for determining turbulent flow rates in conduits consists
of measuring the pressure ditference across a flow nozzle such as that shown in
Fig. 7-9. Derive an expression for the volumetric flow Q in terms of A 1 , A2 ,

Areo = A1 A 2 =oreo

7-7

Fig. 7-9

and the pressure ditference measured by the manometer. Consider carefully

L,
the relative merits of the momentum and mechanical energy balances, for one
is much superior to the other for this particular analysis.
7-1 O. What is the resultant force at the connection between the nozzle assemb1y and
the pipe in Fig. 7-1 O, beca use of the water issuing from the jet? The average
velocity in the 2-in. l. D. section is 1Oft/sec, the mass of the nozzle assembly is
7 lbm, and the interior volume is 38 in.3
7-11. The device illustrated in Fig. 7-11 is used to mix two miscible fluids. Deter-
mine the resultant horizontal force on this device owing to the entering and
Ieaving fluid streams.
Given:
PA = 70 lbm/ft 3
PB = 59 lbm/ft 3
gauge pressure at A = 15 psig
gauge pressure at B = 20 psig
gauge pressure at e = o
E--- - L2 - -- - -- -.-
The average velocity atA is 10 ft/sec; at B, it is 15 ft/sec.
Ans: Fx = -28 lb1
280 Macroscopic Balances-lnertial Effecu Chap. 7 Problems

L.
Grovity

Flg. 7-10
rate is the same in both cases.) E
funtion of O, and assume that the m
over the surface of the exit as indi
Fig. 7-13b for assumption (b).

momentum flux (a) 1 ~l1 -


Ans: momentum flux (b) = 2 (1 - <

8
0=1in.
Flg. 7-11

7-12. Water is being pumped into a tank as shown in Fig. 7-12 ata rate of 8 ft 3/sec.
lfthe diameter ofthe pipe is 6 in. and the inside diameter ofthe tank is 30 in.,
what is the force required to hold the tank in position for h equal 24 in ..
excluding the gravitational force acting on the tank itself?
7-13. In solving macroscopic balance problems, we always attempt to place the
suace of the control vo1ume so that the streamlines are parallel and the
momentum flux is easily determined. As an example of the error that may
occur if the streamlines are not parallel, compute the momentum flux at the
exit of a circular diverging section assuming (a) the stream1ines are straight
lines converging at point O, and (b) the streamlines are parallel. (Note: the (b)
magnitude ofthe ve1ocity vector must be adjusted so that the volumetric flow Fig.
lcrclScc,Dic Balances-lnertial Effects Chap. 7 Problems 281

L. t
h(t)

-t 12in.
~~~--~=---~_l

Fig. 7-12

7-10
rate is the same in both cases.) Express the ratio of momentum fluxes as a
function of O, and assume that the magnitude of the velocity vector is constant
over the surface of the exit as indicated in Fig. 7-13a for assumption (a) and
Fig. 7-13b for assumption (b).

Ans:
momentum flux (a)
momentum flux (b)
= -
1 -l - J
cos2 O
2 (1 - cos 0) 2
In (secO)

0=2in.

7-11

shown in Fig. 7-12 ata rate of 8 ft 3/sec.


the inside diameter of the tank is 30 in.,
the tank in position for h equal 24 in.,
on the tank itself?
we always attempt to place the
the streamlines are parallel and the
"o"~-:_--1!__-
....................
-----
,,
As an example of the error that may
compute the momentum flux at the
assuming (a) the streamlines are straight
the streamlines are parallel. (Note: the
be adjusted so that the volumetric flow Fig. 7-ll
282 Macroscopic Balances-lnertial Effects Chap. 7 Problems

7-14. Fluid distributing systems often consist of an array of perforated pipes


providing discrete distribution at any number of points. A single perforation
Air at 8 psig
in a pipe is shown in Fig. 7-14 . lf the diameter of the hole is smaller than the
pipe-wall thickness, the jet is very nearly perpendicular to the main stream.
Assuming that the volumetric ftow rate through the perforation can be 8ft _::::;>
expressed ast Q 0 = A 0 v 2(p 1 - p 0 ) / p, where Q 0 is the volumetric flow rate
through the hole, A 0 is the area of the hole, and p 0 is the ambient pressure, use
the momentum and mechanical energy balances to obtain an expression for
the pressure ditference, p 1 - p 2 Neglect viscous etfects and explain wh y you
Air at 11 psig
obtain two ditferent answers .

---.,.:e.-

1 in.

Fig. 7-14

7-15 . Calculate the discharge rate from the upper reservoir to the lower reservoir
in Fig. 7-15. Take the fluid to be water and neglect viscous etfects.
v0 = 25 ftlsec -
-.-t_______,
Ans: Q = 0.18 ft 3 /sec.
7-16. Does a converging nozzle on a garden hose place the hose in tension or compres-
sion at the junction between the hose and the nozzle? Don 't guess-analyze.
7-17. The inclined flat plate in Fig. 7-17 is moving toward the plane jet at 10 ft/sec,
and the jet velocity (relative toa fixed reference frame) is 25 ftsec . lf the jet is
1 in. thick, determine the force per unit width exerted on the plate by the stream
of water. F
Ans: fx = -106 lb11ft
7-18. Two miscible turbulent streams of densities p1 and p2 are flowing in a wide,
rectangular duct separated by a thin plate. Neglecting viscous effects, use the
momentum balance to calculate the pressure change in the mixing region in
terms of (v.), (v. )2 , h1 , and h2 . Assume the pressure is uniform across the
channel and the velocity profiles of the unmixed and completely mixed
streams are flat.
t This is an approximate version of a formula to be derived in Chap. 8. Fl1. 7-18. Mixing
Balances-lnertial Effects Chap. 7 Problems 283

of an array of perforated pipes


number of points. A single perforation
Air ot 8 psig
diameter of the hole is smaller than the
nearly perpendicular to the main stream.
rate through the perforation can be 8ft __::::>
p, where Q0 is the volumetric flow rate
hole, and p0 is the ambient pressure, use
rgy balances to obtain an expression for
eglect viscous effects and explain why you
Air ot 11 psig Nozzle diometer
= 2in

Fig. 7-15

1 in.

v0 = 25 ft/sec -
the upper reservoir to the lower reservoir --~-------------

and neglect viscous effects.

hose place the hose in tension or compres-


and the nozzle? Don't guess-analyze.
is moving toward the plane jet at 10 ft/sec,
reference frame) is 25 nsec. If the jet is
it width exerted on the plate by the stream
Fig. 7-17

f densities p1 and p2 are ftowing in a wide,


in plate. Neglecting viscous effects, use the
he pressure change in the mixing region in
~ssume the pressure is uniform across the
~ of the unmixed and completely mixed

~rmula to be derived in Chap. 8. Flg. 7-18. Mixing of two turbulent streams.


284 Macroscopic Balances-lnertial Effects Chap. 7

7-19. A laminar jet ofwater issues from a horizontal nozzle as shown in Fig. 7-19.
The velocity profile of the jet as it emerges from the nozzle is parabolic, but
the profile becomes flat sorne distance from the nozzle owing to viscous
effects. Neglect the effect of gravity and the ambient air and derive an
expression for the final area A 1 .
Ans: A 1/A 0 =!
Macroscopic j
Nozzle oreo, A 0
Visco

---------
Fluid jet
orea, A,
Fig. 7-19. The laminar jet.

7-20. Use the mechanical energy balance (rather than Bemoulli's equation) to show
that v1 = v2 = v3 for the flow shown in Fig. 7.8-4.
7-21. Solve the problem of the moving water scoop illustrated in Fig. 7.9-3 by
using Eq. 7.9-8 instead of Eq. 7.2-10. In Chap. 6, we presented a qualit
7-22. Derive Eq. 7.10-6 by integrating the x-direction stress equation of motion showed that the time-averaged stress
over the y,z-surface. Neglect viscous effects. as the instantaneous stress equations
equations could not be used to det
because the dependence of the turb
velocity is unknown. Empirical exp
length equation, may be used to detc
cannot be applied to any arbitrary gc
In solving turbulent flow problem:
are important, we will always require
of this chapter is the formulation e
experimental data and the applicatior
viscous effects must be considered.

*8.1 Friction Factors-Deflnit

In examining the momentum at


7.2-10 and 7.3-27), we see that knowJ,

Jten> dA a
""(t)
2
Chap. 7

horizontal nozzle as shown in Fig. 7-19.


from the nozzle is parabolic, but
from the nozzle owing to viscous
and the ambient air and derive an

Macroscopic Balances:
Viscous Effects 8

than Bernoulli's equation) to show


in Fig. 7.8-4.
scoop illustrated in Fig. 7.9-3 by
In Chap. 6, we presented a qualitative description of turbulent ftow and
x-direction stress equation of motion showed that the time-averaged stress equations of motion too k the same forro
effects. as the instantaneous stress equations of motion. However, the time-averaged
equations could not be used to determine time-averaged velocity profiles,
because the dependence of the turbulent stress i<t> on the time-averaged
velocity is unknown. Empirical expressions, such as the Prandtl mixing
length equation, may be used to determine velocity profiles; however, they
cannot be applied to any arbitrary geometry and their usefulness is limited.
In solving turbulent ftow problems under conditions where viscous effects
are important, we will always require sorne experimental data. The objective
of this chapter is the formulation of a consistent method of interpreting
experimental data and the application of this result to sorne problems where
viscous effects must be considered.

*8.1 Friction Factors-Definition

In examining the momentum and mechanical energy balances (Eqs.


7.2-10 and 7.3-27), we see that knowledge of the integrals

f t<nl dA
""(t)
and f <1>
?'".(t)
dV

285
286 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.1 Friction Factors-Definition

is required if we are to obtain satisfactory solutions to these equations. In For convenience, the total drag force
the previous chapter, viscous effects were neglected and these integrals were force
easily evaluated. While reasonable results were obtained for a variety of
problems, the flows investigated were restricted to cases for which we knew Frorm =A J
a priori that inertial effects predominated. In accounting for viscous effects, A,+A,
we must relate the surface stress integral and the dissipation integral to
experimentally determined pressure drops and flow rates.
A flowing fluid will exert a force on the solid surfaces that it contacts. In
Ftrlctlon = -A f
A,+
this chapter we are primarily interested in the component of this force in the The drag force contains the negati~
direction of the mean flow, and it will be helpful to define a unit vector A beca use it is defined as a force the ftu
which points in the direction of the mean flow Eq. 8.1-3 represent the force whicl
Because ofthe way in which the drag
(8.1-1) n in Eqs. 8.1-4 through 8.1-6 is direc
The separation of the drag force
Before defining the drag or friction force that the fluid exerts on the solid, made because the former depends rr
Eq. 7.2-10 will be rewritten in a form which is especially convenient for proportional to pu~, while the latter
interpreting experimental data. Substituting Eqs. 7.3-7, 7.8-15, and 7.8-20 proportional to .tu 0 The word "forn
into Eq. 7.2-10, and using the divergence theorem for a scalar, we get of the drag force is greatly influencc
while the word "friction" indica tes th
~~ Jpv dV + Jpv(v -- w) n dA primarily on the area of the so lid s
The drag force for flow in clos
-;'"o(t) A,(t)
(8.1-2) correlated in terms of a dimensionle
= J{-n[(p - p0 ) + pcp] + n T} dA
!=
.llfo(t)

provided the density is constant. It should be kept in mind that for turbulent where
flow, T represents i + i(t>, and p represents p. Splitting the area integral on A* = a characteristic a
the right-hand side of Eq. 8.1-2 into the area of entrances and exits A.(t) and KE* = a characteristic 1(
the area of solid surfaces (both fixed and moving) A, + A,(t), we get
There are many important processe!
~~ Jpv dV +Jpv(v- w) n dA= J{-n[(p- p + pcp] + n T} dA 0)
conduits of constant cross-section2
teristic kinetic energy per unit vol
-;'"G(t) A,(t) A,(t)
given by
(8.1-3)
KE*
+ J{ -n[(p - p0) + pe/>]} dA + J n T dA
A*= t'
A,+A,(t) A,+A,(t)
Because A will be orthogonal to th
We now define the drag force F D as the force which the fluid exerts on the so lid surfaces, F n is given by
so/id in the A-direction.t
Fn=
Fn = ). J n[(p - p 0) + pcp] dA - A .J n T dA (8.1-4)
for conduits having a constant eros
A,+A,(t) A,+A,(t)
t In the design of airfoils, an engineer is naturally interested in a lift force, the force only for mathematically smooth co
the fluid exerts on the solid in a direction perpendicular to A.. for real conduits.
Balances-Viscous Effects Chap. 8 Sec. 8.1 Friction Factors-Definition

solutions to these equations. In For convenience, the total drag force is split into a form force and a friction
neglected and these integrals were force
were obtained for a variety of
r.otrut .. n to cases for which we knew

In accounting for viscous effects,


Frorm =A J n[(p- p0 )
A,+A,(t)
+ pcp] dA (8.1-5)

and the dissipation integral to


and flow rates.
the solid surfaces that it contacts. In
Frrlction = -A f n 't' dA
A,+A,(t)
(8.1-6)

in the component of this force in the The drag force contains the negative of the Iast two terms in Eq. 8.1-3,
be helpful to define a unit vector A beca use it is defi.ned as a force the fluid exerts on the so/id, while the terms in
flow Eq. 8.1-3 represent the force which the surroundings exert on the fluid.
Because ofthe way in which the drag force has been defi.ned, the normal vector
(8.1-1) n in Eqs. 8.1-4 through 8.1-6 is directed from the fluid into the solid.
The separation of the drag force into a form force and a friction force is
that the fluid exerts on the solid, made because the former depends mainly on inertial effects and is roughly
which is especially convenient for proportional to pu~, while the latter depends on viscous effects and is roughly
Eqs. 7.3-7, 7.8-15, and 7.8-20 proportional to .u0 The word "form" results from the fact that this portion
theorem for a scalar, we get of the drag force is greatly influenced by the geometry of the solid surface,
while the word "friction" indica tes that this portion of the drag force depends
w) n dA primarily on the area of the solid surface.
The drag force for flow in closed conduits is generally represented and
(8.1-2) correlated in terms of a dimensionless friction factor f defi.ned by
pcp] + n 't'} dA
f= ___!j_ (8.1-7)
A*KE*
be kept in mind that for turbulent where
p. Splitting the area integral on A* = a characteristic area
area of entrances and exits A.(t) and KE* = a characteristic kinetic energy per unit volume
moving) A. + A.(t), we get
There are many important processes dealing with the flow of fluids through
conduits of constant cross-sectional area. For such conduits the charac-
{-n[(p- p0) + pcp] + n 't'} dA teristic kinetic energy per unit volume and characteristic area are generally
(t)
(8.1-3) given by

dA+ f D't'dA
A,+A,(t)
KE* = tp(z\) 2
A* = t wetted surface
(8.1-8)
(8.1-9)
Because A will be orthogonal to the outwardly directed unit normal at the
force which the fluid exerts on the
so lid surfaces, Fn is given by

-J n
f
] dA - A n 't' dA
A,+A,(t)
(8.1-4)
Fn = -A
A,
't' dA (8.1-10)

for conduits having a constant cross-sectiona1 area. Equation 8.1-1 O is valid


interested in a lift force, the force only for mathematically smooth conduits and thus is only an approximation
pc:nlncuJtai to A. for real conduits.
288 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.1 Frl~ion Factors-Definltion

Assuming that the stress at the wall


the numerator of Eq. 8.1-12 as

Z=O Z=L
-A .f
A,
n 'T dA= -A;

The wall shear stress 7'0 was previous

---,\ D and we may use this definition to ex)

!=
This result simply indicates that the
dimensionless wall shear stress.

The momentum balance


Fig. 8.1-1. Flow in a circular tube.
It will be helpful at this point to a
Asan example, we will apply these ideas to flow in the circular tu be shown shown in Fig. 8.1-1. A rather care
in Fig. 8.1-1. The characteristic area is given by presented here, and future discussiom
that these points are understood. F
A*= !rrDL (8.1-11)
8.1-1, the momentum balance reduc~
and substitution of Eqs. 8.1-8, 8.1-1 O, and 8.1-11 into Eq. 8.1-7 allows us to
express the friction factor as O= f{ -n[(p- Po)+ p<f>] + n '

f = __
F
D_ = _
-A J n 'T dA
__:A~!....,-_ _
(8.1-12)
.4- f
+ - n[(p
A*KE* (!rrDL)(tp(v.i)
""'
For a mathematically smooth circular tube, there is only one nonzero Here, we have assumed that the v
component of the outwardly directed normal at the so lid surface, identica1, so that the momentum fl1
Forming the scalar product with A a
n1 = nr = 1} (8.1 13a)
Eq. 8.1-18 to
n 2 = n 8 = O at the tube wall
n3 = n. =O
(8.1-13b)
(8.1-13c)
FD =A f -n[(p- Po)

and A has on1y one nonzero component.


""'
Dividing by A*KE*, we obtain th
A1 = Ar =O (8.1-14a) factor:
A2 = Ae =O (8.1-14b)
A3 = A = 1 (8.1-14c)
ic Balances-Viscous Effects Chap. 8 Sec. 8.1 Frl~lon Factors-Definition 289

Assuming that the stress at the wall is independent of () and z, we may write
the numerator of Eq. 8.1-12 as

Z=L -A .f n 't" dA= -Ai(n;T;i)TTDL= -TrzTTDL (8.1-15)


A,

The wall shear stress T0 was previously defined in Sec. 6.5 as

at r = r0 (8.1-16)

and we may use this definition to express the friction factor as

4r0
f = tp(!\)2
(8.1-17)

This result simply indicates that the friction factor may be interpreted as a
dimensionless wall shear stress.

The momentum balance


a circular tube.

It will be helpful at this point to apply the momentum balance to the flow
to flow in the circular tu be shown shown in Fig. 8.1-1. A rather careful analysis of the stress terms will be
given by presented here, and future discussions of the momentum balance will presume
(8.1-11) that these points are understood. For the control volume indicated in Fig.
8.1-1, the momentum balance reduces to
8.1-11 into Eq. 8.1-7 allows us to
O= f {-n[(p- p0) + ptf>] + D't"} dA

(8.1-12) +f-
A,
n[(p- p0) + ptf>] dA+ J n 't"dA (8.1-18)

tube, there is only one nonzero Here, we have assumed that the velocity profiles at points 1 and 2 are
at the solid surface, identical, so that the momentum flux term in Eq. 8.1-3 is identically zero.
(8.1 13a) Forming the scalar product with A and using the definition of Fn, we reduce
Eq. 8.1-18 to
(8.1-13b)
(8.1-13c)
Fn =A .J -n[(p- p0) + ptf>] dA +A .J n 't" dA (8.1-19)
A.
Dividing by A* KE*, we obtain the following expression for the friction
(8.1-14a) factor:
(8.1-14b)
(8.1-14c)
f = i(&'1 - &'2) -+- [A J n1 't" dA
Aentrance
+A J n 2 't" dA
.dexlt
J1A *KE*
(8.1-20)
Sec. 8.1 Friction Factors-Definition
290 Macroscopic Balances-Viscous Effecu Chap. 8

In obtaining this result we have used the definition and Eq. 8.1-20 may be written as

fJJ = ((p) - Po) + p(cp) D 1


(8.1-21) f = L ( &\ - !/.J - A* Kl
!-p(t\)2
and the area integral over the entrance and exit has been represented as two If the structure of the turbulence does 1
separate integrals. stress terms will cancel and we are left
The stre~s terms are best treated in terms of index notation,
(8.1-22)
where
Thus, the friction factor may also be in
drop.
Remembering that the turbulent stress was given in Sec. 6.2 as
-(t)
T;; = -pV;V;
1 1 Dimensional analysis for the frictio~

and the time-averaged viscous stress is given by


Before examining the experimental
f;; = !!:.(ov; + ov;) of dimensional analysis to determine
2 OX; OX; friction factor. Writing out Eq. 8.1-12
we may write Eq. 8.1-22 as

A (n ~) = A;n;~(~:: + ~;) - pv;v~J (8.1-23)


sr(-~~
oo
We now wish to evaluate this quantity at the entrance and exit where n has J=---
(t1TD1
only one nonzero component
Forming the dimensionless variables
n1 = nr =O } (8.1-24a)
n2 = n8 = O at the entrance and exit (8.1-24b) Z=
n3 = n. = 1 (8.1-24c)
R=
Carrying out the summations indicated in Eq. 8.1-23, and making use of
Eqs. 8.1-14 and 8.1-24, we get
~ (n ~) = (~ oz
1\
av. -~
- pv.v. 1
) (8.1-25)
=

In obtaining this result, we have assumed that the velocity field is described by Nne=
Vr = Ve = O, v. -::j=. O
allows us to express Eq. 8.1-29 as
v;, v~, v; -=1=- O
The time-averaged velocity in the z-direction will be independent of z for all
points downstream of the entrance region (i.e., in the region of one- != (~)(~)(Ll { o
dimensional flow); thus,
This result indicates immediately that
ov. =o (8.1-26) Reynolds number, but we must know
oz
Balances-Viscous Efrects Chap. 8 Sec. 8.1 Friction Factors-Definition 291

and Eq. 8.1-20 may be written as


2
(8.1-21) f = D(/JlJ
L vr-
OJl)
vr2 -
1 (< -;-;) ( -;-;))7TD
A*KE* pv.v. 1 - pv.v. 2 4 (8.1-27)

exit has been represented as two If the structure of the turbulence does not change from 1 to 2, the turbulent
stress terms will cancel and we are left with the result
of index notation,
(8.1-22) (8.1-28)

Thus, the friction factor may also be interpreted as a dimensionless pressure


drop.
given in Sec. 6.2 as

Dimensional analysis for the friction factor

Before examining the experimental values off, it will be wise to make use
of dimensional analysis to determine what parameters will infiuence the
friction factor. Writing out Eq. 8.1-12 gives

+ :;) - pv;v;J (8.1-23)


(8.1-29)
the entrance and exit where n has

(8.1-24a) Forming the dimensionless variables

entrance and exit (8.1-24b) Z=_::


D
(8.1-24c)
in Eq. 8.1-23, and making use of R=.!:.
D

-ov. - pv.v.
-,') (8.1-25) =!!L
<v.>
the velocity field is described by _ p(v.)D
NRe---
#
allows us to express Eq. 8.1-29 as
2
f= (~)(!l.)(-
1 ) JL/DJ "(- o)
will be independent of z for all
(i.e., in the region of one- R d8 dZ (8.1-30)
7T L NRe
o o
oR R=l/2

This result indica tes immediately that f depends u pon the ratio L/ D and the
(8.1-26)
Reynolds number, but we must know upon what parameters o.foR depends
292 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.2 Friction Factors-Experimental

to complete the investigation. The continuity equation and the Navier- lf the velocity profile is fully develope
Stokes equations in dimensionless form are given by be independent of Z. lntegration wit
VU=O (8.1-31) yield the factor L/ D, which just cano
entrance length for turbulent ftow is
au +u. vu = -VfJJ + - 1- V' 2U (8.1-32) number and is on the order of 50 tub
ae NRe
we may neglect the effect of the entra
The boundary conditions for this ftow may be expressed in both dimensional
f=f(NRe)
and dimensionless form as follows:

B. C. 1: V= 0, *8.2 Friction Factors: Experi


(8.1-33)
U=O, If the friction factor is measured u
values of L/ D, we soon discover that
B.C. 2: V = V1(r, (), t), Z=0 represented by a single plot off vers
(8.1-34) performed the first experimental ~tu
U = U 1(R, (), e), Z=O
we immediately wonder, "Where dtd
B.C. 3: z=L As is often the case with a mathematic
L (8.1-35) the boundary conditions. In Eq. 8.1
Z=-
D zero at r = D/2, and while this seems
true only for a mathematically smootl
B.C. 4: p = Pl - pcp, z= O
written
(8.1-36)
[JJ= fJJ, z=o
B.C. 1': V= O, r
Although we cannot solve Eqs. 8.1-31 and 8.1-32 for U and fJJ, we can state
that these two dependent variables are functions of the independent variables
(R, (), Z, e), the parameters in the differential equations (NRe), and the where the roughness function e(O, z) l
parameters in the boundary conditions (L/ D). Thus, we can write that describes the roughness of the co
magnitude of e is of the order of 0.001

U= u(R, (), Z, e, NRe ~) (8.1-37a)


"small effect," but for turbulent ftow i
do not lead to small effects." This ex~
it is a classic example of a plausib
[JJ = [JJ(R, (), Z, e, N Re ~) (8.1-37b) erroneous conclusion.
The early work by Darcy and other
experimental study of the friction facto
Ifthe time-average ftow is steady, will be independent ofthe dimensionless
tubes. The rough tubes were obtaine<
time e, and the gradient at the wall may be expressed as
possible with sand of a definite grain si

o. 1
oR R=l / 2
= function of (e, Z, NRe .!:)D
(8.1-38)
Nikuradse obtained a roughness ce
encountered in commercially availab
plotted as the friction factor versus th
In Eq. 8.1-30, the () and Z dependence will be eliminated by integration, and results are characterized by the relati
the functional dependence of the friction factor is
1. H. P. G. Darcy; for an account, see 1
(New York: Dover Publications, Inc., 1963),
(8.1-39) 2. J. Nikuradse, VDI-Forschungsh. 361, A
original paper is available as NACA Tech. M
Chap. 8
Sec. 8.2 Friction Factors-Experimental 293

and the Navier-


lf the velocity profile is fully developed at Z = O, the velocity gradient will
be independent of Z. Integration with respect to Z in Eq. 8.1-30 will then
(8.1-31) yield the factor L/ D, which just cancels the multiplying factor, D/ L. The
entrance length for turbulent flow is nearly independent of the Reynolds
(8.1-32) number and is on the order of 50 tu be diameters. For large values of L/ D,
we may neglect the effect of the entrance region and write
be expressed in both dimensional
F or large L/ D (8.1-40)

*8.2 Friction Factors: Experimental


(8.1-33)
If the fiction factor is measured using a variety of pipes all having large
values of L/ D, we soon discover that the experimental values off cannot be
z=O represented by a single plot off versus N Re as Eq. 8.1-40 implies. Darcy 1
(8.1-34) performed the first experimental studies leading to this conclusion, and
Z=O
we immediately wonder, "Where did our dimensional analysis go wrong ?"
Z= L As is often the case with a mathematical analysis, the trouble les with one of
L (8.1-35) the boundary conditions. In Eq. 8.1-33 we specified that the velocity was
Z=-
D zero at r = D/2, and while this seems like quite a reasonable statement, it is
z=O true only for a mathematically smooth tube. In actual fact, we should have
(8.1-36) written

8.1-32 for U and ~. we can state


B.C. 1': V= 0, r = lD + e(O, z) (8.2-1)
~mct1011s of the independent variables
equations (NRe), and the where the roughness function e(O, z) is sorne unknown function of O and z
D). Thus, we can write that describes the roughness of the conduit. For most commercial pipes the
magnitude of e is of the order of 0.001 of the pipe diameter. Surely this is a
"small effect," but for turbulent flow in pipes we will find that "small causes
(8.1-37a)
do not lead to small effects." This example should be remembered well, for
it is a classic example of a plausible intuitive hypothesis leading to an
(8.1-37b) erroneous conclusion.
The early work by Darcy and others led Nikuradse 2 to carry out a detailed
experimental study of the friction factor for smooth and artificially roughened
independent of the dimensionless
be expressed as tubes. The rough tubes were obtained by covering the surface as tightly as
possible with sand of a definite grain size glued onto the wall. In this manner,
Nikuradse obtained a roughness considerably more uniform than that
(o. z, NRe ~) (8.1-38) encountered in commercially available pipe. The experimental results are
plotted as the friction factor versus the Reynolds number in Fig. 8.2-1. The
be eliminated by integration, and results are characterized by the relative roughness parameter, ef D, where e
factor is
l. H. P. G. Darcy; for an account, see H. Rouse and S. Ince, History of Hydraulics
(New York: Dover Publications, Inc., 1963), p. 170.
(8.1-39) 2. J. Nikuradse, VDI-Forschungsh. 361, Ausgabe B, Band 4, 1933. A translation of the
original paper is available as NACA Tech. Mem. 1292, 1950.
Sec. 8.2 Friction Factors-Experimental
294 Macroscopic Balances-Viscous Effects Chap. 8

This equation is an indication that the


12xl0-~ c/0=9 . 86xl0~4 to affect severely the parabolic veloc'
~-8~ JJ .-=
1 o =8.34xl03
1 1~ o = 1.98xlo- 3 = 1.63xlo- 2 X l}/0 =
4 pipes (say, efD > 0.10), we could ce
9
8 \ 3 . 97xlo- 3 6 3.34xlo- 2, (commerciolly rough)_ laminar friction factor; however, this e
r. ' .
If we use Eq. 8.1-17, in conjunction wi
stress, we obtain
8J.L (v.)
4 To=--
Eq. 5.6-~ ~~ ~ D

3 1\ indicating that the drag force is propo


2.5 \ Eq. 8.2 /8 -. . loo... _... _. o;li and the velocity. Under these conditio
r'"'
2.0xlo-2 the friction force; hence,
!"" llfti!ll .... fll
~rx::
.5,~~~~--~~~~--~~~~,~~~m-.~~~Krr~~-~~
1
FD = Frriction
2 ~~~--+---~~~--+-~~~~~-.7
8*.2~-~9~-~i~~a~~~
1.0 x lo-2LW,..L---J-__-!---J-!--L--+--+-+~~~~:=;:=f::L.:~
4 6 8103 2 4 6 8 104 2 4 6 8 105 2 4 6 8 106 2 Critica! region
NflA>=u0 0/v

Fig. 8.2-1. Friction factor versus Reynolds number for sand-roughened Pipe roughness is not a factor
tu bes. number, and the transition to turb
critica! Reynolds number may depen
is the height of the sand grain. Because all the tu bes were roughened in the the pipe, upstream conditions such as
same manner-i.e., the sand grains were placed as close together as possible- spurious disturbances such as buildin
the single parameter ef D was sufficient to characterize the roughness. there is a lower bound for the criti
The student should give sorne serious thought to these results, for they however, recent studies 3 have shown t
are a clear indication that we can easily perform an apparently rigorous to Reynolds numbers of 2 X 104 by
analysis, which in fact neglects important effects. lt is quite common in flow free of disturbances. In practic
engineering analysis to treat surfaces as mathematically smooth and impose disturbance and the transition to tur
boundary conditions of the type given by boundary condition l. Generally, region between NRe = 2100 and NRe
such an approach is satisfactory (as it is for laminar flow); however, for it is here that the transition is comple
turbulent flow in tubes, a small effect such as wall roughness can lead to a and in the critica! region the flow
striking effect in the pressure drop-flow rate relationship. In the absence of turbulent regimes.
experimental studies, the average investigator is not likely to formulate the
boundary condition as indicated by boundary condition 1'; thus, we must be
constantly on the alert for flaws and limitations in a mathematical analysis. Transition and rough-pipe regions
Several aspects of this friction factor plot must be discussed before we
continue with the results for commercial pipes. For smooth tubes there is only o
the two previously mentioned-i.e.,
Laminar region flow existing for NRe > 4 X 103 H
the flow continues to change as the R1
We note first that all the data in the laminar flow region fall on a single the relative roughness is less than O.
line, the equation for which was derived in Chap. 5, smooth pipe curve for a region in w
f = 64 (8.2-2) 3. R. J. Leite, J. Fluid Mech., 1959, 5:81
N Re
ic Balances-Viscous Effects Chap. 8 Sec. 8.2 Friction Factors-Experimental 295

This equation is an indication that the values of ef D were never large enough
to affect severely the parabolic velocity profile. By using extremely rough
pipes (say, ef D > 0.10), we could certainly find an effect of ef D on the
laminar friction factor; however, this case is not of great practica! importance.
If we use Eq. 8.1-17, in conjunction with Eq. 8.2-2, to express the wall shear
stress, we obtain
8JJ. (v.) for laminar fiow (8.2-3)
To= - - -
D
indicating that the drag force is proportional to the product of the viscosity
and the velocity. Under these conditions, the drag force results entirely from
the friction force; hence,

FD = Ffriction for laminar flow (8.2-4)

2
Critica! region
=u0 0/v

number for sand-roughened Pipe roughness is not a factor in determining the critica! Reynolds
number, and the transition to turbulent flow starts at NRe = 2100. The
critica! Reynolds number may depend strongly on the inlet conditions to
all the tubes were roughened in the the pipe, upstream conditions such as valves and bends, and the presence of
placed as close together as possible- spurious disturbances such as building vibrations. lt is well established that
to characterize the roughness. there is a lower bound for the critica! Reynolds number of about 2100;
thought to these results, for they however, recent studies 3 have shown that laminar flow can be mintained up
perform an apparently rigorous to Reynolds numbers of 2 x 104 by taking extreme care to ~eep the inlet
effects. lt is quite common in flow free of disturbances. In practica! cases, there are numerous sources of
mathematically smooth and impose disturbance and the transition to turbulent flow starts at Nfi.e = 2100. The
by boundary condition l. Generally, region between NRe = 2100 and NRe = 4000 is called the critica! region and
is for laminar flow); however, for it is here that the transition is completed. The transition is ,not a sharp one,
as wall roughness can lead to a and in the critica! region the flow alternates between the laminar and
rate relationship. In the absence of turbulent regimes.
is not likely to formulate the
condition 1'; thus, we must be
ns in a mathematical analysis. Transition and rough-pipe regions
must be discussed before we
For smooth tubes there is only one more region of flow in addition to
the two previously mentioned-i.e., the region of fully developed turbulent
flow existing for NRe > 4 x 103 . However, for rough pipes the nature of
the flow continues to change as the Reynolds number is increased. Provided
laminar flow region fall on a single the relative roughness is less than 0.01, the friction factor curve follows the
in Chap. 5, smooth pipe curve for a region in which the rough pipe could be considered
64
(8.2-2) 3. R. J. Leite, J. Fluid Mech., 1959, 5:81.
296 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.2 Friction Factors-Experimental

hydraulica/ly smooth. As the Reynolds number is increased, each curve numbers, the major source of resista
eventually departs from the smooth pipe curve, progresses through a transi- turbulent eddies with the wall rough
tion region, and finally reaches a constant value depending only on the relative compared to the form drag. In the 11
roughness. This latter region will be called the rough-pipe region offlow, and factor is constant, and Eq. 8.1-17 i1
the zone between the smooth-pipe and rough-pipe regions will be called the proportional to the density times the
transition region.
T0 = (~) p(v.) ~n
2

Low Reynolds number Comparing Eqs. 8.2-3 and 8.2-5, we s


momentum transport exist in pipe flo
the other of inertial effects. Also, it i.
the formula developed in Chap. 7 fo1
which indicated that
force exert
{on the pla

where v and A are the velocity and


parison,. we might interpret Eq. 8.2-
force per unit
exerted on the
{
protrusions

There is obviously sorne similarity be

Commercial pipes

Pipes and tubes used in engineer


smooth nor be considered rough in
Flg. 8.2-2. Qualitative desription of wall effects in roughened pipes. The roughness in <
turbulent ftow. it subject to direct measurement. H1
to any given pipe if the friction fac
We can explain the existence of the transition and rough-pipe regions only Moody 4 has made an extensive inve~
in a qualitative manner; however, the explanation will help to explain the cial pipes, the results of which appe<
nature of turbulent pipe flow and should be valuable for this reason alone. widely used in engineering design
At low Reynolds numbers the laminar sublayer is generally thicker than the Values of the relative roughness are
average roughness e, and the wall shear stress To consists primarily of viscous pipe.
stresses. As the Reyno!ds number is increased, the energy of the turbulent When using these charts, an engi
eddies increases, and they penetrate more closely to the wall. This process of e/ D are only approximate and th
causes the thickness of the laminar sublayer to decrease, and the irregularities relative roughness for any given typ<
of the tube wall begin to protrude through the laminar sublayer, as Fig. probable variation injfor smooth t1
8.2-2 illustrates. When the turbulent eddies come into contact with these 10 per cent is to be expected for co
protrusions, the interaction of the fluid and solid changes. At high Reynolds 4. L. F. Moody, "Friction Factors for
Chap. 8 Sec. 8.2 Friction Factors-Experimental 297

number is increased, each curve numbers, the major source of resistance to flow is the interaction of the
curve, progresses through a transi- turbulent eddies with the wall roughness and friction drag becomes small
value depending only on the relative compared to the form drag. In the rough-pipe region of flow, the friction
the rough-pipe region offlow, and factor is constant, and Eq. 8.1-17 indica tes that the wall shear stress is
regions will be called the proportional to the density times the velocity squared.

_ (/_) ( )2 In the rough-pipe region


7 o- P v. of turbulent flow (8.2-5)
8
Comparing Eqs. 8.2-3 and 8.2-5, we see that two very distinct mechanisms of
momentum transport exist in pipe flow-one a result of viscous effects and
the other of inertial effects. Also, it is of interest to compare Eq. 8.2-5 with
the formula developed in Chap. 7 for the force exerted by a jet on a plate,
which indicated that
force exerted} = v2 A (8.2-6)
{on the plate P

where v and A are the velocity and area of the jet, respectively. For com-
parison, we might interpret Eq. 8.2-5 as

force per unit area}


exerted .on the = -
(f) p(v.)
2
(8.2-7)
{protrus10ns 8
There is obviously sorne similarity between the two flows.

Commercial pipes

Pipes and tubes used in engineering practice can neither be regarded as


smooth nor be considered rough in the same sense as Nikuradse's sand-
in roughened pipes. The roughness in commercial pipes is not uniform, nor is
it subject to direct measurement. However, a value of ef D can be assigned
to any given pipe if the friction factor is known in the rough-pipe region.
and rough-pipe regions only Moody 4 has made an extensive investigation of friction factors for commer-
l'lJJiamtuuu will help to explain the cial pipes, the results of which appear in Fig. 8.2-3. This particular chart is
be valuable for this reason alone. widely used in engineering design and is often called the Moody chart.
is generally thicker than the Values of the relative roughness are given in Fig. 8.2-4 for various types of
-r0 consists primarily of viscous pipe.
the energy of the turbulent When using these charts, an engineer must keep in mind that the values
closely to the wall. This process of ef D are only approximate and there may be significant variations in the
to decrease, and the irregularities relative roughness for ny given type of pipe. Moody has indicated that the
the laminar sublayer, as Fig. probable variation inJfor smooth tubing is 5 per cent, anda variation of
come into contact with these 1Oper cent is to be expected for commercial steel pipe. Corrosion may also
solid changes. At high Reynolds 4. L. F. Moody, "Friction Factors for Pipe Flow," Trans. ASME, 1944, 66:671.
298 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.2 Friction Factors-Experimental
Pipe diamete
04 06
0 .2 0.3 0.5 0.81

1\

Riveted
steel
1"'-111.
"
11 ro.
1/ \,) N
" Concrete

~ Woo d~
l"i 1'\stave

~
-;;
;::.
:
~ ~
8
o
a:
<: ...o
....
Q)
.S
.o
E "'...o
::>
e t;
en ~
'O
oe =
o
>.
Q)
a:: ]
,.;
,:.
..
:

Pipe diomet1
Fig. 8.2-4. Relative roughr

cause large variations in ef D with tir


analyzing pipelines which have been in
time.
J JO:lO~ UO!:lDJ.:::I There are two useful empirical form
5. C. F. Colebrook and C. M. White,""
Pipes with Age," J. /nst. Civ. Engrs. (London)
Balances-Viscous Effects Chap. 8 Sec. 8.2 Friction Factors-Experimental 299
Pipe dio meter in feet, O
0.4 0.6 25
0.1 0 .2 0.3 0.5 0.81 2 3456810 20
5
o.o 1agga!ffi~lnm~nmtEf1Eitlno.o7
0.04k:
0.03 ro. 0 .06
0 .02 [\ 0 .05

40 60 100 300
Pipe diomeler in inches, O
Fig. 8.2-4. Relative roughness for commercial pipes.

cause large variations in ef D with time,5 and we must be cautious when


analyzing pipelines which have been in operation for an extended period of
time.
There are two useful empirical formulas giving the friction factor in terms
5. C. F. Colebrook and C. M. White, "The Reduction of the Carrying Capacity of
Pipes with Age," J. Inst. Civ. Engrs. (London), 1937, 7:99.
300 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.2 Friction Factors-Experimental

of the Reynolds number for smooth tu bes. For Reynolds numbers less than Reynolds number can then be determi1
105 , Blasius6 gives the following equation assumption for f may be examined. If
second trial must be made. In the follo
f = 0.316Nfi~ 4 For smooth tubes (8.2-8) type of calculation.
This equation is plotted in Fig. 8.2-1 and is in excellent agreement with
Nikuradse's data for smooth tu bes provided N Re :::;:; 105. For higher Reynolds Determination of the flow rate for a
numbers, the Prandtl equation7 is useful: pressure drop

JJ1 = 2.0 log (Jj N Re) - 0.8 For smooth tubes (8.2-9) Let us determine the volumetric fic
750-ft length of 4-in. diameter (nominal
drop of 23.5lbcfin. 2 We shall assume d
This equation is also plotted in Fig. 8.2-1, and shows good agreement with tional effects need not be considered. 1
the experimental data. The derivation of Eq. 8.2-9 is semitheoretical and
should give satisfactory results for arbitrarily high Reynolds numbers. The D = 4.03 in. (a
Prandtl equation requires a trial-and-error solution to determine J; however, L= 750ft
if the Blasius equation is used to obtain the first estmate, the final value p.= 0.95 centi
can generally be determined with only one additional calculation.
A useful equation for determining the friction factor in the transition and p = 62.4 lbm/f
rough-pipe regions has been developed by Colebrook. 8 t:J.p = 23.5 lbr/in
.E_= 0.0004
JJ = -2.0 log (e/DD + NRe,J]
1 2.51 ) For transition and
rough-pipe regions (8.2-10)
D
We may rearrange Eq. 8.1-28 to yield
This emprica! relationship is, in fact, the basis for the curves shown in
Fig. 8.2-3. lt requires a trial-and-error solution for f; however, such pro-
cedures are straightforward and this equation may prove useful if a digital
computer is used for pipe-sizing calculations. and the Reynolds number becomes
Problems dealing with turbulent flow in pipes generally fall into two _ p(v,)D _
categories: determination of the pressure drop, given the flow rate, the N Re- -
p.
physical properties of the fluid, and the geometry of the system (i.e.,
For this example, we obtain
the length, diameter and relative roughness of the pipe); or determination
2
of the flow rate, given the pressure drop, the physical properties of the fluid, 1 { [(2X23.5 lbr/in. )
N --
and the geometry of the system. The first case is straightforward because Re - ,Jj (0.95 cen
the Reynolds number can be calculated and the friction factor deter-
mined immediately. The second type of problem requires that an initial guess x [ ( centipoise-ft se1
be made for f, allowing us to calculate the average velocity by Eq. 8.1-28. The 0.672 X 10-3 ]1>
or
6. H. Blasius, "The Law of Similarity for Frictional Processes in Fluids," Forsch. 4.
Arb. lngr.-Wesen (Berlin, 1913), 131:361. NRe=-
7. L. Prandtl, "The Mechanics of Viscous Fluids," in Aerodynamic Theory, F. W.
Durand, ed. (Berlin: Springer-Verlag, 1935), 3:143.
8. C. F. Colebrook, "Turbulent Flow in Pipes, with Particular Reference to the Transi- Now we need only find the value ofjJ
tion Region between the Smooth and Rough-Pipe Laws," J. lnst. Civ. Engrs. (London), equation and the relationship between
1938, 11:133. The mnimum value off in this case
Balances-Viscous Effects Chap. 8 Sec. 8.2 Frlctlon Factors-Experimental 301
For Reynolds numbers less than Reynolds number can then be determined, and the accuracy of the initial
assumption for f may be examined. If the two values differ significantly, a
second tria! must be made. In the following example, we consider this type
(8.2-8) type of calculation.
and is in excellent agreement with
N Re ::;; 105. For higher Reynolds Determination of the flow rate for a given
pressure drop

For smooth tubes (8.2-9) Let us determine the volumetric flow rate of water at 75F through a
750-ft length of 4-in. diameter (nominal) commercial steel pipe for a pressure
2-1, and shows good agreement with drop of 23.5lbr/in. 2 We shall assume the pipe is horizontal so that gravita-
tional effects need not be considered. lt is given that
of Eq. 8.2-9 is semitheoretical and
"ly high Reynolds numbers. The D = 4.03 in. (actual diameter)
solution to determine f; however,
L= 750ft
n the first estmate, the final value
one additional calculation. p = 0.95 centipoise
friction factor in the transition and p = 62.4 lbm/ft 3
by Colebrook.8 l:lp = 23.5 lbrfin. 2
2.51 ) For transition and !_ = 0.0004
NReJJ rough-pipe regions (8.2-10) D
We may rearrange Eq. 8.1-28 to yield
the basis for the curves shown in
1 (21:1p D)
112
solution for f; however, such pro- (vz) = Jj ---;L
may prove useful if a digital
and the Reynolds number becomes
in pipes generally fall into two
drop, given the flow rate, the N - p(v.) D - _!._ (21:lpp Da)I/2
the geometry of the system (i.e.,
Re- p -Ji p2L
of the pipe); or determination For this example, we obtain
the physical properties of the fluid, 3
N Re=~ { [(2X23.5 lbr/in.~62.4lbm/ft X4.03 in.)
3
]
rst case is straightforward because
and the friction factor deter- -J (0.95 centipoise) 2(750 ft)
problem requires that an initial guess
averagevelocity by Eq. 8.1-28. The
2
x [ ( centipoise-ft sec ) (32.2 lbmft) (__.!!__) J}1 2
'
0.672 X 10-3 lbm lbr sec2 12 in.
or
Frictional Processes in Fluids," Forsch.
_ 4.13 X 104
Fluids," in Aerodynamic Theory, F. W.
N Re- Ji
43.
with Particular Reference to the Transi- Now we need only find the value ofjfor ejD = 0.0004 that will satisfy this
Laws," J. Jnst. Civ. Engrs. (London), equation and the relationship between f and N Re given in the Moody chart.
The mnimum value off in this case is 0.0 16, which yields a value for the
302 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.2 Friction Factors-Experimental

Reynolds number of pipe sizes is available in handbooks, 9 -


of facing page) will be cited here for a
N - 4.13 X 104 - 3 26 105
Re - .Jo.016 - . x The variation in inside diameter is s
designing any real system. Throughout
Examination of the friction factor chart for s/ D = 0.0004 and N Re = dimensions of any system are assumed t~
3.26 x 105 gives
!= 0.0178 Flow in closed conduits of noncircul
which in turn yields the second approximation for the Reynolds number,
Closed conduits of noncircular
NRe =
4.13 X 104 = 3.08 X 105 although they have received much les
.Jo.0178 laminar flow, each noncircular conduit
Returning to the chart, we see that a Reynolds number of 3.08 x 105 gives problem, for the result from one case e
a friction factor of 0.0178 for this particular relative roughness. Having and Katz11 list severa! solutions of t
established the Reynolds number, we may now determine the volumetric dimensional laminar flow in noncircula
flow rate, Q. Bateman12 give a more thorough discu
severa! solutions. Since an engineer i.
(7T~ )(v.) = (NR~:D/l)
2

Q= pressure drop-flow rate relationship, t


suggested by Gaydon and NuttaP 3 is of
= [(3.08 x 105)(3.14)(4.03 in.)(0.95 centipoise)J odd-shaped conduits. The method re
(4)(62.4Ihm/fe) effort to determine the pressure drop-f
3
0.672 X 10- lbm ) ( ft )] per cent.
X [( centipoise-ft sec 12 in. Turbulent flow has been studied by
rectangular and circular notched condUJ
= 0.83 fejsec
in Fig. 8.2-5. For each conduit, the fr
Nominal pipe diameters J=-
At
It is unfortunate but true that the nominal pipe diameter is not the actual where A* and .KE* are given by Eqs
pipe diameter; we must take this into account when specifying pipe diameters. number is given by
In the previous example, the actual diameter of a 4-in. commercial steel pipe 4
was 4.03 in. The difference between the nominal and actual diameters NRe=-
depends upon the diameter and the so-called schedule number of the pipe.
The latter is simply a measure of the wall thickness and is therefore an indi- 9. Chemical Engineers' Handbook, 4th ed.
cation of the pressure that the pipe can safely withstand. Information on lnc., 1963), Sec. 6.
10. L. S. Marks, ed. Mechanical Engineers
Nominal Outside Schedule Inside Book Company, Inc., 1958).
Pipe Size, in. Diameter, in. No. Diameter, in. 11. J. G. Knudsen and D. L. Katz, Flu1
McGraw-Hill Book Company, Inc., 1958), Cl
6 6.62 SS 6.41 12. H. L. Dryden, F. D. Murnaghan, a
lOS 6.36 Dover Publications, Inc., 1956), Chap. 2.
40ST,40S 6.06 13. F. A. Gaydon and H. Nuttal, "Viscou
soxs, sos 5.76 Cross Sections," Trans. ASME Ser. E 81, 19
120 5.50 14. L. Schiller, Z. Angew. Math. Mech., 1
160 5.19 15. J. Nikuradse, Ingr.-Arch., 1930, 1:3
XX 4.90 16. L. Prandtl, Proc. lntern. Congr. Appl.
Sec. 8.2 Friction Factors-Experimental 303

pipe sizes is available in handbooks, 9 - 10 and only a single example (see foot
of facing page) will be cited here for a 6-in. diameter steel pipe.
The variation in inside diameter is significant and must be considered in
designing any real system. Throughout the remainder of this text the nominal
ef D = 0.0004 and N Re = dimensions of any system are assumed to be identical to the actual dimensions.

Flow in closed conduits of noncircular cross section


v,. .... ~.UV'U for the Reynolds number,
4 Closed conduits of noncircular cross section are occasionally used,
l0 = 3.08 X 105 although they have received much Iess attention than circular tubes. For
laminar flow, each noncircular conduit must be Iooked upon as a separate
Reynolds number of 3.08 x 105 gives problem, for the result from one case cannot be extended to others. Knudsen
particular relative roughness. Having and Katz11 Iist severa! solutions of the Navier-Stokes equations for one-
may now determine the volumetric dimensional laminar flow in noncircular conduits. Dryden, Murnaghan, and
Bateman12 give a more thorough discussion of these problems and present
severa! solutions. Since an engineer is generally concerned with only the
pressure drop-flow rate relationship, the approximate method of solution
suggested by Gaydon and NuttaP 3 is of great value in obtaining solutions for
odd-shaped conduits. The method requires only a small computational
effort to determine the pressure drop-flow rate relationship to within a few
per cent.
Turbulent flow has been studied by severa! investigators14- 16 for triangular,
rectangular and circular notched conduits; the experimental results are shown
in Fig. 8.2-5. For each conduit, the friction factor is defined by
FD (8.2-11)
f= A*KE*
nominal pipe diameter is not the actual
where A* and KE* are given by Eqs. 8.1-8 and 8.1-9, and the Reynolds
l""~v .. cu when specifying pipe diameters.
,.,uu....,... of a 4-in. commercial steel pipe number is given by
nominal and actual diameters _ 4p(v,)Rh
N Re- (8.2-12)
so-called schedule number of the pipe. ;,
wall thickness and is therefore an indi- 9. Chemical Engineers' Handbook, 4th ed. (New York: McGraw-Hill Book Company,
can safely withstand. Information on lnc., 1963), Sec. 6.
10. L. S. Marks, ed. Mechanical Engineers Handbook, 6th ed. (New York: McGraw-Hill
Schedule lnside
Book Company, Inc., 1958).
No. Diameter, in.
11. J. G. Knudsen and D. L. Katz, Fluid Dynamics and Heat Transfer (New York:
McGraw-Hill Book Company, lnc., 1958), Chap. 4.
5S 6.41
12. H. L. Dryden, F. D. Murnaghan, and H. Bateman, Hydrodynamics (New York:
lOS 6.36
Dover Publications, lnc., 1956), Chap. 2.
40ST,40S 6.06
13. F. A. Gaydon and H. Nuttal, "Viscous Flow Through Tubes of Multiply Connected
80XS, 80S 5.76
Cross Sections," Trans. ASME Ser. E 81, 1959, 4:573.
120 5.50
14. L. Schiller, Z. Angew. Math. Mech., 1923, 3:2.
160 5.19
15. J. Nikuradse, Ingr.-Arch., 1930, 1 :306.
XX 4.90
16. L. Prandtl, Proc. Jntern. Congr. Appl. Mech., Zrich, 1927.
10.0
304 Macroscopic Balances-Viscous Effects Chap. 8 8.0
6.0
4.0

"
r J 6-38
Eq . 5.
0. 1 '
"k~ 2.0 t'-....: ~
~"""-
_)--- Eq 8. 1-8 ~...
O. 1
0. 08
"" " "'\ ~ 'So
o
~

,_'
~
~o

O. 06
["-. ./' Co 1.0
0.8
1--:-:::-.,_ equ 11 atera -o-_ ~-1f
~ J \
o. 6
lb.. r-< r--
kC!i~ ~~!~
'<
1"" o. 4
O. 01

O. 01
""" ~"~''''"''lJt~
..... 1 oy -o-:. :!!_tched
~!r
1 r....
le._1
1"--
o. 2
O. 008
--.,__In. ~ o.1
"'
! f 2 4 6 81Q2 2 4 6 11103 2
10
~ r-<
O. 006 1--
o N1kuradse . Re f. 15
O. 004
Schl11er , Ref . 14 Fig. 8.2-6. Drag coefficien

O. 001
1 1 1 1
1 The characteristic area A* for imme
projected area in the direction of fio
Flg. 8.2-5. Friction factors for noncircular conduits.
A*- TTD2
- 4 '

where Rh is the hydraulic radius. The results illustrated in Fig. 8.2-5 indicate A*= DL,
excellent agreement with the Blasius equation; in addition, the values in the
The characteristic kinetic energy is ta
laminar fl.ow region fall on the line given by
velocity far removed from the imme
and 8.2-15, the drag force acting on
(8.2-13)
FD=eD(
These results must not be construed as verification of the friction factor chart
for all shapes of closed conduits, for it is well known that laminar ftows may and the drag force acting on a cylindl
deviate significantly from Eq. 8.2-13. However, we may conclude that if the
shape of the conduit is not far removed from circular the friction factor chart FD= eD
can be used to obtain :.atisfactory results.
Experimental values of eD for spher
Dimensional analysis would again 1
Drag coefficients for spheres and cylinders function of Reynolds number and reJa
has not been studied extensively bec
with reducing the drag on immersed 1
The friction factor for solid bodies immersed in a flowing fluid is tradi- of accomplishing this end.t
tionally referred to as a dimensionless drag coefficient and defined by
t This is not always the case, for somet
drag. This phenomenon, along with the c
e - __fJ_ (8.2-14) detail in Chap. 11.
D- A*KE*
10.0
Chap. 8 8.0
6.0
4.0

"' ~

Co
2.0

1.0
' ""' ~" .... Circular cylinder

"
0 .8
0.6 Sphere

0.4 t--- _....,

0 .2 \.11

o.1
10 2 4 6 81Q2 2 4 6 8Q3 2 4 6 8 104 2 4 6 8 105 2 4 6 8Q6
NRe
Fig. 8.2-6. Drag coefficients for spheres and cylindl:rs.

6 8
The characteristic area A* for immersed bodies is generally taken as the
projected area in the direction of ftow; thus,
for noncircular conduits.
2
A* = 7T D for a sphere (8.2-15a)
4 '
results illustrated in Fig. 8.2-5 indicate A* = DL, for a cylinder (8.2-15b)
; in addition, the values in the
The characteristic kinetic energy is taken to be !pu;,, where U 00 is the fluid
velocity far removed from the immersed body. On the basis of Eqs. 8.2-14
64 and 8.2-15, the drag force acting on a sphere is
(8.2-13)
(8.2-16)
verification of the friction factor chart
is well known that laminar ftows may
and the drag force acting on a cylinder of length L and diameter D is
However, we may conclude that if the
from circular the friction factor chart (8.2-17)
Experimental values of eD for spheres and cylinders are shown in Fig. 8.2-6.
Dimensional analysis would again lead us to the conclusion that en is a
function of Reynolds number and relative roughness. The effect of roughness
has not been studied extensively beca use engineers are generally concerned
immersed in a flowing fluid is tradi- with reducing the drag on immersed bodies; keeping them smooth is a means
drag coefficient and defined by of accomplishing this end.t
t This is not always the case, for sometimes increasing the roughness can decrease the
drag. This phenomenon, along with the curves shown in Fig. 8.2-6, will be discussed in
(8.2-14) detail in Chap. 11.

305
306 Macroscopic Balances-Viscous Effects Chap. 8
Sec. 8.3 Pipeline Systems

For Reynolds numbers less than 1.0, the inertial terms in the Navier-
Stokes equation can be neglected under certain circumstances, which leads to a
linear set of equations first solved by StokesY The drag coefficient for this
condition is
en- - 24
and the drag force acting on the sphere is given by
FD =
NRe

31TflU 00 D
(8.2-18)

(8.2-19)
T
2 00'

Equation 8.2-19 is known as Stokes law and has been used extensively by
engineers to describe the motion of solid spheres moving through gases and
Iiquids.
The discussion here has been quite brief; however, a detailed treatment of
flow around immersed bodies is given in Chap. 11 . l_
2'
8.3 Pipeline Systems

One ofthe standard problems that an engineer may encounter is the design
T 20'--

of a piping system such as the one illustrated in Fig. 8.3-1. In the design of a Pump
new system, the flow rate and the physical properties of the fluid are generally flg. 8.3-1. 1
given, and the engineer must determine the pipe size and the power require-
ment for the pump. The final solution will naturally be subject to certain
economic constraints beyond the scope of this text. The main economic If we now take the positive z-coord
problem hinges on the fact that the cost of pumps and pipes depend on the gravity vector, the potential energy fl
size and that the final design of the system should satisfy the physical con-
ditions at a minimum cost.
The mechanical energy balance is the most suitable macroscopic balance
w use in solving pipeline problems, because the troublesome viscous effects
In addition, we will consider control
are nicely lumped into a positive dissipation term. If we restrict ourselves to
entrances and exits is normal to the e
steady flow, fixed control surfaces at entrances and exits, and incompressible
fluids, the mechanical energy balance (Eq. 7.3-27) takes the form
Net outftow of
kinetic energy f V t<al dA=

f 2
(lpv )v n dA= f
A.
V t(n) dA
Rate of work done on the
system at the entrances
and exits
A.

Substituting Eqs. 8.3-2 and 8.3-3 intt


-f
A.
prf>vndA
Rate of work done by
gravitational forces,
or the net outftow of
potential energy
(8.3-1)
volume with one en trance (designa te<
we obtain
Rate of work done on the
system by moving solid
surfaces
Rate of viscous
-, dissipation to
interna! energy = -p
17. G. G. Stokes; for an account, see H. Lamb, Hydrodynamics, 6th ed. (New York:
Dover Publications, Inc., 1945), pp. 597-604. If the velocity profiles are nearly fla
Chap. 8
Sec. 8.3 Pipeline Systems 307
1.0, the inertial terms in the Navier-
certain circumstances, which leads to a
StokesP The drag coefficient for this

(8.2-18)

(8.2-19)
T200'

law and has been used extensively by


spheres moving through gases and

brief; however, a detailed treatment of


in Chap. 11. j_
2'

20~
engineer may encounter is the design
T
in Fig. 8.3-1. In the design of a Pump
properties of the fluid are generally Flg. 8.3-1. Pipeline system.
the pipe size and the power require-
will naturally be subject to certain
of this text. The main economic If we now take the positive z-coordinate to be oppositely directed to the
of pumps and pipes depend on the gravity vector, the potential energy function, cp, takes the form
should satisfy the physical con-
cp = gz . (8.3-2)
most suitable macroscopic balance
the troublesome viscous effects
term. If we restrict ourselves to In addition, we will consider control volumes such that the velocity at the
entrances and exits is normal to the control surface; thus,
trances and exits, and incompressible
7.3-27) takes the form

Rate of work done on the


J V t(n) dA = - J pv 'n dA (8.3-3)

systetn at the entrances


and exits

Substituting Eqs. 8.3-2 and 8.3-3 into Eq. 8.3-1, and considering a control
Rate of work done by
dA gravitational forces, volume with one entrance (designated by 1) and one exit (designated by 2),
or the net outftow of (8.3-1)
potential energy we obtain
Rate of work done on the
system by moving solid
surfaces
Rate of viscous
lp(v3 )zAz - lp(v 3 )A + pgz (v)zAz -
2 pgz1 (v)A
dissipation to (8.3-4)
internal energy
= -pz(v)zA 2 + p 1 (v) 1A1 + W- E"
Hydrodynamics, 6th ed. (New York:
If the velocity profiles are nearly flat, we may use the macroscopic mass
308 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.3 Pipeline Systems

balance a valve, a pump, etc. Writing Eq. 8.3


(8.3-5)
(v? + ]!_ + z} - {(v)2+
to reduce Eq. 8.3-4 to {2g pg 2 2g

(v): + P2 + z2} {_!_ (v)~ + Pt + Z} = W - E, (v)2 + ]!_ + z} - {(v)2 +


{_!_
2g . pg
-
2g pg pgQ pgQ
(8.3-6) { 2g pg 3 2g

Each term in Eq. 8.3-6 has the units of length, or may be converted to units ----------------
of length by multiplying g., if English units are used. Thus, the tcrm
(v)2 + .E_ + z} - {(v)2 + 1?.
{ 2g pg N-1 2g P8
}!_or _P_
pg pgfg.
(v)2 + ]!_ + z} - {(v)2 + 1?.
is often called the head or pressure head, the designation resulting from the { 2g pg N 2g p
fact that any pressure can be related to a column of static Iiquid by the Summing all these equations, we see
equation, remain on the left-hand side, while th
}!_ = h, for a static fluid the head loss and HID terms.
pg
(v)2 + ]!_ + z} - {(v)2 + ]!_ .
{ 2g pg N 2g pg
and the fact that the height of a static liquid above a datum plane is tradi-
tionally referred to as the liquid head. The combination of terms, pf pg + z, Using !i to indicate differences bet~
is called the piezometric head, and (v) 2/2g is referred to as the velocity head. Changc in
Changc in
In keeping with this traditional nomenclature, the dissipation term E, vclocity pressurc
hcad
Changc in
elcvation
hcad
divided by pgQ is called the head /oss, and is generally separated into two
!1{~~2} +
!ip
terms, h1 and h,.. We will use h1 to refer to frictional head losses in straight
sections of conduits having a uniform cross section, and hm to refer to head pg
+ !iz =: 1

losses occurring at sudden expansions, gate valves, etc. These losses are
often called minor losses beca use the major source of viscous dissipation is in
the comparatively long sections of straight conduit. The change in head
w pgQ owing to the power supplied by a pump or the power delivered to a
turbine, is designated as Htl), the subscript w indicating that this term refers At this point, the student should begi
to the rate of work done on or by the fluid. The term H will be positive for
ti) the derivation of the Navier-Stokes
pumps that supply energy to the fluid and negative for turbines that extract energy equation, and the macrosco
energy from the fluid. Using this nomenclature and simplifying the subscripts each was an important, logical step
for the entrance and exit, we may rewrite Eq. 8.3-6 in the form the complexity of flow through a pip
analyze such a system with the ease

{(v,)' +.E_+ z} - {(v) + J!.. + z} = Htl)- h


2 result. With the aid of experimental
1
- hm (8.3-7)
2g pg 2g
2 pg 1
coefficients we can use Eq. 8.3-9 to de:
system.
Before discussing the head loss terms in detail, it will be helpful to indicate To determine the friction head 1<
the method of application of Eq. 8.3-7. Let us consider a pipeline to be split scopic momentum and mechanical
up into N - 1 separate sections, each of which contains either a straight shown in Fig. 8.1-1. The momentu11
section of conduit or one of the numerous special elements such as an elbow,
Sec. 8.3 Pipeline Systems 309

a valve, a pump, etc. Writing Eq. 8.3-7 for each section of the system gives
(8.3-5)
- +-p + Z }
(v)2
{ 2g -
{(v)2
- + -p + Z } = {Hw- h,- hm}1
pg 2 2g pg 1

W -
+ -P1pg + Z} = -pgQ E (8.3-6)
-" - +-p + Z }
(v)2 -
{(v)2
- + -p + Z } = {Hw- h,- hm} 2
pgQ { 2g pg 3 2g pg 2

of length, or may be converted to units


units are used. Thus, the tcrm
2
(v)
{ 2g
+ 1!... + z} - {(vJ2 + 1!... + z} = {Hw- h,- hm}N- 2
pg N-1 2g pg N-2

the designation resulting from the


- + -p + Z }
(v)2
{ 2g -
{(v)2
- + -p + Z } = { H w - h, - hm } N-1
pg N 2g pg N-1
to a column of static liquid by the
Summing all these equations, we see that only the first and the Nth terms
remain on the left-hand side, while the right-hand side consists of the sum of
a static ftuid the head loss and Hw terms.
2 2

{(v) + 1!... + z} _ {(v) + 1!... + z} =! Hw-! (h 1 + hm) (8.3-8)


liquid above a datum plane is tradi- 2g pg N 2g pg 1 N N

The combination of terms, p/ pg + z, Using d to indicate differences between the outlet and the inlet, we obtain
is referred to as the velocity head.
Change in Change in
the dissipation term E. velocity pressure Change in
head head elcvation
and is generally separated into two
d{~~2} +
to frictional head losses in straight dp Sum o head losaea

cross section, and hm to refer to head pg


+ dz = !Hw
N
or gains or pumpa
and turbines

gate valves, etc. These losses are - !h,


Sum o losaea or
allstraighl (8.3-9)
source of viscous dissipation is in N sectiona

conduit. The change in head - !hm


Sum oC!osaea for
all valuea,
a pump or the power delivered to a N ftuings, etc.

w indicating that this term refers At this point, the student should begin to appreciate all the effort involved in
The term Hw will be positive for the derivation of the Navier-Stokes equations, the differential mechanical
and negative for turbines that extract energy equation, and the macroscopic mechanical energy balance, because
~nc:Iature and simplifying the subscripts
each was an important, logical step in arriving at Eq. 8.3-9. If we consider
Eq. 8.3-6 in the forro the complexity of flow through a pipeline, it is an achievement to be able to
analyze such a system with the ease and simplicity suggested by the above
J!... + z} = Hw- h1 - hm (8.3-7)
result. With the id of experimentally determined friction factors and loss
pg 1 coefficients we can use Eq. 8.3-9 to design a new pipeline or analyze an existing
system.
in detail, it will be helpful to indicate To determine the friction head loss, h1, we must apply both the macro-
Let us consider a pipeline to be split scopic momentum and mechanical energy balances to the control volume
of which contains either a straight shown in Fig. 8.1-1. The momentum balance is given by Eq. 8.1-28, which
special elements such asan elbow,
310 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.3 Pipeline Systems

we may rearrange to get condense the expression for hm to t

{_E_+
pg
z} -
1
{L + z}
pg 2
=!(L) (v)2
D 2g
(8.3-10)

where where K is a function of the Reynold


(v) = (v.)1 = (v.)2 that enter into the boundary conditi
Substitution of this result into Eq. 8.3-7, Hw and hm being zero, gives
The sudden expansion
h1 =!(L) (v)2 for circular tubes (8.3-11)
D 2g' The sudden expansion in a pipe!
This result is called the Darcy-Weisbach 18
formula. For noncircular con- which we can compute the head loss
duits, we must remember that the term L/ D resulted from momentum balance predicted the f

i wetted surface !7T DL L


cross-sectional area = - -2 = -
7TD D
4
Making use of the definition of the hydraulic radius Rh> we have a more
o,
general form of Eq. 8.3-11,
_ (_.!:._) (v) 2
for cir~ular and noncircular ( _
83 12)
i
h, - f ' condUits.
4R~~, 2g
Equation 8.3-12 and the friction factor-Reynolds number chart provide us Fig. 8.3-2. S
with sufficient information to determine the friction head losses in a pipeline.
The minor losses must also be determined experimentally and tabulated thus, we may make use of both
for all the different types of fittings. The equation for hm for a Newtonian balance to determine hm. The dime
fluid is

h = , =
f
"Y
___
ci> dV
=
2-t f Vv: d dV
_.:...."~'____ (8.3-13) and
~p = 2 [ ( ~J - J(
1

m pgQ pg(v)A

Using R~~, as a characteristic length and (v) as the characteristic velocity, we


pg( v)A
~p = [(Dl)4
D
2
- t] + ~
7TD 1 p(
may put the volume integral in dimensionless form to obtaint Equating the two results gives an e~
2
hm = (v) {.....::!__
2g p(v)A
f
"~'
VU: D dV) (8.3-14)
h
m
= 4Ev
7T D~pg( v.)
-tR~~, and the predicted loss coefficient, K
Since p (v )A/-tR~~, is directly proportional to the Reynolds number, we may
t Here the volume dV is actually dimensionless so that the term enclosed by braces { }
is dimensionless.
18. H. Darcy, "Experimental Researches on the Flow of Water in Pipes," Compt. The result is in error by about 20 I
Rend., 1854, 38:1109. Reynolds numbers.
Chap. 8 Sec. 8.3 Pipeline Systems
311

condense the expression for hm to the traditional form,


+ z} =!(.!:_) (v)2 (8.3-10) h =K (v)2
2 D 2g m (8.3-15)
2g
where K is a function of the Reynolds number and any dimensionless groups
that enter into the boundary conditions.
, Hw and hm being zero, gives
The sudden expansion
for circular tubes (8.3-11)
The sudden expansion in a pipeline, shown in Fig. S.3-2, is one case for
18
formula. For noncircular con- which we can compute the head loss hm directly. In Sec. 7.5 we found that the
Lf D resulted from momentum balance predicted the pressure rise with reasonable accuracy;
i1rDL L
=--=-

radius R,., we have a more


~t~~'
01
?b==---..: 1
Dz

for circular and noncircular


conduits.

Reynolds number chart provide us


(8.3-12

the friction head losses in a pipeline.


)

!6==---= l
Fig. 8.3-2. Sudden expansion.
~rnunt:o experimentally and tabulated
equation for hm for a Newtonian thus, we may make use of both the momentum and mechanical energy
balance to determine hm. The dimensionless pressure difference is given by

f
2,u Vv:d dV
"Y (8.3-13)
Momentum balance
(Eq. 7.5-13)
and
pg(v)A
Mechanical energy balance
' (v) as the characteristic velocity, we (Eq. 7.5-31)
form to obtaint
Equating the two results gives an expression for Evf pgQ,

(8.3-14) h
m
=
7T
4Ev
Dipg(v.)
= [1 _ (D1)
D2
2
]
2
Mi
2g
(8.3-16)

and the predicted loss coefficient, K, is given as


to the Reynolds number, we may
so that the term enclosed by braces { } (8.3-17)
on the Flow of Water in Pipes," Compt. The result is in error by about 20 per cent and may be used only for large
Reynolds numbers.
Sec. 8.3 Pipeline Systems
312 Macroscopic Balances-Viscous Effects Chap. 8
It is convenient to express the head 1
When a pipe discharges into a large tank ora reservoir, K= 1, and the cient K may be represented as
loss is 1 velocity head,
(v)2 K=C{
h = - (8.3-18)
m 2g
and Ct should tend toward 1.0 as ()
This amount is termed the exit loss and physically represents the complete
decreases indicating a decrease in the
dissipation of kinetic energy to interna! energy by viscous forces.
streamlined. For values of () equal te
the minor loss are of comparable roa
The diffuser that the form of Eq. 8.3-19 requires
in brackets tends to zero while the f1
If a gradual enlargement or diffuser is used to accomplish the transition
Generally the losses in sudden el
from a smaller pipe toa largerone, the loss can be reduced significantly over
vortex formation that occurs when ft
that obtained for a sudden expansion. The experimentally determined coeffi-
be supplied to maintain the vortex m
cient Ct is shown in Fig. 8.3-3. These results were obtained by Gibson 19- 2o
and this energy dissipation shows u
for water at high Reynolds numbers, and the friction and minor losses were
visualize an increase in vortex moti
represented as
therefore expect Ct to increase witl
h
f
+ h m = ef [1 - (D1)2]2
D2
(v)2
2g
(8.3-19) larger than 1.0 present a puzzling s
vortex inotion than that occurring i
increasing values of Ct with dec1
reasonable. Gibson suggested a pos
but he based his arguments on the a
(see Birkhoff intuitive hypothesis II 1

values of B. It would be difficult to


previously mentioned movie "Fiow
Aside from providing an excellent d
this movie drives home the point tt
topology before we can construct a

11'12 1T The reducer


()

The loss in head at a gradual re<


that for a diffuser. The difference n
and hence vortex formation, does 1
accounted for by the friction loss h1

The sudden contraction

The sudden contraction illustratt


reducer, for the abrupt change in
Fl1. 8.3-l. Gradual expansion or diffuser. t lt is a plausible intuitive hypothesis 1
t Prepared by Professor S. J. Kline of
19. A. H. Gibson, Proc. Roy. Soc. (London) Ser. A, 1910, 83 :366. tional Services, Inc., 47 Galen St., Watert
20. A. H. Gibson, Engineering, 1912,93 :205.
Sec. 8.3 Pipeline Systems 313
ic Balances-Viscous Effects Chap. 8

It is convenient to express the head loss in this forro, because the loss coeffi-
tank ora reservoir, K= 1, and the cient K may be represented as

(8.3-18) (8.3-20)

physically represents the complete and C should tend toward 1.0 as () -+ TT. As () -+ O, the coefficient Ct first
energy by viscous forces. decreases indicating a decrease in the minor losses as the system beco mes more
streamlined. For values of Oequal to approximately 7, the friction loss and
the minor loss are of comparable magnitude and C begins to increase. Note
that the forro of Eq. 8.3-19 requires the C-+ oo as O-+ O, because the term
is used to accomplish the transition in brackets tends to zero while the friction loss remains finite.
Ioss can be reduced significantly over Generally the losses in sudden expansions are explained in terms of the
The experimentally determined coeffi- vortex formation that occurs when flow separation takes place. Energy must
results were obtained by Gibson 19- 20 be supplied to maintain the vortex motion against the action ofviscous forces,
the friction and minor Iosses were and this energy dissipation shows up in the term, hm. It is not difficult to
visualize an increase in vortex motion as the angle O is increased,t and we
_ (D1)2]2 (v1)2 (8.3-19)
therefore expect C to increase with increasing O. However, values of C
larger than 1.0 present a puzzling situation, for it indicates a more intense
D2 2g
vortex motion than that occurring in a sudden expansion. furthermore, the
increasing values of C with decreasing values of A 2/A 1 do not seem
1=2.25
reasonable. Gibson suggested a possible explanation for this phenomenon,
but he based bis arguments on the assumption that the flow field is symmetric
(see Birkhoff intuitive hypothesis III, Sec. 1.3), and it is not, for intermedia te
values of O. It would be difficult to describe the flow phenomenon, and the
previously mentioned movie "Flow Visualization" is strongly recommended.:J:
Aside from providing an excellent description of the flow field in a diffuser,
this movie drives borne the point that we must have sorne idea of the flow
topology before we can construct a meaningful analysis.

1T
The reducer
8
"2n
The Ioss in head at a gradual reduction or reducer differs markedly from
that for a diffuser. The difference results from the fact that flow separation,
and hence vortex formation, does not occur. The loss in a reducer can be
accounted for by the friction loss h1, and we need not include the term, hm.

The sudden contraction

The sudden contraction illustrated in Fig. 8.3-4 is very different from the
reducer, for the abrupt change in cross-sectional area does Iead to vortex
ppansion or diffuser.
t It is a plausible intuitive hypothesis regarding the flow topology.
n) Ser. A, 1910, 83:366.
t Prepared by Professor S. J. Kline of Stanford University and distributed by Educa-
05.
tional Services, Inc., 47 Galen St., Watertown 72, Mass.
314 Macroscopic Balances-Viscous Effects Chap. 8
Sec. 8.3 Pipeline Systems
-- - ---------=~------------~-

hm

Square-edged
Jb
~

~
~
Well rounded
g
I-~--r ~

~
Fig. 8.3-4. Head loss in a sudden contraction.
Reentrant
formation. The loss coefficient K determined by Weisbach 21 is tabulated in ~
Table 8.3-1.
An important point to notice in Fig. 8.3-4 is that the major portion of the
loss takes place downstream from the abrupt change in cross-sectional area.
The jet formed by the sudden contraction attains a minimum area about one
Table 8.3-1
a minimum cross-sectional area ~
Loss CoEFFICIENTS FOR A SUDDEN CONTRACTION degree of vortex motion, which, of co
In practice, these minor head
~~ ~ ~ ~ ~ ~ 1~ 1~ 1~ 1~ 1 ~ 1 ~ 1 -
0
however, we can learn from these
A 2 /A 1 0.1 0.2 1 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 reduce the vortex motion and t
relatively unimportant in pipeline p.
diameter (D 2) downstream from the contraction. This mnimum area is
importance in the study of flow arm
called the vena contracta. Further downstream, viscous effects cause the jet
Numerous types of fittings may l
to spread and fill the conduit.
for st:veral appear in Table 8.3-2. T
Obviously the loss in a contraction is severely influenced by the geometry
been stlJ<:lied as thoroughly as the
of the contraction, for the geometry controls the degree of vortex motion
there may be considerable variatio
and hence energy dissipation. Three pipeline entrances from a large reservoir
8.3-2 and the actual value for a give
are illustrated in Fig. 8.3-5 along with approximate values for the head loss.
Regarding the actual desi!n of
The loss coefficient for a square-edged entrance is generally taken as K ~ 0.5,
existing pipeline, the engineer may
while the coefficient for a well-rounded entrance is K ~ 0.04 and the losses
The essential elements of the proble
can be neglected. In this latter case, as in the gradual reduction in a pipeline,
streamlining prevents the occurrence of a vena contracta and the accom- (a) geometry of the system (pipe
panying vortex formation. The re-entrant opening is the extreme case, giving entrances and exits);
21. J. Weisbach, Die Experimental Hydraulik (Freiberg, Germany : J. S. Englehardt,
(b) pumping or power requirem
1855), p. 133. (e) fluid properties (fl, p) and flo
(d) the economic constraint.
Balances-Viscous Effects Chap. 8
Sec. 8.3 Pipeline Systems 315

Squore-edged

Well rounded

(v)~
Reentront hm ~ 0 .8 Zg
by Weisbach 21 is tabulated in

.3-4 is that the major portion of the


change in cross-sectional area. Fig. 8.3-5. Losses at pipeline entrances.
attains a mnimum area about one
a mnimum cross-sectional area for the vena contracta and a maximum
degree of vortex motion, which, of course, gives rise to a maximum head loss.
0.10 0.06 0.02 o In practice, these minor head losses do not play an important role;
however, we can learn from these examples that streamlining may greatly
0.8 0.9 l.O reduce the vortex motion and the accompanying energy losses. While
This mnimum area is relatively unimportant in pipeline problems, these ideas will be of primary
importance in the study of fl.ow around immersed bodies.
viscous effects cause the jet
Numerous types of fittings may be present in a pipeline; loss coefficients
infl.uenced by the geometry for s~::veral appear in Table 8.3-2. The head losses for these fittings have not
the degree of vortex motion been st11ciied as thoroughly as the simple expansion and contraction, and
entrances from a large reservoir there may be considerable variation between the value of K listed in Table
~proximate values for the head loss.
8.3-2 and the actual value for a given piece of equipment.
is generally taken as K~ 0.5, Regarding the actual desig,n of a new pipeline, or the analysis of an
is K~ 0.04 and the losses existing pipeline, the engineer may be confronted with severa! possibilities.
The essential elements of the problem are:
the gradual reduction in a pipeline,
a vena contracta and the accom- (a) geometry of the system (pipe length and diameter, valves and fittings,
opening is the extreme case, giving entrances and exits);
(Freiberg, Germany: J. S. Englehardt, (b) pumping or power requirements;
(e) fluid properties (p,, p) and fl.ow rates;
(d) the economic constraint.
Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.3 Pipeline Systems
316

Table 8.3-2 Power requirements for a pipeline


Loss COEFFICIENTS FOR FITIINGS AND VAL VES"
In this example, we wish to det
Type of Fitting or Valve Loss Coefficient, K
system shown in Fig. 8.3-1 if it is t
galjmin of oil. The following are the
45 ell
standard 0.35 Q = 100 galjmin (volumetric flo
long radius 0.20 p = 60 lbm/ft3 (density)
90 ell .t = 10 centipoise (viscosity)
standard 0.75 D = 4 in. (pipe diameter)
long radius 0.45
square or miter 1.3 The gate valve is fully open, and the
180 bend, close return 1.5 smooth. The average velocity in the ~
Tee, standard
along run, branch blanked off 0.4
used as ell, entering run 1.3
used as ell, entering branch 1.5
branching fiow = 184fj
Coupling 0.04
Union 0.04 and the Reynolds number is
Gate valve
open 0.20 N Re= p(v) D = {(60 lbm/ft')(1840
topen 0.90 .t (10 cen
topen 4.5
4
topen 24.0 x { (4.78 x 10 centipoise)
Diaphragm valve lbr-secjft2
open 2.3
topen 2.6 Under these conditions the flow is tu
topen 4.3
formulas discussed in this section. 1
topen 21.0
Globe valve
diagram yields a friction factor of
open 6.4
topen 9.5
!=
From the macroscopic mechanical en
a From Chemlcal Englneers' Handbook, 4th ed. (New York: McGraw-Hill Book

~{~~2} =o
Company, lnc., 1963), Sec. S, p. 33.
initial ar

The route to a solution simply requires the application of the macroscopic ~p =0 The pr
entrance
mass and mechanical energy balances, along with experimental values of the pg
various head loss coefficients and friction factors. In the analysis of existing ~z =200ft
systems, the economic constraint is often of little or no concern and the 2
engineer needs to determine either (b) or (e). In the design of new systeins, L h1 = f (v) LL
(e) is usually given and the economic constraint requires that the job be done D 2g
at a "mnimum cost." This generally involves a multitude of factors and = (0.033) (1840
makes the application of the economic constraint the most difficult part of (4 in.) (2X3
the problem. However, the physics of the design are adequately taken
care of by Eq. 8.3-9, which is all we concern ourselves with in this text. = 3.11 ft
Balances-Viscous Effects Chap. 8 Sec. 8.3 Pipeline Systems 317

8.3-2 Power requirements for a pipeline system


FITIINGS ANO VALVES0

Loss Coefficient, K In this example, we wish to determine the power requirements for the
system shown in Fig. 8.3-1 if it is to handle a maximum flow rate of 100
galfmin of oil. The following are the specifications:
0.35 Q = 100 gal/min (volumetric flow rate)
0.20 p = 60 lbm/ft3 (density)
p. = 10 centipoise (viscosity)
0.75
D = 4 in. (pipe diameter)
0.45
1.3 The gate valve is fully open, and the pipe is to be considered hydraulically
1.5
smooth. The average velocity in the pipe is
0.4
(v) = 4Q = 4(100 gal/min) (231 inf
1.3
1.5 7TD 2 (3.14)(4in.) 2 gal

0.04 = 1840 in.fmin


0.04 and the Reynolds number is
0.20 N Re= p(v) D = {(60 lbm/ft~(1840 in./min)(4 in.)}
0.90 p. (10 centipoise)
4.5
24.0
X {
4
(4. 78 x 10 centipoise)
lbr-secjft 2
(_Jf_) (60minsec) ( 32lbrlbmsecft
144 in. 2
2
) } = 7600
2.3
2.6
Under these conditions the flow is turbulent, and we may use the head loss
4.3
21.0 formulas discussed in this section. The smooth pipe curve on the Moody
diagram yields a friction factor of
6.4
9.5 != 0.033
From the macroscopic mechanical energy balance given by Eq. 8.3-9 we note
ed. (New York : McGraw-Hill Book

~{~~2} =o initial and final velocities are zero

the application of the macroscopic ~p =0 The pressure in the reservoirs at the


entrance and exit Ievels is a constan!
with experimental values of the pg
factors. In the analysis of existing ~z =200ft
of little or no concern and the
(e). In the design of new systeins, 2, h1 = f (v? 2, L
requires that the job be done D 2g
involves a multitude of factors and 2
= (0.033) (1840 in./min) ( ft)
constraint the most difficult part of 310
(4 in.) (2)(32.2 ftjsec 2)
the design are adequately taken
ourselves with in this text. = 3.11 ft
318 Macroscopic Balances-Viscous Effects Chap. 8
Sec. 8.3 Pipeline Systems

The su m of the minor losses takes a particularly simple form beca use the pipe
diameter is constant.
Gate Entrance Exit
Elbows valve loss loss
2
,L hm = (v?
2g
,.L K= (v) {2(0.75)
2g
+ 0.2 + 0.5 + 1.0}
2
= (1840 in./min) ( 3 .2 ) = 0 .32 ft
2
(2)(32.2 ft/sec )
As often happens, the minor losses are quite small, and, in fact, disregarding
these losses would have caused an error of only 10% in the viscous losses.

Substitution of these results into Eq. 8.3-9 yields


L Hw =200ft + 3.11 ft + 0.32 ft
= 203ft
Solving for the power required, W, gives
w= (203ft) (pgQ) Fig. 8.3-6.
= 2,720 lbt ft/sec
or
W = 5.0hp would require specification of either
a given ftow rate at the single outlet.
Hence the pump must be capable of delivering 5.0 hp to the fluid if the maxi-
conditions must be satisfied.
mum ftow rate is to be obtained.
In studying this example, the student should stop and give sorne thought
l. The sum of the pressure dron
to the ideas behind it. We might list them as follows:
the pressure is a continuous, ,
l. the /aws of mechanics leading to the Navier-Stokes equations and the 2. The mass balance must be sa1
mechanical energy equation; 3. The mechanical energy balan(
2. the principie of conservation ofmass;
3. kinematics-the Reynolds transport theorem leading to the Navier- In single pipelines, only condition 3
Stokes equations; represent obvious physical concepts
4. the concept of stress; networks is simply one of algebraic
5. mathematics-the general transport theorem leading to the macro- The problem is solved by first sp
scopic balances. en ter and lea ve the system and of tht
The solution of pipeline problems by Eq. 8.3-9 is obviously a straightforward values are chosen so that the mas
problem; however, understanding what one is doing is by no means a simple assumed ftow distribution is used to
matter. system that generally does not satis
calculation, adjustments are made
conditions are satisfied. An efficie1
Pipeline networks
described by Venard, 23 and we shall r
Very often piping systems may be much more complex than the one 22. H. Cross, "Analysis of Flow in 1
illustrated in Fig. 8.3-1, and the ftow from a given outlet may come from 11/inois Eng. Expt. Sta. Bu/l., 286, 1936.
severa! sources. Such a system is illustrated in Fig. 8.3-6, the design of which 23. J. K. Venard, "One-Dimensional
Streeter, ed. (New York: McGraw-Hill Bo
ic Balances-Viscous Effects Chap. 8
Sec. 8.3 Pipeline Systems 319
!y simple form because the pipe
Gate Entrance Exit
valve loss loss

1.0}

uite small, and, in fact, disregarding


of only 1O% in the viscous losses.

Fig. 8.3-6. Pipeline network.

would require specification of either the pipe or pump size required to provide
5.0 hp to the fluid if the maxi- a given flow rate at the single outlet. In pipeline networks, the following three
conditions must be satisfied.
should stop and give sorne thought
as follows : l. The sum of the pressure drops around each circuit must be zero-i.e.,
the pressure is a continuous, single-valued function.
Navier-Stokes equations and the 2. The mass balance must be satisfied at each junction.
3. The mechanical energy balance must be satisfied for each section.
theorem leading to the Navier-
In single pipelines, o ni y condition 3 has to be considered; however, 1 and 2
represent obvious physical concepts, and the problem of designing pipeline
networks is simply one of algebraic complexity.
theorem leading to the macro-
The problem is solved by first specifying the flow rates of all streams that
en ter and leave the system and of the streams within the system. These initial
8.3-9 is obviously a straightforward values are chosen so that the mass balance requirement is satisfied. The
is doing is by no means a simple assumed flow distribution is used to determine the pressure everywhere in the
system that generally does not satisfy condition l. On the basis of the first
calculation, adjustments are made and the process repeated until all the
conditions are satisfied. An efficient method, worked out by Cross, 22 is
described by Venard, 23 and we shall not discuss this subject further in this text.
much more complex than the one
22. H. Cross, "Analysis of Flow in Networks of Conduits or Conductors," Univ.
a given outlet may come from 11/inois Eng. Expt. Sta. Bu/l., 286, 1936.
in Fig. 8.3-6, the design of which 23. J. K . Venard, "One-Dimensional Flow," Handbook of Fluid Mechanics, V. L.
Streeter, ed. (New York: McGraw-Hill Book Company, lnc., 1961).
Sec. 8.4 Unsteady Flow in Closed Conduiu
320 Macroscopic Balances--Viscous Effecu Chap. 8

In industrial practice, pipeline networks can become exceedingly complex, The boundary conditions for this proll
and digital computers are required to perform the analysis. Ingels and B.C. 1 : v. = O, t = O,
Powers 24 have presented a useful method of digital computation. B.C. 2: v. is finite, for
B.C. 3: v. =o, r =ro
8.4 Unsteady Flow in Closed Conduits
B.C.4:
op
-=---
(Pi -
oz
In general, engineers would like to deal with processes operating at or near The solution of this equation is beyon
steady state conditions. Unstable or oscillating systems are usually undesir- a knowledge of partial differential equ
able, and considerable effort is made to control processes at sorne steady by Szymanski, 25 and it may be express
state. Unsteady state conditions naturally occur during the start-up and
shut-down of any process, or if the steady-state operating conditions are (v.) = 1 - (0.963e-0 3681
changed. Very often it is important to know the time required to reach the (v.)oo
where (v.) 00 is the average velocity for t
of an infinite series of exponential ter
for satisfactory results. The dimensio
. .
t
0 =-

~---------L------~1
h and/00 is the friction factor at t = oo.

Joo =
___ _l____ p=p, If the final fl.ow is turbulent, the
involving an initial transition in the la
the turbulent region. lf the final Reync
will be turbulent during most of the tr
approximate solution by the mechani<
If the liquid depth h undergoes a r
Fig. 8.4-1 . Unsteady flow in a pipeline.
the pressure drop in the tube may be

steady state, and we will apply the mechanical energy balance to two transient flp =Jo
pg
fl.ows to illustrate the important steps in the analysis. Figure 8.4-1 illustrates where
a supply and pipeline arrangement for which we wish to determine the time
tlp
required for the ftow to reach the steady state value.
pg
lf the ftow is always laminar, and the entrance length is small compared
to the tube length, the differential equations of motion reduce to Taking the tube as the control volum

(8.4-1)
~dt J(tpv 2
) dV =
-r
25. P. Szymanski, "Sorne Exact Solutim
24. D. M. Ingels and J. E. Powers, "Analysis of Pipeline Networks," Chem. Eng. Prog.,
Cylindrical Tube," J. Math. Pures Appl. Ser.
1964, 60:65.
Macroscopic Balances-Viscous Efrects Chap. 8 Sec. 8.4 Unsteady Flow in Closed Conduits 321

can become exceedingly complex, The boundary conditions for this problem are
to perform the analysis. Ingels and B.C. 1: v. = O, t = O, for O :==:::: r :==:::: r0
method of digital computation.
B.C.2: v. is finite, for O~ r :==:::: r0
B.C. 3: v. = 0, _ r = r0

B.C. 4: Op = _ _,_,_(P-=--i----"'P-"'-o)
t >o
oz L
to deal with processes operating at or near The solution of this equation is beyond the scope of this text, for it requires
or oscillating systems are usually undesir- a knowledge of partial differential equations. The result has been presented
to control processes at sorne steady by Szymanski, 25 and it may be expressed as
naturally occur during the start-up and
steady-state operating conditions are (v.) = 1- (0.963e-03691 oo + 0.036e-1. 9091 oc + ) (8.4-2)
to know the time required to reach the (v.)oo
where (v.) 00 is the average velocity for t = oo. The complete solution consists
of an infinite series of exponential terms, but only the first two are necessary
for satisfactory results. The dimensionless time 0 is defined as

0 = t(v.) 00
(8.4-3)
D
~---------L--------~~1 and loo is the friction factor at t = oo.

foo=~ (8.4-4)
NRe,oo
If the final flow is turbulent, the transition will be a complex process
involving an initial transition in the laminar region and a final transition in
the turbulent region. Ifthe final Reynolds number is large (say, 106), the flow
will be turbulent during most of the transition period, and we may obtain an
approximate solution by the mechanical energy balance.
If the liquid depth h undergoes a negligible change during the transition,
the pressure drop in the tube may be expressed as

~'-"''all ...,a,
energy balance to two transient ~p = f oo (
pg
L) (v.);,
D 2g
(8.4-5)
in the analysis. Figure 8.4-1 illustrates
where
which we wish to determine the time
state value. ~p = h
the entrance length is small compared pg
of motion reduce to Taking the tube as the control volume, the mechanical energy balance gives

(8.4-1) ~~ f (!pv 2
) dV = f V t(n) dA - E, (8.4-6)
-y A,

of Pipeline Networks," Chem. Eng. Prog., 25. P. Szymanski, "Sorne Exact Solutions of the Equations of Motion for Flow in a
Cylindrical Tube," J. Math. Pures Appl. Ser. 9, 1932, 11:67.
322 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.4 Unsteady Flow in Closed Condut

Assuming the velocity profile to be flat (it certainly won't be during the early Both Eqs. 8.4-14 and 8.4-2 indicate t
stages), and representing the stress vector at the entrance and exit as the time becomes infinite; however, f~
say that steady state has been reache~
t(n) = -np (8.4-7) of the final value. The time is therefo
we find Cl 11.8
~ {tp(v,) AL} =
2
(v,) ~pA - Ev
1:/gg% = f 00 '
(8.4-8)
dt
Cl 5.3
Carrying out the differentiation and dividing by p(v, )Ag, we have 1:/gg% =f 00 '

(f)
g
d(v,)
dt
= ~P
pg
_ Ev
pgQ
(8.4-9) It is difficult to comment on these two
time for two distinct types of flow.
If we remember that
11.8JJ
= --
Ev = h, =
pgQ
!( L) (v,)
D 2g
2 t99%
0 (v,)oo

5.3
and if we substitute Eq. 8.4-5 into Eq. 8.4-9 and use the dimensionless time t99% =
o
--j
(v,)oo
0, we can write Eq. 8.4-9 in the form
To ga in sorne familiarity with the
could say for turbulent flowt
(8.4-iO)
foo = (
Solving this equation would be a difficult task beca use of the complex varia- D =(
tion of the friction factor f with the Reynolds number. A numerical solution (v,) = (
would be straightforward (and might make an interesting class project),
but to obtain an analytic solution we will assume Under these conditions, the time wo1

j_ >:::! 1 (8.4-11) t99% = O[(0.00:(5.


foo
We separate variables and integrate to obtain = 0[1.77 n
From this order of magnitude calcul~
2f 1 ~;2 =Joof d0 +e
z
(8.4-12) becomes significant for large diamet
that this analysis assumed that the le
The approximation given by Eq. 8.4-11 will not be valid at short times when the entrance length and is therefore 1
the flow is laminar; on the other hand, if the major portion of the transient
time is associated with the rough-pipe region of flow, the approximation Oscillating systems-the U-tube r
might be quite satisfactory. Integration of Eq. 8.4-12, and application of the
boundary condition Various fluid systems are subject
latory motion. Actual processes ma)
B.C. 1: U,= O, 0 =0 (8.4-13) analogue or digital computers may be
yield the final solution even approximately. The purpose ol
gain sorne insight into the import
(8.4-14)
t Read (0.005) as "on the order of rm
Balances-Viscous Effects Chap. 8 Sec. 8.4 Unsteady Flow in Closed Conduits 323

(it certainly won't be during the early Both Eqs. 8.4-14 and 8.4-2 indicate that the final velocity is not reached until
at the entrance and exit as the time beco mes infinite; however, for practical purposes it is satisfactory to
say that steady state has been reached when the velocity is within 1 per cent
(8.4-7) of the final value. The time is therefore given as
1 8
0 1. laminar flow
= (v.) 6.pA - Ev (8.4-8) 99% = foo '
5.3
viding by p(v. )Ag, we have 0 99% = foo ' turbulent flow

6.p _ Ev
(8.4-9) It is diffi.cult to comment on these two values, for they represent the transient
pg pgQ time for two distinct types of flow. The real times for the two cases are
11.8D "!> n~ b re
!(.!:.)D (v.)2
2g
t 99 % = ( -J.) ,
v. 00 00
laminar

5.3D
t ---- turbulent
8.4-9 and use the dimensionless time 99% - (
vz
)
00
J. '
00

To gain sorne familiarity with the orders of magnitude of these times, we


(8.4-O) could say for turbulent flowt
/oo = O(0.005)
task beca use of the complex varia- D = 0 (l ft)
number. A numerical solution
(v.) = O(1 ftjsec)
make an interesting class project),
Under these conditions, the time would be

(8.4-11) 199
% =
o[ (5.3)(1 ft) min
(0.005)(1 ftjsec) 60 sec
J
= 0[1.77 min]
From this order of magnitude calculation we see that the transient time only
(8.4-12) becomes significant for large diameter tubes and low flow rates. Remember
that this analysis assumed that the Iength of the pipe was long compared to
will not be valid at short times when the entrance length and is therefore not valid for short pipes.
if the major portion of the transient
region of flow, the approximation Oscillating systems-the U-tube manometer
of Eq. 8.4-12, and application of the
Various fluid systems are subject to upsets which may give rise to oscil-
latory motion. Actual processes may become extremely complex, and large
0 =0 (8.4-13)
analogue or digital computers may be required to determine the fluid motion
even approximately. The purpose of studying the U-tube manometer is to
gain sorne insight into the important aspects of oscillatory systems. In
(8.4-14)
t Read 0(0.005) as "on the order of magnitude of 0.005."
324 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.4 Unsteady Flow in Closed Conduits

addition, this example is of sorne interest because it can easily be studied For the moving control volume i'a(t)
experimentally, thus introducing the student to the errors incurred in an that the gas-liquid interface is horizon
approximate solution. 8.4-2. Actually, it is not because the
The U-tube manometer to be analyzed is shown in Fig. 8.4-2. Once again, interface will be as shown in the inset.
the differential equations describing the motion are too complex to be solved, Considering Eq. 8.4-15, we note th1
and the mechanical energy balance is required. We start with the complete
equation and note that W is zero to obtain V=W

~t J(!pv J and the convective transport term on


dV +
2 2
) (!pv )(v - w) n dA energy balance is identically zero. If vi!
i'"0 (t) A,(t) are neglected, we may assume that
= J V t(n) dA- J pgzv n dA - Ev (8.4-15) V ~n> hert =
A,< O .91' aW
and the first term on the right-hand

( - - - - -- - - '
__("' Gas pressures suddenly equilibrated
assumption is open to question, becaus
depends upon whether it is advancing
solve the detailed problem associated v
assume it is ftat and make the assump

Tho
Final position of
manometer _________________ l __ _ z=O
1 If the flow in the manometer were t
fluid

8.4-17 could be replaced by average


term could be expressed as in the pr
ftow in a manometer is likely to be la
mate analysis for laminar flow. Assur
1 - - - i + - - lnitial pasition of
manometer fluid one-dimensional, we write

v(r, t) = 2(v)l
Here, v(r, t) represents the component
Expressing the transient velocity profil
is a common technique (and one whi
obtaining approximate solutions to eng
method depends on the variations witl
velocity profile is never far from par
Manometer liquid Advancing liquid Receding liquid we shall explore the validity of this as!
The average velocity (v) in Eq. 8.4

u
defined as positive when the flow is fr<

"''""' '"'"'"" 11111 V n heft

and the position of the two interfaces


=

Liquid Liquid
Fig. 8.4-2. U-tube manometer. Z = -h heft
ic Balances-Viscous Effects Chap. 8 Sec. 8.-4 Unsteady Flow in Closed Conduits 325

because it can easily be studied For the moving control volume "Ya(t) we choose the fluid itself and assume
to the errors incurred in an that the gas-liquid interface is horizontal and well defined as shown in Fig.
8.4-2. Actually, it is not because the fluid adheres to the tube wall, and the
interface will be as shown in the inset.
Considering Eq. 8.4-15, we note that
V=W on ~a(t)

and the convective transport term on the left-hand side of the mechanical
oenergy balance is identically zero. If viscous effects at the gas-liquid interface
.are neglected, we may assume that
dA- f pgzv n dA- E" (8.4-15) V t.:n) hert = -V t<n> lrlght (8.4-16)
""(t) and the first term on the right-hand side of Eq. 8.4-15 is also zero. This
Gas pressures suddenly equilibrated
assumption is open to question, beca use the nature of the interface obviously
depends upon whether it is advancing or receding. However, we must either
solve the detailed problem associated with the moving interface, or we must
assume it is flat and make the assumption indicated by Eq. 8.4-16 to obtain

T ~f (!pv 2
f
= -
) dV pgzv n dA - E, (8.4-17)
ho 'j'"G(t) Jll'G(t)

______ l __ _ z=O If the flow in the manometer were turbulent, the various ve1ocities in Eq.
8.4-17 could be replaced by average velocities and the energy dissipation
term could be expressed as in the previous example. However, oscillating
J flow in a manometer is likely to be laminar, and we must devise an approxi-
of mate analysis for laminar flow. Assuming the flow to be quasi-steady and
one-dimensional, we write

v(r, t) = 2(v)[ 1 - (~)] (8.4-18)

Here, v(r, t) represents the component of the velocity vector along the tube.
Expressing the transient velocity profile in terms of the steady state solution
is a common technique (and one which should be used with caution) for
obtaining approximate solutions to engineering problems. The success ofthis
method depends on the variations with time being small enough so that the
velocity profile is never far from parabolic. After obtaining the solution,
Receding liquid
we shall explore the validity of this assumption.
The average velocity (v) in Eq. 8.4-18 is a function of time, and will be

u L iquid
defined as positive when the flow is from left to right. Since
V n hen

and the position of the two interfaces is given by,

Z = -h hert>
= -V

z=
n lrlght

+h lrtght
(8.4-19)

(8.4-20)
Sec. 8.4 Unsteady Fl 0 w in Closed Conduits
326 Macroscopic Balances-Viscous Effects Chap. 8

time 0 as,
We may simplify the area integral fn Eq. 8.4-17 to obtain
h
H=-
~1(4p7TL(v)2f" [1 - (~rJ r dr) = - 2pgh(v)7Tr~ - J
2 ho
<1> dV (8.4-21)
o r-.(t) . 0 =t
From Table 7.3-1, we find the dissipation function to be and Eq. 8.4-24 reduces to

(!r ovao + ~)r + (av.) d H + 2{3 dH


2
= 2~-'[av~ +
2 2
8
<1> ]
or oz d02 dG

+ ~-'[' :,(;) + ~(:;r) r + ~-'[;(:;) + (::8) r where

+ ~-'[ (::r) + (:~)


In this case, there is only one component ofvelocity, v.
r = v, and <1> reduces to
In terms of the dimensionless variables,
B.C. 1': H= 1,

B.C. 2': dH =O
16J,L(v)2 (~)
2

4> = (8.4-22) d0 '


The solution of this second-order, ordin
Substituting this result into Eq. 8.4-21 and carrying out the integration, we by assuming solutions of the form
get
H=
(8.4-23) which when substituted into Eq. 8.4-26
eme(m2 + 2{3
Because the time rate of change of height of the gas-liquid interface is equal
to the average velocity, This equation will be satisfied, and E
differential equation, if m is chosen so t
dh
- = (v) m2 + 2{3m
dt
Thus, m has two possible values, given
We may rearrange Eq. 8.4-23 to obtain
m1 = -({3 +
dh+ (6v)dh + (3g)h = 0
2
(8.4-24)
dt 2 r~ dt 2L m2 = -({3-

This equation is to be solved subject to the boundary conditions Provided m 1 and m2 are distinct (i.e., {3

B.C. 1: h = h0 , t=O (8.4-25a) H = Cemle


However, if the roots of Eq. 8.4-30 are
B.C. 2: dh =0 t =o (8.4-25b)
dt , H = (C1 +
It will be helpful to put the differential equation and the boundary conditions 26. L. R. Ford, Differential Equations (N
Inc., 1933), p. 70.
in dimensionless form. To do so, we define a dimensionless distance H and
Chap. 8
Sec. 8.4 Unsteady Flpw in Closed Conduits 327

. 8.4-17 to obtain
time e as,
h
H=-
ho
= -2pgh(v)1Tr~- J<1> dV (8.4-21)
7".(1) E>=t
~-u
!3i
and Eq. 8.4-24 reduces to

+ (~:n d2H
d82
+ 2{3 dH + H = O
dE>
(8.4-26)

)r + ~t[;(~~) + (~:8) r where


fJ = 3v f2L
r~ ~Ji
In terms of the dimensionless variables, the boundary conditions are

ofvelocity, v. = v, and <1> reduces to


B.C. 1': H= 1, e =O (8.4-27a)

B.C. 2:: dH =O 8=0 (8.4-27b)


(8.4-22) dE> ,
The solution of this second-order, ordinary, differential equation is obtained
and carrying out the integration, we by assuming solutions of the form
H= eme (8.4-28)
(8.4-23) which when substituted into Eq. 8.4-26 gives

of the gas-liquid interface is equal


em 9 (m 2 + 2flm + 1) = O (8.4-29)
This equation will be satisfied, and Eq. 8.4-28 will be a solution of the
differential equation, if m is chosen so that
m2 + 2flm + 1 = O (8.4-30)
Thus, m has two possible values, given by

m1 = -({3 + .Jf32 - 1) (8.4-31a)


(8.4-24)
m2 = -({3- .Jf3 2
- 1) (8.4-31b)
the boundary conditions Provided m1 and m 2 are distinct (i.e., fJ =1= 1), the general solution is

t =o (8.4-25a) H = C1 em1 e + C emsE> 2 {8.4-32)


However, if the roots of Eq. 8.4-30 are not distinct, the solution is 26
t =o (8.4-25b)
H = (C1 + C2 8)e-e (8.4-33)
and the boundary conditions 26. L. R. Ford, Differential Equations (New York: McGraw-Hill Book Company,
a dimensionless distance H and lnc., 1933), p. 70.
328 Macroscopic Balances-Viscous Effects Chap. 8
Sec. 8.4 Unsteady Flow in Closed Conduits

The solution is again given by


complex function containing bo
part independently satisfies the
conditions, and takes the forro

This result indica tes that H takes


which decrease exponentially wi
damped."

Without the preceding analysis, we


-i.OL------------:8::--------

-
Fig. 8.4-3. Position of the gas-liquid interface for an
oscillating manometer.
could occur in a mano meter. If viscous
that the fluid should slowly and steadil
But if inertial effects predominate, we
behave somewhat like a pendulum, os
tion. Since {3 is, in effect, the ratio o
dependence of the motion on {3 is just
The motion of the fluid in the manometer may have three distinct modes
though the analysis agrees with our int
depending on the value of {3. The three types of motion are illustrated
question, "U nder what conditions can
graphically in Fig. 8.4-3. We may discuss them as follows .
We can answer this question, in part,
l. {3 > l. For this case, both m1 and m 2 are real negative numbers, and analysis.
the constants of integration in Eq. 8.4-27 are readily determined to
yield Order of magnitude analysis
(m 2emt9 _ m1em28)
H=~::.:.-----2..::...~ (8.4-34)
(m 2 - m1) Two major assumptions were impo
This type of motion is called "over damped," and the fluid approaches l. The gas-liquid interface was flat.
its equilibrium position exponentially. 2. The velocity profile was parabol
2. {3 = l. Here, the roots are indistinct, and the solution is obtained by
application of the boundary conditions to Eq. 8.4-33, yielding The first assumption is best investigat1
the second assumption can be discuss1
H = (1 + 0)e-9 (8.4-35) away from the interface and the curve1
The fluid motion is similar to the previous case in that the equilibrium flow will be one-dimensional, and the t
position is approached exponentially. However, this solution repre-
sents the most rapid approach to equilibrium without "overshooting" av --+
p-=
Jp
P<
the equilibrium position. Such motion is called "critically damped." Jt Jz
3. {3 < l. Under these conditions, the roots become complex, and we
From previous work we know that if
must take special care in obtaining a solution. We write the roots as
compared to the viscous term, the velo
{3<1 (8.4-36) Making use of Eq. 8.4-18, we can p
tude analysis using the solution to Eq. 8
ic Balances-Viscous Effecu Chap. 8
Sec. 8.4 Unsteady Flow in Closed Conduiu 329

The solution is again given by Eq. 8.4-32; however, the result is a


complex function containing both real and imaginary parts. The real
part independently satisfies the differential equation and the boundary
conditions, and takes the form

fJ<l
(8.4-37)
This result indica tes that H takes on both positive and negative values
which decrease exponentially with time. This motion is called "under
damped."

Without the preceding analysis, we might easily guess that such motions
o
-
gas-liquid interface for an
could occur in a mano meter. If viscous effects predominate, it seems natural
that the fluid should slowly and steadily approach the equilibrium position.
But if inertial effects predominate, we would certainly expect the system to
behave somewhat like a pendulum, oscillating around the equilibrium posi-
tion. Since fJ is, in effect, the ratio of viscous forces to inertial forces, the
may have three distinct modes dependence of the motion on fJ is just what we would expect to find. Even
types of motion are illustrated though the analysis agrees with our intuition, we should still ask the important
them as follows. question, "Under what conditions can we expect this result to be accurate?"
We can answer this question, in part, by performing an order of magnitude
m2 are real negative numbers, and analysis.
Eq. 8.4-27 are readily determined to
Order of magnitude analysis
(8.4-34)
Two major assumptions were imposed on this analysis.
l. The gas-liquid interface was fiat.
2. The velocity profile was parabolic-i.e., quasi-steady fiow.
and the solution is obtained by
to Eq. 8.4-33, yielding The first assumption is best investigated experimentally, but the validity of
the second assumption can be discussed mathematically. At sorne distance
0)e- 9 (8.4-35)
away from the interface and the curved portion of the manometer tu be, the
previous case in that the equilibrium fiow will be one-dimensional, and the equations of motion reduce to
y. However, this solution repre-
equilibrium without "overshooting" av a (rav.)
- -op + pg. + p,-
is called "critically damped."
P-
ot = oz -
or or (8.4-38)
the roots become complex, and we
a solution. We wriie the roots as From previous work we know that if the local acceleration term is small
compared to the viscous term, the velocity profile will be parabolic.
fJ<l (8.4-36) Making use of Eq. 8.4-18, we can perform the following order of magni-
tude analysis using the solution to Eq. 8.4-26 to estmate the magnitude ofthe
330 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.5 Flow Rate Measurement

terms in Eq. 8.4-38. Attacking the local acceleration term first, we writet Here we see that the inequality given b

ov =o( P o<v>
) =o( P d
but will be satisfied for all dimensionl
P ot ot dt
h) 2

2
(8.4-39) damped flows we may expect reasonabl)
experiment, except for very short times
The viscous terms are expressed as

(8.4-40) 8.5 Flow Rate Measurement


Our requirement that the local acceleration term be small compared to the
viscous term immediately leads us to the inequality, There are severa! common devices
meter, Pitot tube, and the sharp-crested

1 ~~ ~ ~ ~~
1 1 1 (8.4-41)
velocities. It is of interest to analyze
engineering practice, and since these
or, in dimensionless form, experimentally, we have an opportunity
(8.4-42)
Venturi meter
If the system is under damped ({3 < 1), the first derivative dHfd0 takes
on zero values; at these times the inequality can never be satisfied. For the
The Venturi meter, illustrated in Fig
under damped case, we can expect only qualitative agreement between the
constriction that accelerates the flow
derived result and a real manometer.
pressure; this process is followed by a
For the critically damped case our solution indicates that
loss to be minimized. The pressure di
d2H = -(1- El)e-9, dH = -ee-9 related to the flow rate by the appli
d8 2 dEl This presents a situation for which
Thus, mentum balance because the pressure

~~~~ ~ ~~~1 for e -o


"1"
...
~~~~ ~ ~~~1
1
for e- 00 1
1

and we can again expect only qualitative agreement between theory and
1
" -
experiment. This situation is left rather wide open, as it should be. If the
Z4--------+----
agreement between theory and experiment were poor, we should not be sur- l
prised. Neither should we be surprised if the agreement were good, for it 1
L--/.
...:
would only mean that significant deviations from the parabolic profile did
not greatly alter the value of the viscous dissipation term.
The analysis for the over damped flow ({3 > 1) is more complex than z2 _____ ----- -----
the other two cases; however, if {3 ~ 1 we may write

z1---------

t The symbol O should be read as "the order of magnitude of." This type of analysis
will be discussed in detail in Sec. 11.3. Fil. 8.5-1. Ver
ic Balances-Viscous Effects Chap. 8 Sec. 8.5 Flow Rate Measurement 331

acceleration term first, we writet Here we see that the inequality given by Eq. 8.4-42 is not satisfied for 0--+- O,
but will be satisfied for all dimensionless times greater than {3-1 For over
(8.4-39) damped flows we may expect reasonably good agreement between theory and
experiment, except for very short times.

.t(v)) = O(/!:._ dh) (8.4-40)


( r~ r~ dt 8.5 Flow Rate Measurement
term be small compared to the
the inequality, There are severa! common devices-such as the orfice meter, Venturi
meter, Pitot tube, and the sharp-crested wier-used to measure flow rates and
'}/ dh 1 (8.4-41) velocities. lt is of interest to analyze these systems as an introduction to
1 r~ dt engineering practice, and since these metering devices have been studied
experimentally, we have an opportunity to compare analysis with experiment.
(8.4-42)
Venturi meter
< 1), the first derivative dHfd0 takes
can never be satisfied. For the
qualitative agreement between the The Venturi meter, illustrated in Fig. 8.5-1, consists of a relatively abrupt
constriction that accelerates the flow and produces a decrease in the fluid
so1ution indicates that pressure; this process is followed by a gradual expansion allowing the head
loss to be minimized. The pressure difference between points 1 and 2 can be
-dH = -0e-9 related to the flow rate by the application of the macroscopic balances.
d0 This presents a situation for which we should definitely not use the mo-
mentum balance because the pressure (and thus the surface force) varies
for 0--+- O
"1" "2"
1
1
for 0 --+- oo 1
Control volume
1

agreement between theory and


wide open, as it should be. lf the --,
were poor, we should not be sur- ---:-'--- - - - - - - -
if the agreement were good, for it
from the parabolic profile did
s dissipation term.
flow (/3 > 1) is more complex than
1 we may write

21---------
1
Gravity

of magnitude of." This type of analysis Z=O-----------

Flg. 8.5-1. Venturi meter.


332 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.5 Flow Rate Measurement

from 1 to 2 in an unspecified manner.t The mechanical energy balance is well include an estimate of the effect of t
suited to this type of problem, and for steady, incompressible flow, the pressure; however, we shall neglect this
various terms in Eq. 7.3-31 are by the centerline pressure to obtain

f lpv2y n dA
.ti.,
= -lp(v
3
) 1 A1 + lp(v 3
) 2A2 (8.5-1) Q=A2a

If this formula is expressed in terms o


f V t<n> dA= +(pv) 1A:- (pv) 2A 2 (8.5-2) z 2 - z1 yields
.ti.,
Q =A 2 J2(pm l
p[l
Jpcf>v n dA= -p(<f>v) A 1 1 + p(cfv) 2A2 (8.5-3)
where Pm is the density of the manome
..~~.. Experimental studies indicate that
For the present, we shall assume that viscous effects are negligible and write accurate for engineering practice, anda
to give
E" = o (8.5-4)
Q = CaA2 {2(;;:
Substituting Eqs. 8.5-1 through 8.5-4 into Eq. 7.3-31, and assuming flat .J~
velocity profiles, we find
A plot of Ca versus Reynolds numbe~
lp[(v)~A 2 - (v)~A 1 ] = (v)1(p) 1A1 - (v)a(p) 2A2 meter with (A 2/A 1) = 0.25. We see tha
-p[(cf)2(v)2A2- (cf)(V) 1A1 ] (8.5-5) sis is good because Ca is nearly unity.
we should not expect the analysis to ho
Application of the macroscopic mass balance velocity profiles are no longer flat. A
(v) 2A 2 = (v) 1A 1 (8.5-6) profile into aC<:ount, can lead to the
rapidly as laminar flow is approached; l
allows Eq. 8.5-5 to be simplified to exercise for the student.
2
(v)~ - (v)~ = p [(p + pcf>)t - (p + pcf>)2] (8.5-7) 1.00

The mass balance may be used to eliminate (v) 1 and obtain an expression for
the volumetric flow rate

J 2 [(p + pcf>) + pcf>)2]


/'
Q= (v) A = A (p 0.90
1 -
(8.5-8)
2 2 2 p[l - (A2/At)2] /
Now, 4> is a linear function in z; therefore, the average value is the centerline
value, and
(8.5-9)
V
0.80
However, the pressure at point 2 is not a linear function of z because the
streamlines are curved,t and strictly speaking we cannot replace the average
pressure by the pressure at the centerline. A more detailed analysis might 2
1/ 4 681Ql 2 4
NRe=
t Does this suggest a differential-macroscopic balance as a possibility?
! See Sec. 7.4, Eq. 7.4-21. Fl1. 8.5-:Z. Venturi ~
Chap. 8 Sec. 8.5 Flow Rate Measurement 333

The mechanical energy balance is well include an estimate of the effect of the curved streamlines on the average
for steady, incompressible flow, the pressure; however, we shall neglect this effect and replace the average pressure
by the centerline pressure to obtain

(8.5-1) Q- A
-
2
J 2(p - P2)
p[1 - (A 2/A1) 2]
(8.5-10)

If this formula is expressed in terms of the height of the manometer fluid,


(8.5-2) z 2 - z1 yields

(8.5-11)
(8.5-3)
where Pm is the density of the manometer fluid.
Experimental studies indicate that Eq. 8.5-11 is not always sufficiently
viscous effects are negligible and write accurate for engineering practice, and a discharge coefficient e, is introduced
to give
(8.5-4)
into Eq. 7.3-31, and assuming flat (8.5-12)

A plot of e, versus Reynolds number is shown in Fig. 8.5-2 for a Venturi


- (vMp)2A2 meter with (A 2/A 1) = 0.25. We see that at high Reynolds numbers the analy-
- (t/>)1(v)1A 1 ] (8.5-5) sis is good because e, is nearly unity. For Reynolds numbers less than 2300,
we should not expect the analysis to hold because the flow is laminar and the
velocity profiles are no longer flat. A more carefulanalysis, takingthevelocity
(8.5-6) profile into account, can lead to the conclusion that e, should decrease
rapidly as laminar flow is approached; however this problem will be left as an
exercise for the student.

(8.5-7) 1.00
lo-"
~
(v)1 and obtain an expression for ~v
V
(8.5-8) 0.90
V
/
eti
the average value is the centerline

(8.5-9)
0 .80
V
~~"""'F.we cannot replace the average
A more detailed analysis might 2
V 4 6 8 103 2 4 6 810" 2 4 6 8 1dl
Nrt.= (v) 1D1/V
balance as a possibility?
Fl. 8.51. Venturi discharge coefficient.
Sec. 8.5 Flow Rate Measurement
334 Macroscopic Balances-v 1scous Effects Chap. 8

Orfice meter
t---.... !'---
The sharp-edged orfice shown in Fig. 8.5-3 is another device for accelerat- 0 .80
ing the flow and obtaining a measurable pressure difference related to the
flow rate. The orfice is often used as a metering device in preference to a !'--t---.... t-.
Venturi meter because of its versatility and low cost of construction and
installation. Because the area of the jet at point 3 (the vena contracta) is

r--- ~J
Control volume t--r--..,
"1'' "2"

... r--..t- ... t--r--.., t-11


1
1
1
1
0.60
...
l-..

2 4 6 8 4 2 4 6
10

Fig. 8.5-4. Discharge coefficien '

Our analysis would indicate that

Ca.=~
'V [1- (

Flg. 8.5-3. Sharp-edged orifice. Because Ce changes with A2/A 1 it is ratl


with experimental results. However, as
unknown, the solution to this flow problem will not be as accurate as that for coefficient should equal the contraction e
the Venturi meter. Calibration of orfice meters is therefore a necessity, and, simplified. Figure 8.5-4 shows sorne exp1
in general, they are considered less reliable than Venturi meters. sharp-edged orfice. For small values o
If we express the area of the jet at the vena contracta as approximately 0.6, a reasonable value
(8.5-13) again the analysis is only satisfactory at

where A2 is the area of the orfice and Ce is the contraction coefficient, the
flow rate is given by Flow nozzle

Q-
_ CA
e
J
2
2(p - Pa)
p[1 - (CeA2/A1) 2]
(8.5-14) A flow nozzle, such as that illustrat
difficulty of an unknown contraction cot
Here we have applied the mechanical energy balance, neglected viscous parallel as the fluid emerges from the nm
effects, and replaced average pressures with centerline pressures. Because the is applied as before to yield
contraction coefficient is a function of the ratio of areas A2/ A1 , the equation
is generally written in terms of a discharge coefficient
Q = Ca.A 2~ 2(p 1 - p3 )/p (8.5-15)
Sec. 8.5 Flow Rate Measurement 335

-..... r-...
8.5-3 is another device for accelerat- 0.70
0 .80
pressure difference related to the 1~
a metering device in preference to a
and Iow cost of construction and
1'-. ~
0 .60
jet at point 3 (the vena contracta) is

... ..............

r-- 0 .40
........
...... t- """'... ..._........ 0.20
0.60
...... 0.10
0.05
2 4 6 8104 4 6 8105
2 2 4 6 81o6 2

NR =<v>1D,fv
Fig. 8.5-4. Discharge coefficients for a sharp-edged orifict;.

Our analysis would indicate that

(8.5-16)

Because c. changes with A 2/A 1 it is rather difficult to compare our analysis


with experimental results. However, as A 2/A 1 becomes small, the discharge
coefficient should equal the contraction coefficient, and matters are somewhat
meters is therefore a necessity, and, simplified. Figure 8.5-4 shows sorne experimental values of C,.for a standard
than Venturi meters. sharp-edged orfice. For small values of A 2/AI> the discharge coefficient is
approximately 0.6, a reasonable value of the contraction coeffi.cient. Once
(8.5-13) again the analysis is only satisfactory at high Reynolds numbers.

c. is the contraction coefficient, the


Flow nozzle

(8.5-14) A flow nozzle, such as that illustrated in Fig. 8.5-5, does not offer the
difficulty of an unknown contraction coeffi.cient, because the streamlines are
energy balance, neglected viscous
parallel as the fluid emerges from the nozzle. The mechanical energy balance
with centerline pressures. Because the is applied as before to yield
the ratio of areas A 2/A 1 , the equation
coefficient
(8.5-17)
(8.5-15)
336 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.5 Flow Rate Measurement

"1" "2"
1
The experimental values of Cd shown
1 is in error by about 10 per cent or less,
area to the duct area, A 2/A 1 As the n
erated more severely, and inertial effec
the Reynolds number is greater tha1
viscous dissipation begins to show i
decreases.

Pitot-static tube

The pitot-static tube shown in Fi


velocities ata point; it may be used t
pro file is flat or if the en tire profile is m
at point 2 which is the stagnation poin
holes are drilled around the peripher_)l
static pressure. The pressure at point 2
difference between these two pressure
and relatd to the fluid velocity. In thi
the pressure along two streamlines rat

Fig. 8.5-5. Flow nozzle.

1.00 1
A)A,Lb1.~ 11
~ """"
~
A2/A1 1 1_1
1 -0.50

1 ~~ u
A2l;A 11=0.65
0 .90 f-A 2 !A 1 = 0 .30

0.80
Th
l_
2 4 6 8104 2 4 6 8105 2 4 6 81o6 2
NRe =<v>D,Iv
Fl1. 8.5-6. Discharge coefficients for a flow nozzle. Fig. 8.5-7. Pit
Sec. 8.5 Flow Rate Measurement
337

The experimental values of C shown in Fig. 8.5-6 indicate that the analysis
is in error by about 1Oper cent or less, depending u pon the ratio of the nozzle
area to the duct area, A 2/A 1 As the ratio becomes smaller the fluid is accel-
erated more severely, and inertial effects very definitely predominate, provided
the Reynolds number is greater than 105 For lower Reynolds numbers,
viscous dissipation begins to show its effect and the discharge coefficient
decreases.

Pitot-static tube

The pitot-static tube shown in Fig. 8.5-7 is a device for measuring fluid
velocities ata point; it may be used to determine the flow rate if the velocity
profile is flat or if the en tire profile is meas u red. The tu be has a small opening
at point 2 which is the stagnation point, i.e., the velocity is zero. A series of
holes are drilled around the periphery of the tube at point 3 to measure the
static pressure. The pressure at point 2 is called the dynamic pressure and the
difference between these two pressures may be measured by the manometer
and related to the fluid velocity. In this particular system, we wish to compute
the pressure along two streamlines rather than over a control surface, and we

Flow nozzle.

Boundory loyer (thickness is exoggeroted)

"1" "2"
~========::::-

~.

Th
4 6 e105 2 4 6 e o6 2 1__
1
=<V>D11v

for a flow nozzle.


Fig. 8.5-7. Pitot-static tube.
338 Macroscopic Balances-Viscous Effects Chap. 8 Sec. 8.5 Flow Rate Measurement

shall use Bernoulli's equation. Application of Eq. 7.4-17 to the streamline Sharp-crested weir
between 1 and 2 yieldst

PI + !pv~ + pgzi = P2 + tpv~ + pgz2 (8.5-18) Flow rates in open channels may
the channel over which the fluid m
Since v2 = O, the pressure at point 2 is monly used weirs, but we will consi
Fig. 8.5-8. It represents a rather e
P2 = PI + !pv~ + pg(zi - Z2) (8.5-19)
sence of experience and knowledge,
Along the streamline from points 1 to 3 we write

PI + fpv~ + pgzi = Pa + !pv~ + pgza (8.5-20)


To continue this analysis, we must know something about the flow field
around the tube. We need not guess the flow topology, for experimental and
theoretical studies have been carried out 27 indicating there is a thin region
(called the boundary !ayer) around the tube where the velocity changes
rapidly from the free stream value to zero. At high Reynolds numbers, the
boundary 1ayer thickness is small and the pressure difference across it is
negligible. Under these conditions, the pressure at point 3 is very nearly
equal to the pressure at the static hole at the surface of the Pitot-tube. We
take the streamline from points 1 to 3 to be outside the boundary !ayer; thus,
(8.5-21)
and the pressure at the static hole is given by
Pa = PI + pg(zi - Za) (8.5-22)

Substituting Eq. 8.5-19 into Eq. 8.5-22 and solving for the velocity, we get
//7////////ff//,ij
vi = J1.[(P2 - Pa) + pg(z2 - za)J . (8.5-23) Fig. 8.5-8. S
p
As in the other systems studied in this section, the analysis is not exact and attack the problem. Both the mon
Pitot-static tu bes must be calibrated in terms of the equation, would seem to be suitable tools fo
equation and lea ve the application o
vi= e J2(p2; Pa) (8.5-24) Along any streamline we may w

p + !pv
The gravitation terms have been dropped because they are always negligible.
If a Pitot-static tube is properly designed, thf" coefficient e can be very close provided viscous effects are negligi
to unity. The implication is not that the analysis is exact; it simply indica tes the thickness of the jet at the weir
that severa1 errors have cancelled. the weir will be small. U nder th
specified by applying Eq. 8.5-25 at
t Remember that Bernoulli's equation neglects viscous effects.
27. H. Schlichting, Boundary Layer Theory, 4th ed. (New York: McGraw-Hill Book
Company, Inc., 1955).
ic Balances-Viscous Effects Chap. 8 Sec. 8.5 Flow Rate Measurement 339

tion of Eq. 7.4-17 to the streamline Sharp-crested weir

(8.5-18) Flow rates in open channels may be measured by a weir, an obstruction in


the channel over which the fluid must ftow. There are severa! types of com-
monly used weirs, but we will consider only the sharp-crested weir shown in
(8.5-19) Fig. 8.5-8. It represents a rather complicated ftow problem, and in the ab-
sence of experience and knowledge, it is not at all obvious how we should

Pa + tpvi + pgza (8.5-20)


know something about the ftow field
flow topology, for experimental and
out27 indicating there is a thin region
the tube where the velocity changes
zero. At high Reynolds numbers, the
the pressure difference across it is
pressure at point 3 is very nearly
at the surface of the Pitot-tube. We
to be outside the boundary !ayer; thus,
(8.5-21)

(8.5-22)

and solving for the velocity, we get

Fig. 8.5-8. Sharp-crested weir.


(8.5-23)

section, the analysis is not exact and attack the problem. Both the momentum and mechanical energy balance
terms of the equation, would seem to be suitable tools for the analysis, but we will use Bernoulli's
equation and Ieave the application ofthe macroscopic balances asan exercise.
(8.5-24) Along any streamline we may write

p + tpv 2 + pgz =c. (8.5-25)


because they are always negligible.
thP coeffi.cient C can be very close provided viscous effects are negligible. If the channel is deep compared to
analysis is exact; it simply indicates the thickness of the jet at the weir (H ';J> h), the velocity far upstream from
the weir will be small. Under these conditions, the constant c. may be
specified by applying Eq. 8.5-25 at sorne point far upstream from the weir.
viscous effects.
4th ed. (New York : McGraw-Hill Book
C, = p + pgz (8.5-26)
340 Macroscopic Balances-Viscous Effects Chap. 8 Problems

Since the velocity in the z-direction is negligible there, the pressure is just the While experiments must be carried 01
hydrostatic pressure, this analysis does have value in that i
rate on (H - L). In Chap. 9, we w
P =Po +
pg(H - z) (8.5-27)
show how practica! flow rate measuri
and we see that every streamline has the same constant given by theory, thus providing a more reliab
The results presented in this sectio
C, =Po+ pgH (8.5-28)
on fluid metering 28- 30 to obtain both
and a thorough discussion of the m
We may now write Eq. 8.5-25 as
tions.
p + l pv + pgz =Po +
2
pgH (8.5-29)
PRO
At the crest of the weir the fluid pressure is unknown; however, the pressure
at both the top and bottom of the jet is atmospheric, and the pressure within 8-1. To determine the pressure drop-fl<
the liquid should not be too far removed from this value. Applying Eq. bundle used in heat transfer e
8.5-29 to the streamlines at the crest of the weir and setting p = p 0 , we find configurations similar to that
shown in Fig. 8-1. The tubes are
v = ..}2g(H- z) (8.5-30) arranged in a triangular pattern
with the mean flow directed along
The volumetric flow rate over the weir is given by the diagonal of the square. De-
z ~ L+h
scribe how you would interpret the

Q= b f
z= L
v i dz (8.5-31)
experimental results obtained i
such a system-i.e., how would
you define the friction factor, what
parameters would f be a function
where bis the width ofthe weir. While Eq. 8.5-30 may be a satisfactory expres- of, etc.? Neglect wall effects and
sion for the magnitude of the velocity vector, it does not provide us with the assume the tubes are infinitel)i
x-component of the velocity vector. If we make the approximation long.
(8.5-32) 8-2. Derive an integral expression for
tube shown in Fig. 8-2. Express
Eq. 8.5-30 may be substituted into Eq. 8.5-31 and the integration carried out
to yield
Q = fb..}2g{[H- L]3 i 2 - [H- (L + h)] 3i 2} (8.5-33)
---
EJ
In practice, the height of the weir L is known, and the depth far upstream H
is measured with a depth gauge ; however, the position of the free surface at Q

the weir remains asan unknown. Traditionally, we assume -----.:


A1
H- (L + h) ~ H- L (8.5-34)
Flg. S.l. Flow iJ
and introduce a discharge coefficient to obtain
28. " Flowmeter Computational Handt
(8.5-35) lication, New York, 1961.
29. " Fluid Meters : Their Theory and
We expect the discharge coefficient to be less than unity, and experiments New York, 1959.
indicate that 30. H. W. King and E. F. Brater, Ham
(8.5-36) Book Company, Inc., 1963).
Balances-Viscous Effects Chap. 8 Problems 341

negligible there, the pressure is just the While experiments must be carried out to determine the discharge coefficient,
this analysis does ha ve value in that it predicts the correct dependence of flow
pg(H- z) (8.5-27) rate on (H- L). In Chap. 9, we will analyze the broad-crested weir, and
show how practica! flow rate measuring devices can be altered to fit available
same constant given by theory, thus providing a more reliable method of measurement.
The results presented in this section are brief; we must refer to handbooks
(8.5-28)
on fluid metering 28 - 30 to obtain both accurate values for discharge coefficients
and a thorough discussion of the many metering devices and their applica-
tions.
=Po+ pgH (8.5-29)
PROBLEMS
is unknown; however, the pressure
atmospheric, and the pressure within
8-1. To determine the pressure drop-flow rate relationship for flow through a tu be
from this value. Applying Eq.
bundle used in heat transfer equipment, experiments are performed on
the weir and setting p = p 0 , we find configurations similar to that
shown in Fig. 8-1. The tubes are Tube diometer is O
(8.5-30) arranged in a triangular pattern /'/'##/$#$!'#/#/&
with the mean flow directed along
the diagonal of the square. De- o o o o
scribe how you would interpret the Mean
experimental results obtained in - - -
o o o
(8.5-31) flow
such a system-i.e., how would o o o o
you define the friction factor, what
parameters would f be a function o o o o o
Eq. 8.5-30 may be a satisfactory expres-
of, etc.? Neglect wall effects and
vector, it does not provide us with the
assume the tubes are infinitely
//7////////////////////////////
we make the approximation long. Flg. 8-1. Flow through a tube bundle.

(8.5-32) 8-2. Derive an integral expression for the pressure drop in the expanding circular
tu be shown in Fig. 8-2. Express !lp as a function of Q, A1 , A2, and f The
. 8.5-31 and the integration carried out
r
- [H- (L + h)]al2} (8.5-33)

known, and the depth far upstream H


the position of the free surface at
iiuium1a11 , we assume
1
1
1
1
1
1
1
,
1

L,
(8.5-34)
A~~-----------------------'~,,~
Az
Flg. 8-l. Flow in an expanding tube.

_ L)312 (8.5-35) 28. "Flowmeter Computational Handbook," Am. Soc. Mech. Engrs. Research Pub-
/ication, New York, 1961.
be less than unity, and experiments 29. "Fluid Meters: Their Theory and Application," Am. Soc. Mech. Engrs. Paper,
New York, 1959.
30. H. W. King andE. F. Brater, Handbook of Hydraulics (New York: McGraw-Hill
(8.5-36) Book Company, lnc., 1963).
Problems
342 Macroscopic Balances-Viscous Effects Chap. 8

cross-sectional area is given by


A = A1 + (A 2 - A1 )(i)
and yo u are to assume that f is constant and the flow is turbulent.
8-3. Water is pumped through a 1-in. diameter pipe 100ft long at arate of 120
in. 3 /sec. Neglecting entrance effects, calculate the pressure drop for a relative
roughness ej D of 0.002.
8-4. Calculate the pressure drop for the conditions in Prob. 8-3 if the pipe is
replaced by a square duct, 1 in. on a side. Pipe diometer =0 0
Surge tonk diometer =fJ
8-5. Crop-dusting of the artichoke fields around Castroville, California, is tra-
ditionally done by pilots flying at an altitude of 17 ft. If the particles of
insecticide are approximately spherical with an average diameter of 10- 2 cm,
anda density of 127lbm/ft, how long does it take them to fall to the ground?
If the nearest population center is 1 mi from the fields, what is the maximum Flg. 8-7. Use of a surge t
wind velocity (directed from the field toward the town) that can be tolerated
before the dusting operation must be stopped? an approximate solution for the
8-6. The flow rate through a flow nozzle was given in Sec. 8.5 by t~nk h'(t) may be obtained. Det
maximum damping, and comme
2(p - P2) Hint: Assume a solution for h'(t
h'(t)
where the discharge coefficient Cd had to be determined experimentally. and make use of the approximat1
Repeat the derivation retaining the viscous dissipation term in the analysis.
Explain what experiments you would perform, and how you would correlate
the data to obtain a satisfactory method of predicting flow rates by a flow
nozzle. The assumption of flat velocity pro files is a reasonably good one and 8-8. Apply the momentum and mecha
should be incorporated in your analysis. for the flow rate over the shar
8-7. The pump illustrated in Fig. 8-7 has an output given by fully list the assumptions that mu
Q = Q0 + aQ sin wt
where Q 0 is the time averaged flow rate, and aQ represents the amplitude of
the volumetric flow rate variations . The frequency of these variations is w. 8-9. Determine the water fiow rate fa
If the siphon shown in Fig. 8-9. Th
aQ ~ Qo inlet to the siphon is placed 10 1
the friction factor may be taken as a constan t. In the absence of a surge tank, below the water leve! and the ow
the fluctuations occurring at the process are identical to those at the pump let is even with the bottom of th
and the process is therefore difficult to control. You are asked to design a tank. The pipe diameter is 1 ft, th
surge tank which will reduce the amplitude of the fluctuations by a factor of totallength of the piping is 450 f
10. Set up the problem in a general fashion, locating a surge tank of diameter and the relative roughness is 0.0~
D1 at a distance rxL (O : : :; rx ~ 1) from the pum p. Neglect minor losses and The fittings in the lineare standar
inertial effects. An exact solution is difficult to obtain; however, if the depth in 90 elbows. Assume that the levf
the surge tank is represented by a steady term and a time-dependent term of the water in the tank i
h = h0 + h'(t) constan t.
where
h'(t) ~ h0
Balances-Viscous Effects Chap. 8 Problems 343

and the ftow is turbulent.


pipe 100ft long at a rate of 120
-a)L.---..ll r
calculate the pressure drop for a relative

conditions in Prob. 8-3 if the pipe is


\\ ;f'0
si de. Pipe diometer = 0 0
around Castroville, California, is tra- Surge tonk diometer =0 1
altitude of 17 ft. If the particles of
with an average diameter of w- 2 cm,
does it take them to fall to the ground?
from the fields, what is the maximum
toward the town) that can be tolerated Flg. 8-7. Use of a surge tank to damp pump functions.
stopped?
an approximate solution for the amplitude of the fluctuations in the surge
was given in Sec. 8.5 by tank h'(t) may be obtained. Determine the value of oc that will provide the
2(p - P2) maximum damping, and comment on the effect of increasing the ratio D 1/ D 0
Hint: Assume a solution for h'(t) of the form
p[1 - (A 2/A 1) 2]
h'(t) =a,. sin (wt - 8)
had to be determined experimentally.
viscous dissipation term in the analysis. and make use of the approximation
perform, and how you would correlate
of predicting flow rates by a ftow
y, = y-;0 (t + ~ h'(t))
ho
profiles is a reasonably good one and
8-8. Apply the momentum and mechanical energy balances to derive an expression
for the flow rate over the sharp-crested weir discussed in Sec. 8.5. Care-
fully list the assumptions that must be made to obtain a solution in each case.
+ aQ sin wt
and aQ represents the amplitude of
The frequency of these variations is w.
8-9. Determine the water flow rate for
the siphon shown in Fig. 8-9. The
inlet to the siphon is placed 1O ft
In the absence of a surge tan k, below the water level and the out-
are identical to those at the pump let is even with the bottom of the
to control. You are asked to design a tank. The pipe diameter is 1 ft, the
of the ftuctuations by a factor ~r totallength of the piping is 450 ft,
locating a surge tank of diameter and the relative roughness is 0.02.
the pump. Neglect minor losses and The fittings in the lineare standard
to obtain; however, if the depth in 90 elbows. Assume that the level
term and a time-dependent term of the water in the tank is
constan t.

Flg. 8-9. Flow in a siphon.


Macroscopic Balances-Viscous Effects Chap. 8 Problems

8-10. Explain how you might estimate the totalloss owing to the presencc of a flow
nozzle in a pipeline. Hint: Apply both the momentum and mechanical energy
balances to a suitable control volume.
8-11. Determine the cros~-sectional area

r
of the jet illustrated in Fig. 8-11 3
as a function of z. The area of the p = 68.5 lbm/ft
rounded orifice is A 0 , and you are J.L = 1.2 centipoise
to assurne that the velocity profile h
everywhere in the jet is flat. Treat

Ll.t:z::z::~:::z::~____z =o
Sudden contr
the flow as quasi-steady. Neglect 6in. I.D. to
viscous effects and sol ve the prob-
lem using both Bernoulli's equa-
tion and the momentum balance.

8-12. Water at 70F flows through a


Venturi meter as shown in Fig.
8-12. Cavitation occurs when the
local pressure falls below the vapor
pressure of the liquid, allowing
vapor bubbles to form. The vapor Flg. 8-11. Flow through a rounded orifice 5'
in a tank. .L_
Seo wot
pressure of water at 70F is 0.36
psia. Calculate the maximum flow
rate that this meter can handle without cavitation. Flg. 8-13. Flo

"1" "2"
1
: 1
IJ,II:J.IIr.t.liJ.II~11r.t.IU,II.,I.J : /1 (/ti(/! 1111111.1..
~~~wwuu~
'J .

1
2
A1 =0.10 ft !
1 1
Flow
! i A2=0.025 ft
2
i _.......:.r:, _ _
1 1/T~--

~~~~ "~mnmm~~
Fig. 8-12. Flow through a Venturi meter.

8-13. Figure 8-13 shows a pipeline system for pumping sea water into a reservoir.
The Venturi meter has a discharge coefficient of 0.96 anda throat area one-
half that of the pipe. The pipes are 4-in. and 6-in. I.D. commercial steel, and
the manometer fluid is mercury. Determine the horsepower requirement of
the pump and the volumetric flow rate. The density of sea water may be
taken as 68.5 lbm/ft 3 and the viscosity as 1.2 centipoise.
8-14. A i-in. plastic sphere (p = 38 lbm/ft 3) is released from the bottom of a deep
lake. What is the terminal velocity?
8-15. Calculate the horsepower delivered by the pump to maintain a steady flow
ofwater in the system shown in Fig. 8-15. All piping is 6-in. I.D. commercia1 Flg. 8-15. Flo
Balances-Viscous Effecu Chap. 8 Problems
345
totalloss owing to the presencc of a flow
the momentum and mechanical energy

r
h
JL
p =68 .5 1bm/fl 3
= 1.2 centipoise Pump

L~p::z:ZJ-z=o Sudden controction from


6in. I.D. lo 4in . I.D.

Flg. 8-11. Flow through a rounded orifice


in a tank. 5'
.L...
Seo water

Flg. 8-13. Flow in a pipeline.

for pumping sea water into a reservoir.


eoeffic.ient of 0.96 and a throat area one-
and 6-in. I.D. commercial steel, and
the horsepower requirement of
rate. The density of sea water may be
as 1.2 centipoise.

is released from the bottom of a deep

the pump to maintain a steady flow


5. AII piping is 6-in . I.D. commercial Flg. 8-15. Flow in a pipeline.
Macroscopic Balances-Viscous Effects Chap. 8 Problems
346

steel, and all fittings are 90 standard elbows. Assume a value of 0.7 for the 8-20. Calculate the flow rate through tht
contraction coefficient of the 4-in. diameter orifice in the si de of the tan k. submerged orifice in Fig. 8-20
The orifice diameter is 2 in. Esti
8-16. Electrical transmission towers are placed at 1000-ft intervals, and i-in. diam- mate the discharge coefficient.
eter cables are strung between them. Determine the drag force on a single
cable between two towers if the wind velocity is 60 mph.
8-21. Derive an expression for the vol
8-17. When 15 ft 3 /sec ofwater flow in a 12-in. I.D. pipeline, 85 hp are lost in viscous umetric flow rate in a Ventu
dissipation for every 1000 ft of pipe. For these conditions, calcula te the head meter (see Sec. 8.5), assuming th
loss, friction factor, and shear stress at the pipe wall. velocity profile is always paraboli
Compare your result with th
8-18. Calculate the flow rate of water from the tank illustrated in Fig. 8-18. experimental values of the dis
charge coefficient shown in Fig
8.5-2.

Water
l
lOOft
1 in. 1.0. smooth tubing

k----------250ft---------~!
Fig. 8-18

8-19. Calculate the smallest flow rate required to keep the pipeline shown in Fig.
8-19 running full. The pipe is 1-in. I.D. commercial steel. Neglect minor
losses and keep in mind the phenomenon of cavitation.

r-30'----j
t

t
50'
t l 100'

Water at 70 F
LL \
)._
j
Pump

Flg. 8-19. Flow in a pipeline.


ic Balances-Viscous Effects Chap. 8 Problems 347

elbows. Assume a value of 0.7 for the 8-20. Calcula te the flow rate through the
orfice in the side ofthe tank. submerged orfice in Fig. 8-20. Air ot 4 psig Atm. press.
The orfice diameter is 2 in. Est-
at 1000-ft intervals, and i-in. diam- mate the discharge coefficient.
Determine the drag force on a single
velocity is 60 mph.
5'
8-21. Derive an expression for the vol-
.I.D. pipeline, 85 hp are lost in viscous
For these conditions, calculate the head
at the pipe wall.
umetric flow rate in a Venturi
meter (see Sec. 8.5), assuming the
velocity pro file is always parabolic.
j__
the tank illustrated in Fig. 8-18. Compare your result with the
experimental values of the dis-
charge coefficient shown in Fig. Fig. 8-20. Flow through a subrnerged
8.5-2. orifice.

1 in. /.D. smooth tubing

250ft _ _ _ _..,

8-18

to keep the pipeline shown in Fig.


I.D. commercial steel. Neglect minor
of cavitation.

r-30'4
t
50'
t l lOO'
Sec. 9.1 Uniform Flow

the curvatu re of the free surfacet is sm


and rapidly varied ftows, where the cu
In the study of ftow in closed con~
fiow rate, pressure drop, and geometry
will seek a relationship between the
geometry [Q, h(x), b(x), 'l)(x)]. We sha
Open Channel Flow 9 terms of the macroscopic mass, mome
and then go on to consider the flow ir
This willlead us to equations defining
shall analyze rapidly varied ftows, suc

9.1 Uniform Flow

Consider the ftow in a rectangula


sluice gate, such as that illustrated in
gate, the ftow will be nonuniform ar

The study of open channel flow is primarily directed toward applications


in irrigation, ftood control, and water suppty. The analysis of ftow in open
channels is more complex than the analysis of flow in closed conduits for two
reasons : the existen ce of a free surface complicates the analysis; and the
actual channels encountered in practice are comparatively complex. A large

_j~___ -----
portian of the ftow in closed conduits is restricted to pipe flow, where the
diameter and relative roughness describe the geometry fairly well. However,
in dealing with flow in open channels we may encounter anything from a
carefully constructed aquaduct to a grass-lined irrigation ditch. The flow
characteristics of the former are fairly well known; however, those of the Fig. 9.11. Formati
latter must be studied experimentally for every individual case if accurate
results are desired. The study of open channel ftow could easily be the subject distance downstream, a balance .bet
of an entire course, and in a single chapter we can only hope to comment on will be achieved and the flow will be
the most important aspects. a representative rectangular channel
These ftows may be classified as either steady or unsteady, uniform or indicate a rather curious situation. B
nonuniform. With the exception of the solitary shallow-water wave studied wetted surface is the only force opp
in Sec. 9.3, al! the ftows investigated in this chapter will be steady. A uniform leads us to expect the maximum velo
ftow is one for which the fluid depth h above the channel bed is constant.
Nonuniform flows can be further classified into gradually varied flows, where t The term "free surface" denotes an
that the tangential stress is zero and the no
348
Sec. 9.1 Uniform Flow 349

the curvature of the free surfacet is small compared to the depth of the fluid,
and rapidly varied flows, where the curvature is comparable to the fluid depth.
In the study of flow in closed conduits, we sought equations relating the
flow rate, pressure drop, and geometry (Q, !l.p, L, r0 ), while in this chapter we
will seek a relationship between the flow rate, fluid depth, and channel
geometry [Q, h(x), b(x), r(x)]. We shall first analyze a simple uniform flow in
9 terms of the macroscopic mass, momentum, and mechanical energy balances,
and then go on to consider the flow in a channel with a changing geometry.
This willlead us to equations defining gradually varied flows, after which we
shall analyze rapidly varied flows, such as the hydraulic jump.

9.1 Uniform Flow

Consider the flow in a rectangular channel formed downstream from a


sluice gate, such as that illustrated in Fig. 9.1-1. Just downstream from the
gate, the flow will be nonuniform and rapidly varying; ho~ever, at sorne

y directed toward applications


supply. The analysis of flow in open
of flow in closed conduits for two
complicates the analysis; and the
are comparatively complex. A large
is restricted to pipe flow, where the
the geometry fairly well. However~
we may encounter anything from a
lined irrigation ditch. The flow
well known; however, those of the
_j)___ -----------------------
Fig. 9.1-1. Formation of a uniform flow.
for every individual case if accurate
annel flow could easily be the subject distance downstream, a balance .between gravitational and viscous forces
r we can only hope to comment on will be achieved and the flow will become uniform. The velocity profile for
a representative rectangular channel is shown in Fig. 9.1-2, and the curves
ther steady or unsteady, uniform or indicate a rather curious situation. Because the drag force occurring at the
solitary shallow-water wave studied wetted surface is the only force opposing the gravitational force, intuition
chapter will be steady. A uniform leads us to expect the maximum velocity to occur at the centerline on the free
abo ve the channel bed is constan t.
into gradually varied flows, where t The term "free surface" denotes an air-water interface, and has the characteristic
that the tangential stress is zero and the normal stress is - np 0
Open Channel Flow Chap. 9 Sec. 9.1 Uniform Flow
350
M ximurri velocity Time-averaging of all the equation.
because the open channel fiows of p
turbulent. The general macroscopic 1

:J pv dV
?'"a (t)
+ J pv(v - w)
A, (t)
z
h can be dotted with i and applied to t

Because the fiow is uniform,


(v!)2
Unes of constont velocity ( Vx)
and Eq. 9.1-6 reduces toa balance o
Fl. 9.1-l. Velocity profiles in a rectangular channel.

O = pgzAL + f [-
d
surface; however, the maximum velocity is located below the free surface. This
anomaly appears to result from the secondary flow (indicated in Fig. 9.1-2), Assuming that the shear stress at th
which carries the low velocity (vz) fluid upward along the sides and then out should be for uniform fiow), Eq. 9.1
over the free surface.

Mass balance
where S is 'the wetted perimeter, b -1
that
Starting with the general macroscopic mass balance,
(pg)l

f!..
dt
f +fp
?'"4 (t)
dV p(v - w) n dA
A,<t>
.
= O (9.1-1) Defining an average wall shear stres~

l
'~'o=--
S,
and simplifying it for a fixed control volume and incompressible fiow, we
obtain we obtain

J vndA =O (9.1-2)
Following the procedure used in tC
O= pg sin

A,
arrange this equation in dimensionle
Application of Eq. 9.1-2 to the control volume illustrated in Fig. 9.1-1 yields Dimensionless
wall shear stress
the obvious result
(9.1-3)

Since A2 = A1 = bh, we obtain


t The use of the net stress vector is a
(9.1-4) inasmuch as t 1:, is zero at a free surface.
Open Channel Flow Chap. 9 Sec. 9.1 Uniform Flow 351

Time-averaging of all the equations used in this chapter is understood,


because the open channel flows of practica! importance are almost always
turbulent. The general macroscopic momentum balancet

z
:J pv dV
7" (t)
+ J pv(v- w) n dA
A. (t)
= J pg dV
7" (t)
J
+ t~>
.9/ (t)
dA (9.1-5)

can be dotted with i and applied to the fixed control volume to yield

p(v! )2 A - p(v! )1 A = pg.,AL f


+ t~> i dA
.9/
(9.1-6)

Because the flow is uniform,

(v.~:.. ) and Eq. 9.1-6 reduces to a balance of surface forces and body forces,
(9.1-7)
a rectangular channel.

O = pg.,AL + f [-i npg + i (n T)] dA (9.1-8)


located below the free surface. This .9/
(indicated in Fig. 9.1-2), Assuming that the shear stress at the solid walls is independent of x (as it
along the sides and then out should be for uniform flow), Eq. 9.1-8 becomes

(9.1-9)
S
where S is the wetted perimeter, b + 2h. Once again uniform flow requires
that
(pg)] = (pg)2 (9.1-10)
(9.1-1) Defining an average wall shear stress as

(9.1-11)
and incompressible flow, we
we obtain
O= pg sin O AL- -r0 SL (9.1-12)
=0 (9.1-2)
Following the procedure used in the analysis of closed conduits flow, we
arrange this equation in dimensionless form to obtain
illustrated in Fig. 9.1-1 yields
Dimensionless Dimensionless
wall shear stress body force

(9.1-3)
(9.1-13)

(9.1-4) t The use of the net stress vector is a convenience in analyzing open channel ftow
inasmuch as t(:) is zero at a free surface.
Open Channel Flow Chap. 9 Sec. 9.1 Uniform Flow

where the hydraulic radius is given by Flow rate-channel depth calculati


cross-sectional area bh uniform flow
R~ = ----:---:---
wetted perimeter S
Severa! emprica! expressions exist for the friction factor, f The most If the geometry and fluid depth fo
popular is the work of Manning, 1 which we may write in the form of the flow rate is straightforward. If
and solve for (v:z), we get
n2
f = 116 R~13 (9.1-14)

The coefficient n is a measure of the roughness, and values for severa! types
of surfaces are Iisted in Table 9.1-1. Since f is dimensionless, we must be From the channel geometry and fluil
volumetric flow rate. For a rectangul
Table 9.1-1
T ABLE OF ROUGHNESS COEFFICIENTS na
Q= bh
8g sin() R:'
116n 2
Type of Channel n, ft' /6
If the volumetric flow rate and channe
Artificial Channels of Uniform Cross Section
is to be determined, we use a trial-
Sirles and bottorn Iined with well-planed tirnber evenly laid . . . . . . . . . . . . . . . . . . . . 0.009 channel; we multiply Eq. 9.1-15 by bJ
Neat cernen! plaster, srnoothest pipes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0.010
Cernen! plaster (3 cernen! to 1 sand), srnooth iron pipes . . . . . . . . . . . . . . . . . . . . . . 0.011 R~=
Unplaned tirnber evenly laid, ordinary iron pipes .. . .... .. . .. ........ ... .. .... 0.012
Ashlar rnasonry, best brickwork, well-laid sewer pipeb . . . . . . . . . . . . . . . . . . . . . . . . 0.013 and rearrange to obtain
Average brickwork, foul planks, foul iron pipes, ordinary sewer pipes after average
uneven settlernent and average fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0.015
Good rubble rnasonry, concrete laid in rough forrns, poor brickwork, heavily h=![(2h+b
incrusted iron pipes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0.017
Channels Subject to Nonuniforrnity of Cross Section
If we assume a value of h, we can th
value of h and test the validity of th1
1

Excellent clean canals in firrn grave!, of fairly uniforrn section; rough rubble, yield close agreement between the a
"drying paving" . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0.020 fluid depth is determined. More con
Ordinary earth canals and rivers in good order, free frorn large stones and heavy channel, Iead to a more tedious tria!
weeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0.025
Canals and rivers with rnany stones and weeds . . . . . . . . . . . . . . . . . . . . . . . . . . . 0.03-0.04 is straightforward.

a Data fro m L. Marks and T. Baumeister, eds. , Marks' Mechanica/ Englneers' Handbook . 6th ed. (New York :
McGraw-Hill Book Company.Inc. 1958). pp. 3- 81. A more tho rough list of values.is given in Ref. 2. Mechanical energy balance
b This last value should be used for the previous categories, if doubt exists asto the exccllencc of construction
and the maintenance free from slime, rust, or other growths and deposits.

If the flow is incompressible and


careful to remember that n has units of ftl/ 6 Note that the value off given
may use the results of Prob. 7-2 to 1
by Eq. 9.1-14 is independent of the Reynolds number, indicating that it is
only valid for the " rough-pipe" region of flow where inertial effects pre-
dominate. A detailed discussion of the drag force is given by Chow.2 :t fr.w(!pv~ dV +j
A
l. R. Manning, "On the Flow of Water in Open Channels and Pipes," Trans. Inst. Civ.
Engrs. Jreland, 1891,20 :161.
2. V. T. Chow, Open Channel Hy drau/ics (New York : McGraw-Hill Book Cornpany, = f V ttn) dA -
Inc., 1959), Chap. l.
A.,(t)
Open Channel Flow Chap. 9 Sec. 9.1 Uniform Flow 353

Flow rate-channel depth calculations for


bh uniform flow
S
for the friction factor, f The most If the geometry and fluid depth for a channel are given, the determination
we may write in the form ofthe flow rate is straightforward. Ifwe substitute Eq. 9.1-14 into Eq. 9.1-13
2 and solve for (v.,), we get
16~
R~ 13 (9.1-14) 8g sin eR 11
413

<V"') -- 116n2
(9. 1-15)
uuJ~nn.c:::~s. and values for severa! types
Since f is dimensionless, we must be From the channel geometry and fluid depth, we may readily compute the
volumetric flow rate. For a rectangular channel,

Q = bh
8g sin eR:' 3
rectangular channel (9.1-16)
n, ft 1 i 6 116n 2
If the volumetric flow rate and channel geometry are given and the fluid depth
is to be determined, we use a trial-and-error solution. For a rectangular
0.009 channel, we multiply Eq. 9.1-15 by bh, express the hydraulic radius as
0.010
0.011

. b . 0.012
pipe .. .. . ..... .. . . ... . . .. . . . 0.013
ordinary sewer pipes after average
and rearrange to obtain
... .
forms, poor brickwork, heavily
0.015 1[
h = - (2h
b
+ b) (116Q2n2)3/4]2/5
8g sin e
(9.1-17)
0.017
If we assume a value of h, we can then use Eq. 9.1-17 to calculate a second
value of h and test 'the validity of the assumption. A few trials will quickly
uniform section; rough rubble, yield close agreement between the assumed and calculated values, and the
... o 0.020 fluid depth is determined. More complex geometries, such as a trapezoidal
, free from large stones and heavy
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0.025 channel, Iead to a more tedious trial-and-error calculation, but the method
. . . . . . . . . . . . . . . . . . . . . . . . . . . 0.03-0.04 is straightforward.

Mechan/cal Englneers Handbook . 6th ed. (New York :


thorough list of values.is gi ven in Ref. 2.
if doubt exists asto the excellence of construction Mechanical energy balance
and deposits.

ft 116 . Note that the value off given If the flow is incompressible and there are no solid moving surfaces, we
number, indicating that it is may use the results of Prob. 7-2 to writc the mechanical energy balance as
of flow where inertial effects pre-
drag force is given by Chow. 2 :t J(ipv~
r .(t)
dV + J(!pv )(v -
.A:,(t)
2
w) n dA
Open Channels and Pipes," Trans. Inst. Civ.
(9.1-18)
(New York : McGraw-Hill Book Company,
= Jv t7n) dA - Jpcf>v n dA - ,
A , (t)
354 Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow

For the case under consideration, Eq. 9.1-18 quickly reduces to


9~2 Gradually Varied Flow
J(tpv )v n dA = - J
2
PuV n dA - J pfv n dA - Ev (9.1-19)
A, A, In this section, the macroscopic ba
Here we have used the arguments presented in Sec. 8.1 to reduce the stress on illustrated in Fig. 9.2-1. The width an<
the entrance and exit to a simple pressure term, -np 9 Writing the gravita- to a variable fluid depth h(x).t The v
ticnal potential as shown in Fig. 9.2-1 areexaggerated, bec
cp = g(1](X) Z COS (:1] + (9.1-20)
and expressing the gauge pressure as
Po = pg[(h - z) cos (:1] (9.1-21)
we may combine the pressure and gravitational terms in Eq. 9.1-19 to obtain

-J PuV n dA - J pfv n dA = - pg J(17 + h cos (:l)v n dA (9.1-22)


A, A, A, dr
ton (}=- dx
Application of Eqs. 9.1-19 and 9.1-22 to the control volume indicated in
Fig. 9.1-1 gives
tp(v!)2A2 - lp(v!)IAI
= -pg[(1J 2 + h 2 cos (:i)(v.,) 2 A2 - (17 1 +h cos (:i)(v.,) 1 A1] (9.1-23)

=uLMUUm~
1

Inasmuch as the flow is uniform,


(v!)2A 2 = (v!)A 1
(puv.,)2A2 = (p"v.,)A 1 b(x)
(zv.,) 2 A 2 = (zv.,) 1 A 1
(v.,)2 A 2 = (v.,) 1 A 1
The mechanical energy balance reduces to
mnmmmmmlnm,
O= pg[17 lx=x 1
- 1J lx=x,] (v.,)A - Ev (9.1-24)
Dividing by pgQ and noting that
(9.1-25)
1J '"'="' - 1J '"'="' = L sin (:1
we can rewrite Eq. 9.1-24 as
L sin (:1 = h1 (9.1-26) ~rrTm77777,i7771i77777i~
Thus, the viscous dissipation is just equal to the decrease in potentia/ energy Fig. 9.2-1. Grad
for uniform flow. The head loss can be expressed in more familiar terms by
substituting sin (:1 from Eq. 9.1-13 to obtain only if the geometry is changing slowl)
channel must be small compared to the
(9.1-27) of the channel bed must be small comJ
fore treat the flow as one-dimensional, 1
However, the physical significance of the viscous dissipation is best expressed
t The phrase, "variable fluid depth" is th
by Eq. 9.1-26.
balances are in order.
Open Channel Flow Chap. 9
Sec. 9.2 Gradually Varied Flow 355
9.1-18 quickly reduces to

n dA- Jpcpv n dA- Ev (9.1-19)


9-.2 Gradually Varied Flow

In this section, the macroscopic balances will be applied to the channel


nted in Sec. 8.1 to reduce the stress on illustrated in Fig. 9.2-1. The width and slope are allowed to vary giving rise
term, -np 9 Writing the gravita- to a variable fluid depth h(x).t The variations in channel slope and width
shown in Fig. 9.2-1 are exaggerated, beca use the analysis presented here is valid
+ z cosO] (9.1-20)

- z) cosO] (9.1-21)

-
tational terms in Eq. 9.1-19 to obtain

-pg J(r + h cos O)v n dA (9.1-22)


A,
to the control volume indicated in

(v~) 1 A 1
(pgv.,) 1 A 1
b(x) Plan view
(zv.,) 1A1
(v.,)A 1

(9.1-24)

(9.1-25)
Cross section

(9.1-26)
wm7?7777J7777J"7777i77/~
to the decrease in potentia/ energy Fig. 9.2-1. Gradually varied flow.
expressed in more familiar terms by
only if the geometry is changing slowly-i.e., the curvature of the walls of the
channel must be small compared to the width ofthe channel, and the curvature
(9.1-27) of the channel bed must be small compared to the fluid depth. We can there-
fore treat the flow as one-dimensional, even though the streamlines are curved.
viscous dissipation is best expressed
t The phrase, "variable fluid depth" is the clue indicating that differential-macroscopic
balances are in order.
Sec. 9.2 Gradually Varied Flow
356 Open Channel Flow Chap. 9

and Eq. 9.2-1 reduces to


(v.,)bh !"'+t."' -
(al Dividing by D..x and taking the limit D.
h(x) scopic mass balance,
d
Side view dx ((v.,

Momentum balance

For steady flow, we may dot Eq. 9

Free surfoce al x Jpv.,v n dA = f_ p


A, 1"

Starting with the momentum flux term


Perimeler al x tial control volume to obtain

(b)
Perimeler al x+llx
Jpv.,v n dA = p(v!
A,
The body force term yields

Flg. 9.2-2. Control volume. f


-r
pg.,dV =

The area integral of the stress vector


A more detailed version of the control volume to be used is shown in Fig.
9.2-2. It is formed by two parallel surfaces at x and x + D..x perpendicular to
the x-coordinate. f
~
i t<:> dA =- f
~
i 1

Directing our attention to the pressur


Mass balance
l. Pu = O, on the free surface;
Application of Eq. 9.1-2 to this control volume yields 2. i . n = O, on the bottom of the

f V
A,(z)
n dA = f v n dA
A,(z)
1., f
+
A,(z)
v n dA I.,+A., = 0 (9.2-1)
3. i. n = -l db , on the sides of
dx
Under these conditions, we may writt

Ji DPu dA = f Pu dA LA., - j
where the cross-sectional ar~a of flow, A.(x), may be expressed as
A.(x) = bh (9.2-2)
~ A,(z) A

Note that v n must be zero at the free surface for the flow to be steady.
Assuming that the flow is one-dimensional, we may write

VD = v., At the entrance or exit of


(9.2-3)
- (~
the control volurne
Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow 357

and Eq. 9.2-1 reduces to


(v,.)bh !"'+t.'" - (v,.)bh j,. = O (9.2-4)
1 Dividing by t:u and taking the limit ~x ---.. O, we find the differential macro-
1
1
1 scopic mass balance,
1
1 d
1
1 Side view - ((v,.)bh) = O (9.2-5)
1 dx
Momentum balance

For steady flow, we may dot Eq. 9.1-5 with i to obtain

Jpv,.v n dA = f_ pg,. dV +Ji t7n> dA (9.2-6)


A, -r .!#

Starting with the momentum flux term, we apply this equation to the differen-
tial control volume to obtain

J
A,
pv,.v n dA = p(v!)bh lzHz - p(v!)bh ,. (9.2-7)

The body force term yields

J
-r
pg,. dV = pg,.bh ~x (9.2-8)

volume to be used is shown in Fig. The area integral of the stress vector may be expressed as
at x and x + ~x perpendicular to
Ji
.!#
t<:> dA = -Ji .!#
np 11 dA +Ji
.!#
(n 't') dA (9.2-9)

Directing our attention to the pressure term, we note the following:


l. p11 =O, on the free surface;
volume yields 2. i n = O, on the bottom of the channel;
. db . . db
(9.2-1) 3. 1 n = -i-d , on the s1des of the channel, prov1ded- ~l.
X ~

Under these conditions, we may write


may be expressed as
(9.2-2) Ji
.!#
Dp 11 dA J
= Pu dA L&z -
A,(z)
f P11 dA
A,(z)
L Pressure forces on the
entrance and exit of
thc control volume

for the flow to be steady.


we may write (9.2-10)
Pressure forces
At the entrance or exit of on the sides or
the control volume (9.2-3) the channel
358 Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow

lt should be noted that the last term in Eq. 9.2-10 is easily overlooked, for our and applying this result to Eq. 9.2-17,
attention is generally directed to the forces at the entrance and exit of the 1ess form of the differential-macroscop
control volume. 2
If we further assume that pg is independent of position across the channel, dh[g _ (v.,) ] _ (v,.Y
Eq. 9.2-10 takes the form dx g gh gh
For practica! purposes, gradually vari
-J
h(x) h(x) h(x)

f i npg dA =f pgb dz 1 pgb dz 1 - db ~xf pg dz (9.2-11)


() is Iess than 10, and under these con
""' x+.ix x dx
""" o o o
~ = cos
The viscous shear stress term is easily handled by the definition of 'To given in g
Eq. 9.1-11

f
J4
i (n -r) dA = -S ~x To (9.2-12)
which Ieads to
-g,. = sm fJ
g
o
R! t1

Collecting the various'terms in the momentum balance yields

p(v~) bh lx+dx - p(v!) bh lx = pg.,bh ~X


dh (1 -
dx
NFr)- NFr(~)b

- h(x)

f pgb dz
o
LHx - f
o
h(x)

pgb dz l., - e:) ~X h(x)

f pg dz - S
o
)

~X 'To (9.2-13)
We have replaced (v.,) 2 /gh with the Fr
the surface profile depends on whether
for which NFr = 1 are called critical.fiJ
.ftow and NFr < 1 a subcritical jlow.
Dividing by ~x and taking the Iimit ~x ---+ O, we get

f
h(x) h(x)
Mechanical energy balance
p .E_ ((v~)bh) = pg,.bh - !!._ f pgb dz + (db) pg dz - T0 S (9.2-14)
dx dx dx
o o Starting with Eq. 9.1-19, we expre
Expressing the gauge pressure as
pg = pg. [h(x) - z] (9.2-15) f
A.
2
(tpv )v n dA = iP<
carrying out the indicated integration and differentiation, and using the
approximationt Using Eq. 9.1-20 to express e/> and wri
(9.2-16)
we get pg = pg[(h
2 2
d(v,.) d (bh h db - we can combine the pressure and gra'
p(v,.)bh - - = pg,.bh - pg.- - )
+ pg.-- T 0S (9.2-17)
dx dx 2 2 dx
Here we have made use of the fact that - f pgvndA- f pcf>vndA =
A. A0
Q = (v,. )bh = constant (9.2-18)
Applying this result to the differenti
Differentiating Eq. 9.2-18 or Eq. 9.2-5 to obtain collecting the terms in Eq. 9.1-19, we
bh d(v,.) + (v,.)b dh + (v,.)h db = O (9.2-19)
dx dx dx
t Chow (Ref. 2, p. 28) indicates that Eq. 9.2-16 may be in error by as muchas 20 per
cent; hence, any study of real systems must incorporate the necessary correction factor.
Open Channel Flow Chap. 9
Sec. 9.2 Gradually Varied Flow 359

Eq. 9.2-10 is easily overlooked, for our and applying this result to Eq. 9.2-17, we can obtain the following dimension-
forces at the entrance and exit of the less form of the differential-macroscopic momentum balance,

dh[g _ (v.,?J _ (v.,) (~) db =


2
g., _ T 0 S (9.2-20)
dx g gh gh b dx g pgbh
For practica! purposes, gradually varied fiows are restricted to cases where
(9.2-11) O is less than 10, and under these conditions

g. =coso~ 1
g

(9.2-12) -g., = sm
uD ~ tan {) = - (dr)
-
g dx
which leads to
balance yields
dh (1 - NFr)- NFr(~)db = - (dr + ToS ) (9.2-21)
dx b dx dx pgbh
h(x) ) We have replaced (v.,)2/gh with the Froude number, so that it becomes clear
tlxf Pu dz - S tlx To (9.2-13) the surface profile depends on whether NFr is greater or less than one. Flows
for which NFr = 1 are called critica/ fiows; NFr > 1 represents a supercritica/
jlow and NFr < 1 a subcritica/ fiow.
-o, we get
h(x)

dz + (~~) f
o
Pu dz - ToS (9.2-14)
Mechanical energy balance

Starting with Eq. 9.1-19, we express the kinetic energy flux termas

(9.2-15)
and differentiation, and using the
J
A,
2
(ipv )v n dA = ip(v!>AI
x+4x
- ip(v!>AI
"'
(9.2-22)

Using Eq. 9.1-20 to express 4> and writing the gauge pressure as
(9.2-16)
Pu = pg[(h - z) cos O] (9.2-23)
(9.2-17) we can combine the pressure and gravitational terms in Eq. 9.1-19

-JPuV n dA - Jptf>v n dA = - pgJ(r + h cos 6)v n dA (9.2-24)


A, A, A,
(9.2-18)
Applying this result to the differential volume element in Fig. 9.2-2, and
collecting the terms in Eq. 9.1-19, we get
db
+ (v.,)h- =O (9.2-19) iP(<v!>AI - (v;)AI ) = - pg[(r + h cos O)(v.,)AI x+4x
dx x+4x x
.216 may be in error by as muchas 20 per
the necessary correction factor.
- (r + h cos O)(v.,)AIJ - Ev (9.2-25)
360 Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow

Dividing by !::u and taking the limit .x -- O, we have the differential macro- to eliminate the velocity gives
scopic mechanical energy balance

iP .!!._ (<v!>A) = - pg .!!._ [(11 + h cos e)(v.,)A] - e, (9.2-26)


dx dx
where For a given volumetric flow rate and s

e,=
1'1m- (")
Az-o .x
(9.2-27)
h. Exluding negative and complex valu
values of h or none. The solutions are
Assurning that for each value of Q2/2gb 2 there is a mil
is no solution for h. By inspection of
(9.2-28)
comes Iarge the fluid depth may tend t
and using Eqs. 9.2-18 and 9.2-19, we can rearrange Eq. 9.2-26 to obtain

!!_ [ (v.,)2 + h] = - (d1] + ~) (9.2-29)


dx 2g dx pgQ 6

providing e is less than 10. Referring to Eq. 9.1-26, we see that the viscous
dissipation per unit length e, for uniform ftow is given by 5

~=~= sin e= - d7] for uniform flow (9.2-30)


pgQ pgQL dx 4
lf the viscous dissipation for gradually varied flow is very nearly the same as
for uniform flow, we may expect the right-hand side of Eq. 9.2-29 to be small. h
Integration of Eq. 9.2-29 gives

(v.,)
2g
2
+h= -J (d1]
dx
+ ~) dx + constant
pgQ
(9.2-31)

which is generally written


(v.l
--
+ h-E
- Bp (9.2-32)
2g
E 8P is known as the specific energy, and it should be very nearly constant for
gradually varied flow, although the changes in the geometry and roughness of
the channel will certainly lead to small variations. Fig. 9.2-3. Fluid depth as a
Ifthe flow rate and channel width are given and values for ,.o are available,
either Eqs. 9.2-21 and 9.2-18 or Eqs. 9.2-29 and 9.2-18 can be used to solve
We may determine the value of h fe
for the two unknowns, (v.,) and h. In general, the process requires numerical
the derivative dEsp/dh equal to zero a1
integration, although analytic methods may be used to obtain approximate
solutions for simple cases.
dEsp = 1 _
The mechanical energy balance is not as powerful a too! for solving open
channel ftow problems as it was for solving closed conduit problems; dh
however, it can be used to good advantage to illustrate sorne important This condition corresponds to what i
characteristics of open channel ftow. Combination of Eqs. 9.2-18 and 9.2-32 energy, fluid depth, and velocity all a:
Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow 361
--+- O, we have the differential macro- to eliminate the velocity gives

[('17 + h cos O)(v.,)A] - e, (9.2-26)


2
- --
2 2
2gb h
+ h = E sp (9.2-33)

For a given volumetric flow rate and specific energy, Eq. 9.2-33 is a cubic in
h. Exluding negative and complex values of h, this equation yields either two
(!:) (9.2-27)
values of h or none. The solutions are shown in Fig. 9.2-3, where we see that
for each value of Q2/2gb 2 there is a mnimum value of Esv below which there
(9.2-28) is no solution for h. By inspection of Eq. 9.2-33, we can see that as E sp be-
comes large the fluid depth may tend toward either zero or Ew
rearrange Eq. 9.2-26 to obtain

(9.2-29)

Eq. 9.1-26, we see that the viscous


flow is given by

= - d17 for uniform flow (9.2-30)


dx
flow is very nearly the same as
riL-Jitar:tu side ofEq. 9.2-29 to be small. h

~) dx + constant (9.2-31)
pgQ

(9.2-32)

should be very nearly constant for


in the geometry and roughness of Esp

Fig. 9.2-3. Fluid depth as a function of specific energy.


given and values for r 0 are available,
and 9.2-18 can be used to solve
the process requires numerical We may determine the value of h for the mnimum value of E sp by setting
be used to obtain approximate the derivative dE5 vfdh egua! to zero and solving for h.

as powerful a too! for solving open (9.2-34)


solving closed conduit problems;
to illustrate sorne important
This condition corresponds to what is called critica! flow, and the specific
of Eqs. 9.2-18 and 9.2-32
energy, fluid depth, and velocity all assume their critica! values. Expressing
362 Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow

Eq. 9.2-34 in terms of the velocity, we find the critica! depth h. to be

h = (v.,)~
e
g
(9.2-35) ----
and the critica! velocity is
(9.2-36)

Substitution of these results into Eq. 9.2-33 indicates that the critica! specific
energy is always three halves the critica! depth
(9.2-37)

In terms of the Froude number, we note that Eq. 9.2-35 gives


Flg. 9.2-4. Establishm
for critica! flow (9.2-38)

There are perhaps two reasons why the condition NFr = 1 is designated as Assuming that the Manning formula
the critica! ftow. Ifwe remember that Eqs. 9.2-29, 9.2-30, and 9.2-31 indicated ftows, we can combine Eqs. 9.1-13 and!
the specific energy was only approximately constant, we can certainly expect of the volumetric ftow rate to obtain
small variations in Esp from changes in geometry and channel roughness. At
the critica! condition, the slope of h versus E 8 P is infinite; hence, small r0 = 14.5p
variations in E 8 P lead to large variations in h and wave motion is likely to
occur. For this reason, designers try to avoid such ftows. If we further restrict the analysis to ver
A second reason for designating NFr = 1 as the critica! ftow is that it the approximations
separa tes two fairly distinct ftow regimes. When NFr < 1, the fiow is sub-
critica/ or tranquil and disturbances can travel upstream; thus, a change in
downstream conditions can alter the ftow upstream. When NFr > 1, the
ftow is supercritica/ or rapid, and downstream disturbances are not propagated Under these conditions, we can combir
upstream. The speed at which a disturbance is propagated will be discussed in
Sec. 9.3. dh(1- ~)=sil
2 3
dx gb h
We have replaced -drfdx with sin O. T
Establishment of uniform flow
simplified somewhat by noting that
As an example of the application of the equations developed in this
section, we will investigate the gradually varied fiow indicated in Fig. 9.2-4
2 (gb2h
gb 2 = 2

to determine the length required for establishing uniform ftow. This example
because Q 2 fgb 2h~ is the Froude number
will also serve to introduce the student to the variety of surface profiles he
to one. Equation 9.2-41 can now be an
may encounter for any particular fiow system.
We start with the differential momentum balance, expressing the Froude
number in terms of the volumetric fiowrate and restricting the analysis to a ~~[1- (~n =sin o
rectangular channel of constant width.
If we designate hnt as the fluid depth f
dh(l-
dx
~)- -(dr + ~)
gb h 2 3
dx pgbh
(9.2-39) t The subscript n is used because the de
called the normal depth.
Open Channel Flow Chap. 9
Sec. 9.2 Gradually Varied Flow 363
find the critica! depth he to be

(v.,)~
g
(9.2-35)
-- --
(9.2-36)

indicates that the critica! specific


depth

(9.2-37)

that Eq. 9.2-35 gives


Fig. 9.2--4. Establishment of uniform ftow.
(9.2-38)

the condition NFr = 1 is designated as Assuming that the Manning formula can be applied to gradually varied
s. 9.2-29, 9.2-30, and 9.2-31 indicated flows, we can combine Eqs. 9.1-13 and 9.1-14 and express the result in terms
constant, we can certainly expect of the volumetric flow rate to obtain
geometry and channel roughness. At
Q2n2
versus Esp is infinite; hence, small To = 14.5p 2 2 1/ 3 (9.2-40)
in h and wave motion is likely to b h R,.
avoid such flows. If we further restrict the analysis to very wide channels, b > h, we can make
= l as the critical flow is that it the approximations
When NFr < l, the flow is sub- R,.~h
travel upstream; thus, a change in
s~b
flow upstream. When NFr > 1, the
disturbances are not propagated Under these conditions, we can combine Eqs. 9.2-39 and 9.2-40 to find
is propagated will be discussed in
dh( 1 -
- ~
Q2 ) .
= sm Q2n2
() - 14.5 2 1013 (9.2-41)
dx gb h gb h
We have replaced -dr/dx with sin O. The left-hand side of this result may be
simplified somewhat by noting that
of the equations developed in this
varied flow indicated in Fig. 9.2-4 Q2 -
gb2 -
(L)ha- ha
gb2h~ e - e
(9.2-42)
uniform flow. This example
to the variety of surface profiles he because Q 2jgb 2h~ is the Froude number for critica! flow and is therefore equal
system. to one. Equation 9.2-41 can now be arranged in the form
balance, expressing the Froude
and restricting the analysis to a dh[1-
dx
(he)3] =sin ()(1- 14.5Q2n2 )
h gb h 1 sin () 2 10 3
(9.2-43)

If we designate hnt as the fluid depth for uniform flow (h-+ hn as x-+ co),
-(dr + ~) (9.2-39) t The subscript n is used because the depth under uniform ftow conditions is often
dx pgbh
called the normal depth.
364 Open Channel Flow Chap. 9 Sec. 9.2

we may rearrange Eq. 9.1-16 to obtain

(9.2-44)

Our original differential momentum balance may now be written in compact


forro

dh
dx [ 1 - (hhe)3] = sin e[1 - (hhn)10/3] (9.2-45)

which is to be sol ved subject to a boundary condition of the type

B.C. 1: x=O (9.2-46)


Critic

Surface profiles

Before going on to obtain an approximate solution to Eq. 9.2-45, we will


discuss the possible surface profiles this equation may describe. Our purpose
is not to acquaint the student with the nine specific surface profiles associated
with the flow downstream from a sluice gate, but rather to drive borne the
point that flows in open channels can assume many possible forros. In his
discussion of gradually varied flow, Chow (Ref. 2, p. 222) lists severa! dozen Steep
commonly encountered surface profiles. Thus, in analyzing flow in open
channels it is extremely important to have knowledge of experimentally
observed surface profiles, for it is especially difficult to "guess the flow
topology" if the various possibilities are unknown. Fl. 9.2-5. Classifi
For a given volumetric flow in a given channel, it should be intuitively
obvious that the normal depth hn will decrease with increasing slope. This
Examining Eq. 9.2-45, we find the foil
idea is expressed quantitatively in Eq. 9.2-44 and leads us to classify channels
as mild, i.e., having a slope small enough so that the normal depth is greater Ml. dhfdx > O, backwater curve;
than the critica! depth, (hn > he), critica/ (hn = he), and steep (hn < he). M2. dhfdx < O, drawdown curve;
These conditions are illustrated in Fig. 9.2-5, and the flow at the critica! slope M3. dhfdx > O, backwater curve.
is wavy or undular in agreement with our previous comment on the instability
of critica! flow. We now wish to examine Eq. 9.2-45 for mild, critica!, and The condition that h increases with in
steep channels downstream and upstream from a sluice gate. called a backwater curve, while a dect
For mi!d channels, hn > he, and we must consider three possibilities in the The three profiles for mild channels are
gradually varied flow region. approaches the critica} depth, the profil
as Eq. 9.2-45 indicates, the flow cann<
Ml. h > hn >he, subcritical flow; The Ml backwater curve is the typc
M2. hn > h > he, subcritical flow; dam (or the sluice gate shown in Fig
M3. hn > he > h, supercritical flow. curve is important to reservoir desig
Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow 365

(9.2-44)
---

nce may now be written in compact

(9.2-45)

condition of the type

x=O (9.2-46)

solution to Eq. 9.2-45, we will


VAIIllla'""

equation may describe. Our purpose


specific surface profiles associated
gate, but rather to drive home the ........ ....
assume many possible forms. In his .........
.............
(Ref. 2, p. 222) lists severa! dozen Steep
........
.............................
Thus, in analyzing flow in open
have knowledge of experimentally
difficult to "guess the flow
unknown.
Fl1. 9.1-5. Classification of channels.
channel, it should be intuitively
decrease with increasing slope. This
and Ieads us to classify channels Examining Eq. 9.2-45, we find the following restrictions on dhfdx:
so that the normal depth is greater Ml. dhfdx > O, backwater curve;
(hn = he), and steep (hn < he). M2. dhfdx < O, drawdown curve;
and the flow at the critica! slope M3. dhfdx > O, backwater curve.
previous comment on the instability
Eq. 9.2-45 for mild, critica!, and The condition that h increases with increasing distance down the channel is
from a sluice gate. called a backwater curve, while a decrease in h is called a drawdown curve.
consider three possibilities in the The three profiles for mild channels are illustrated in Fig. 9.2-6. As the depth
approaches the critica! depth, the profile is drawn with a dashed line, because,
as Eq. 9.2-45 indicates, the flow cannot be gradually varied as h --+-he.
The Ml backwater curve is the type of profile that exists upstream from a
dam (or the sluice gate shown in Fig. 9.2-4). Knowledge of the backwater
curve is important to reservoir design, because the profile les above the
366 Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow

The C 1 and C2 backwater curves o<

--------- MI
may exhibit hydraulic jump. The C3 ba
from a sluice gate, and a mild, or undull
supercritical to critica! flow.

......__---- , - - - - '
-~-
---- ~--
-------~
-..._....._

-- ---- ---- --
~
--- ----
--... --<:......__
------- __
---------__------
. ......

-- ----
-_-:~ ............_.............._

~
Flg. 9.2-6. Surface profiles for mild channels.
--......

__
------- ---~, , -
normal depth of the channel and the flooding behind a dam may extend
beyond a plane established by the depth of the water in the reservoir.
The M3 backwater curve occurs downstream from a sluice gate. As the
critica! depth is approached a sudden transition from supercritical to sub- Fig. 9.2-7. Surface profilc
critica! flow takes place. This transition is known as a hydraulic jump.
The M2 drawdown curve does not occur for the configuration shown in
Fig. 9.2-4. However, it can occur when a change in channel slope from mild F or steep channels, he > h11 , and th
to critica! takes place.
For critica! channels, h.. = he, and again we have three possibilities in the SI. h > he> h.., subcritical flow;
gradually varied flow region. S2. he > h > h.., supercritical flow;
Cl. h >h .. = he, subcritical ftow; S3. he > h 11 > h, supercritical flow.
C2. h11 = he = h, critica! ftow; These conditions lead to the following
C3. h 11 = he > h, supercritical ftow.
Examination of Eq. 9.2-45 for this channel indicates: SI. dhfdx > O, backwater curvt:;
S2. dhfdx < O, drawdown curve;
Cl. dhfdx >O, backwater curve;
S3. dh fdx > O, backwater curve.
C2. dhfdx > O,t backwater curve;
C3. dhfdx > O, backwater curve. These surface profiles are illustrated in
The three profiles for critica! channels are illustrated in Fig. 9.2-7. The Sl backwater curve is the type t
t This result is easily obtained by application of I'Hospital's rule to Eq. 9.2-45; how- The S2 and S3 profiles can occur downs1
ever, the actual existence of such a profile is open to question. being similar to that illustrated in Fig.
Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow 367

The Cl and C2 backwater curves occur upstream from a dam, and both
may exhibit hydraulic jump. The C3 backwater curve may exist downstream
from a sluice gate, and a mild, or undular, jump occurs at the transition from
supercritical to critica! flow.

for mild channels.

flooding behind a dam may extend


of the water in the reservoir.
from a sluice gate. As the
transition from supercritical to sub-
Fig. 9.2-7. Surface profiles for critica! channels.
s known as a hydraulic jump.
occur for the configuration shown in
a change in channel slope from mild For steep channels, he> hm and the possible types of flow follow:

again we have three possibilities in the SI. h >he > hm subcritical flow;
S2. he > h > hm supercritical flow;
S3. he > hn > h, supercritical flow.

These conditions Iead to the following surface profiles:

indicates: SI. dhfdx > O, backwater curvt:;


S2. dhfdx < O, drawdown curve;
S3. dhfdx > O, backwater curve.

These surface profiles are illustrated in Fig. 9.2-8.


are illustrated in Fig. 9.2-7. The Sl backwater curve is the type that may occur upstream from a dam.
ofi'Hospital's rule to Eq. 9.2-45; how- The S2 and S3 profiles can occur downstream from a sluice gate, the S2 profile
to question. being similar to that illustrated in Fig. 9.2-5.
368 Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varled Flow

allows us to write Eq. 9.2-45 as

To extractan analytic solution from


mation
][10

If t < H < 2, this approximation w


however, it is not excessive when coro
in the use ofEqs. 9.2-16 and 9.2-28 a
factor.
Simplifying Eq. 9.2-47, separatin

where e is the constant of integrat


B.C. 1': H=H0

The integral in Eq. 9.2-49 is available


condition 1', is

Fig. 9.2-8. Surface profi.les for steep channels.

+ j_ [tan_1 (2H + 1)
Once again, the point to be made in this discussion is that a variety of
surface profiles are possible for a given geometrical configuration and it is
J3 J3
necessary to have sorne idea what to expect before plunging into an analysis. F or practica! purposes, we can assu
Returning now to Eq. 9.2-45, we wish to extractan approximate solution when h is within l per cent of h.,. or ll
for the approach to uniform flow downstream from a sluice gate. The fl.ow required to achieve uniform ftow
at the sluice gate cannot be classified as gradually varied; however, at a short designated by L,.
distance downstream fwm the gate the variations in fluid depth are more
gradual and Eq. 9.2-45 is satisfactory. Defining the dimensionless quantities
Entrance lengths

H = -h , dimensionless fluid depth For mild channels, only the M3


hn
gate, and a hydraulic jump occurs ir
X = x sin O dimensionless length type, 9l < 1, and the maximum entr
hn ' because the term in braces ({ }) is al
3. H. B. Dwight, Tables of Integrals a.
fJl = he , ratio of critical to normal depth
Macmillan Company, 1961), p. 42.
hn
Open Channel Flow Chap. 9 Sec. 9.2 Gradually Varied Flow 369

allows us to write Eq. 9.2-45 as


3
dH( 1 - Ha
dX [Jt )
= ( 1 )
1 - Htoa (9.2-47)

To extractan analytic solution from this equation we must make the approxi-
mation
pota~ na (9.2-48)
If i < H < 2, this approximation will lead to an error of about 25 per cent;
however, it is not excessive when compared to the percentage of error incurred
in the use of Eqs. 9.2-16 and 9.2-28 and the Manning formula for the friction
factor.
Simplifying Eq. 9.2-47, separating variables, and integrating, we get

f(~3-=-~ )3

dH = X +e (9.2-49)

where e is the constant of integration to be determined by application of


B.C. 1': X= O (9.2-50)
The integral in Eq. 9.2-49 is available, 3 and the result, after applying boundary
condition 1', is

(H- Ho) +(&la- 1){In [( 1 + H + (1- H2) Ho)2]


for steep channels. 6 1 + H 0 + H~ 1 - H
(9.2-51)
in this discussion is that a variety of
geometrical configuration and it is
before plunging into an analysis. For practica! purposes, we can assume that uniform flow has been reached
sh to extractan approximate solution when h is within 1 per cent of h .. or H is within 1 per cent of unity. The length
m from a sluice gate. The flow required to achieve uniform flow will be caiied the entrance length and
gradually varied; however, at a short designated by L .
variations in fluid depth are more
Defining the dimensionless quantities
Entrance lengths

For mild channels, only the M3 profile occurs downstream from a sluice
gate, anda hydraulic jump occurs in the neighborhood of H = 81. For this
type, ,tjt < 1, and the maximum entrance length will be obtained for 91--+ 1
because the term in braces ({}) is always positive. Letting 81 = 1 to obtain
to normal depth 3. H. B. Dwight, Tables of Integrals and Other Mathematica/ Data (New York: The
Macmillan Company, 1961), p. 42.
370 Open Channel Flow Chap. 9 Sec. 9.3 The Solitary Wave

an upper bound gives


L = (0.99- H 0)h., (9.2-52)
sin 8
and an order of magnitude estimate of the entrance length is

L.=o(~)
sm O
(9.2-53)

Since h., is generally on the order of feet, and sin O may be 0.01 or less, the
entrance length is on the order of 100ft or more for mild channels.
For critical channels, fJt = l and the entrance length is given by Eq.
9.2-52.
For steep channds, we see from Fig. 9.2-8 that there are two possible
profiles leading to a uniform flow.:
S2. fJt ~ H ~ 1;
S3. fJt ~ 1 ~ H.
For these two limiting cases we will take H 0 ~ 1 and H 0 < 1, respectively, to
obtain;
S2. L.= (h.,fsin O) [1.44fft3 + (1.01 - H 0)]; (9.2-54a)
S3. L. = (h.,fsin O) [2.02fft3 + 0.99]. (9.2-54b)
In bath cases, the en trance length can be hundreds of feet long, and it should
be apparent that uniform flow is the exception and varied flow the rule for
most practical cases.

9.3 The Solitary Wave Flc. 9.3-1. Gener:

To understand sorne ofthe phenomena encountered in open channel flow,


we must determine the speed at which a disturbance to the surface profile r------- - - -
moves. We first consider a disturbance caused by a movable wall which flh
boun4s an initially quiescent body of water. This process is illustrated in Fig.
9.3-1, and a sketch of the control volume to be u sed in the analysis is shown
in Fig. 9.3-2. We assume that the wave profile is preserved and choose a
ho h(x}
differential control volume that moves with sorne specific portion ofthe wave.

~"""'"~''
Note that the control volume is not a material volume, for it is the profile
which moves, not the water. In this respect, Watts' has written that:

"Buddhism has frequently compared the course of time to the apparent z


motion of a wave, wherein the actual water only moves up and down,
creating the illusion of a 'piece' of water moving over the surface."
4. A. W. Watts, The Way ofZen (New York: Pantheon Books, Inc., 1957).
Lx
Flc. 9.3-2. Moving control vol
Open Channel Flow Chap. 9
Sec. 9.3 The Solitary Wave 371

(9.2-52)

(9.2-53)
---
'
and sin () may be 0.01 or Iess, the
or more for mild channels.
the entrance Iength is given by Eq.

9.2-8 that there are two possible

H0 ~ 1 and H0 ~ 1, respectively, to

(9.2-54a)
(9.2-54b)
hundreds of feet long, and it should
and varied flow the rule for

Flg. 9.3-J. Generation of a solitary wave.

encountered in open channel flow, Control volume


a disturbance to the surface profile r-----------
caused by a movable wall which flh
. This process is illustrated in Fig.
to be used in the analysis is shown
profile is preserved and choose a 1----c
sorne specific portion ofthe wave. ho h(x)

material volume, for it is the profile


Watts' has written that:

the course of time to the apparent


water only moves up and down,
moving over the surface."
: Panthcon Books, Inc., 1957).
Flg. 9.3-1. Moving control volume for analysis of the solitary wave.
372 Open Channel Flow Chap. 9 Sec. 9.3 The Solltary Wave

Neglecting viscous effects and app,


Mass balance
in Fig. 9.3-2, we bave
We begin the analysis with the macroscopic mass balance d
p dx [(v.,(v.,-

~t Jp dV +Jp{v - w) n dA = O (9.3-1) Expressing the gauge pressure ast


7'"0 (t) A,(t)
pg=
and assume that w is a constant vector such that
and carrying out the integration it
w.,=c, w11 = w. =O (9.3-2)
d
p- [(v.,(v.,
The mass in the control volume is a constant so that Eq. 9.3-1 immediately dx
reduces to
Assuming that the velocity profil
J p(v - w) n dA
A,W
I.,+Az + J
A,W
p{v - w) n dA 1 =O (9.3-3)
((v.,)-e)
Here we have used the fact that Use of Eq. 9.3-7 to eliminate d(v.,
(v - w) n = O at the free surface (9.3-4) (e-
Note that Eq. 9.3-4 does not imply that the control volume moves with the or
fluid (w = v), it simply means that the control surface always coincides with e=
the free surface. Evaluation of the terms in Eq. 9.3-3 gives Provided the amplitude of the wa
depth !1h ~ h0, the fluid velocity
-J
h(z) ll(z)

f
o
(v.,- e) dz 1
z+Az
o
(v., - e) dz 1
.,
=O (9.3-5)
(v.,) ~e, and Eq. 9.3-14 reduces
e
where the density has been removed because the flow is incompressible. The result of our analysis app
Dividing by !1x, expressing the integrals in terms of average velocities, and tions-i.e., the wave profile is pt
taking the limit !1x --+ O, we get speed of a particular point on tt
constant. However, our analysis
d of waves m channels, for observ
dx [((v.,) - c)h] =O (9.3-6)
faster than the leading or trailing
which we may also write, in front of the crest. However, if
h d(v.,) + (v.,) dh = e dh (9.3-7) over the wave, and a reasonable :
dx dx dx

Momentum balance t This equation assumes that the hJ


the variations caused by fluid motion i
t There is sorne tradition associate
The momentum of the fluid in the control volume will be a constant, so time rate of change of position of the w
the x-component of Eq. 9.1-5 becomes word speed for v = .,;:-;: However,

J pv.,(v- w) n dA= Ji [-np., + n ~]dA


misinterpret.
(9.3-8) 5. J. J. Stoker, "The Formation o
1 :l.
..t,(t) .w
JI(
Open Channel Flow Chap. 9 Sec. 9.3 The Solltary Wave 373

Neglecting viscous effects and applying this equation to the control volume
in Fig. 9.3-2, we bave
/0(.,)
ros,:oo:tc mass balance
p ddx [(v.,(v., - e))h] = - ddx Jpg dz (9.3-9)
(9.3-1) o
Expressing the gauge pressure ast
pg = pg[h(x) - z] (9.3-10)
and carrying out the integration in Eq. 9.3-9, we find
(9.3-2)
stant so that Eq. 9.3-1 immediately p ~ [(v.,(v.,- e))h] = -pgh dh (9.3-11)
dx dx
Assuming that the velocity profile is flat, and using Eq. 9.3-6, we get
(9.3-3)
((v.,)- e)h d(v.,) = -gh dh (9.3-12)
dx dx
Use of Eq. 9.3-7 to eliminate d(v.,)fdx leads us to
(9.3-4)
( e- (v.,) ) 2 = gh(x) (9.3-13)
the control volume moves with the or
surface always coincides with e= (v.,) ../gh(x) (9.3-14)
in Eq. 9.3-3 gives
Provided the amplitude of the wave is small compared to the quiescent fluid
depth !l.h ~ h0 , the fluid velocity will be small compared to the wave speed~
(9.3-5) (v.,) ~e, and Eq. 9.3-14 reduces to
e= ../gh(x) (9.3-15)
the flow is incompressible.
The result of our analysis apparently contradicts one of the first assump-
in terms of average velocities, and
tions-i.e., the wave profile is preserved- for Eq. 9.3-15 indicates that the
speed of a particular point on the profile depends upon h(x), which is not
constant. However, our analysis does predict qualitatively the real behavior
(9.3-6) of waves m channels, for observation5 indicates that the wave crest travels
faster than the leading or trailing edge and that the wave tends to get steeper
in front of the crest. However, if !l.h ~ h0 , the speed will change only slightly
(9.3-7) over the wave, and a reasonable approximation for small amplitude waves is

e=~ (9.3-16)
t This equation assumes that the hydrostatic pressure variations are large compared to
the variations caused by fluid motion in the z-direction.
ntrol volume will be a constant, so t There is sorne tradition associated with the use of the word ce/erity to describe the
time rate of change of position of the wave profile, reserving the word velocity for v and the
word speed for v = VV:V. However, the term wave speed or wave velocity is difficult to
misinterpret.
[-np, + n -r) dA (9.3-8) 5. J. J. Stoker, "The Formation of Breakers and Bores," Comm. Appl. Math., 1948
b 1 :l.
37-4 Open Channel Flow Chap. 9 Sec. 9.3 The Solltary Wave

We must clarify the assumption associated with Eq. 9.3-10. Expressing the
pressure as hydrostatic is satisfactory only if pv!fh0 is negligible compared to
pg (inertial effects are therefore small compared to gravitational effects). A
careful study of wave motion indicates that this assumption is satisfied when
the wave length A. is long compared to ~he fluid depth. Such waves are often
called shallow-water waves, and the wave speed is given correctly by Eq.
9.3-16. Waves for which the wave length is less than the fluid depth are called
gravity waves, and their speed is given by .fih-<v.. >

e= /gi, gravity waves (A< h,} (9.3-17)


-J2;
We see that gravity waves move ata slower velocity than shallow-water waves.
If the wave length is very short-say, less than 1 cm-the wave velocity is
governed by surface tension effects and these waves are called ripp/es. Their
wave speed is given by

e -- !5'"pA.(f ' ripples (A < 1 cm} (9.3-18)

Both Eqs. 9.3-17 and 9.3-18 are restricted to the condition that the wave
amplitude is small compared to the wave length, Ah < A.; thus, the results
are applicable only for small disturbances.
An important point to be cleared up before we discuss the result derived
for shallow-water waves is the sign associated with the wave speed. There
is nothing in the analysis which tells us whether the wave generated in Fig.
9.3-1 will travel to the left or to the right, but there is little doubt in our minds Fig. 9.3-3. Propagati
that it will move to the right. One might ask, "If our analysis is correct, why
shouldn't it predict the direction in which the wave moves ?" The answer is The interesting case to consi
that one must investigate carefully the actual generation phenomena in order is given by Eq. 9.2-36 as
to predict the direction of wave propagation. Stoker7 has presented a thor-
ough treatment of this problem indicating that a disturbance always propa-
gates into the "region of quiet," i.e., the undisturbed region. In our problem Thus, for critical flow a disturb
the undisturbed region is to the right of the movable wall; hence, the wave nel; for subcritical flow (such a
moves to the right. may propagate upstream; and ~
An analysis of wave propagation in a flowing channel is similar to the carried downstream. Note tha~
development presented here and Eq. 9.3-14 is again obtained with the velocity waves because they have a small
(v,.) being the normal flow ve1ocity. It is thus indicated that a disturbance may propagate upstream in a e
may propagate downstream at a velocity equal to Jih
plus the stream
velocity, and upstream ata velocity equal to Jih
minus the stream velocity.
This situation is illustrated in Fig. 9.3-3 where a disturbance has been
created by momentarily immersing a gate into the stream. Thus, if the fluid depth is (fe
6. L. M. Milne-Thomson, Theoretica/ Hydrodynamics, 4th ed. (New York: The Such small waves are rapidly
Macmillan Company, 1960}, Chap. 14. that shallow-water waves, grav
7. 1. 1. Stoker, Water Waves (New York: Interscience Publishers, lnc., 1957), p. 303.
in supercritical flow.
Open Channel Flow Chap. 9 Sec. 9.3 The Solitary Wave 375

with Eq. 9.3-10. Expressing the


if pv!fho is negligible compared to
to gravitational effects). A
this assumption is satisfied when
fluid depth. Such waves are often
. speed is given correctly by Eq.
1s less than the fluid depth are called
.g; -<v.)
Jt

waves (A < h1) (9.3-17)

velocity than shallow-water waves.


than 1 cm-the wave velocity is
waves are called ripples. Their

(9.3-18)

to the condition that the wave


length, M ~ A; thus, the results

we discuss the result derived


lociatc:d with the wave speed. There
the wave generated in Fig.
there is little doubt in our minds Fig. 9.3-3. Propagation of a disturbance in a mild channel.
"If our analysis is correct, why
the wave moves ?" The answer is
The interesting case to consider is that of critical flow where the velocity
generation phenomena in order
is given by Eq. 9.2-36 as
Stoker7 has presented a thor-
that a disturbance always propa- (v~) = Jih (9.3-19)
region. In our problem Thus, for critica! flow a disturbance may remain fixed at a point in the chan-
movable wall; hence, the wave nel; for subcritical flow (such as that illustrated in Fig. 9.3-3), a disturbance
may propagate upstream; and for supercritical flow all s~pall disturbances are
flowing channel is similar to the carried downstream. Note that this latter statement also applies to gravity
is again obtained with the velocity waves beca use they ha ve a smaller velocity than shallow-water waves. Ripples
thus indicated that a disturbance may propagate upstream in a critica! channel, provided
equal to Jih
plus the stream
to Jih minus the stream velocity. ,
,.<-
27T(]
(9.3-20)
where a disturbance has been pgh
into the stream.
Thus, if the fluid depth is O(feet), the wave Iength must be about I0- 2 cm.
4th ed. (New York: The Such small waves are rapidly dissipated by viscous effects, and we conclude
that shallow-water waves, gravity waves, and ripples are carried downstream
~crscien<CC Publishers, lnc., 1957), p. 303.
in supercritical flow.
376 Open Channel Flow Chap. 9 Sec. 9.-4 Flow Over Bumps, Crests, a

In effect, we have airead y come to this conclusion in our analysis of the


ftow profiles upstream ofthe weir in Fig. 9.2-4; however, it will be worthwhile 9.4 Flow Over Bumps, Cr
to consider the matter further. In Fig. 9.3-4 the surface profiles upstream of a
dam have been sketched for mild and steep channels. In each case, the dam The presentation in this sectio
causes a dist'urbance in the ftow. For the mild channel, the ftow is subcritical to examine the broad-crested wei

Bockwoter curve Flow over a bump

-.,.....___ .___
-
We wish to examine the fto
------- . ~----

~- whether the surface will hump o


channel. Intuition would natural
appear but this assumption is no
are no variations in channel widt

dh
-(1- NFI
dx
Upstream and downstream from

Under these conditions,


dr
Critico! depth of
( dx
.. flow without dom
Flow cond1t1on controlled ot the ~ 1 The difficult question to answer i.
upstreom end 1 bump ?" The behavior of drfdx
1
decreases on the upstream side of
Hydroulic jump 1

side. Our comments regarding T


and for a start we shall assume th
over the bump. This statement is
Steep or su . forms as illustrated in Fig. 9.4-1
Percntico
slope
dr +.ToS )
( dx pgbh

( dr + 'ToS )
Fig. 9.3 ..... Su~face profiles upstream from a dam. <
dx pgbh
We now consider two cases:
and the disturbance is propagated upstream forming the smooth backwater
curve. In the steep channel, small disturbances are carried downstream; l. NFr > 1, supercritical ftow
however, the dam presents a finite disturbance and a hydraulic jump is 2. NFr < 1, subcritical ftow.
formed which propagates upstream at a velocity just equal to (vz> The t This statement assumes that we e
characteristics of hydraulic jumps will be treated in Sec. 9.5. channel.
Open Channel Flow Chap. 9
Sec. 9.<4 Flow Over Bumps, Crests, and Weirs 377
conclusion in our analysis of the
; however, it will be worthwhile 9.4 Flow Over Bumps, Crests, and Weirs
the surface pro files upstream of a
channels. In each case, the dam
The presentation in this section is mainly qualitative, the objective being
channel, the flow is subcrtical
to examine the broad-crested weir and describe how it works.

Flow over a bump

We wish to examine the flow illustrated in Fig. 9.4-1 and determine


whether the surface will hump or dip when passing over the bump in the
channel. Intuition would naturally lead u's to beleve that a hump would
appear but this assumption is not necessarily correct. We assume that there
are no variations in channel wdth so that Eq. 9.2-21 reduces to

dh (1 - NFr) = - (dr + ToS ) (9.4-1)


dx dx pgbh
Upstream and downstream from the bump the flow is uniform,t and

dh =o (9.4-2)
dx
Under these conditons,

dr) ( r0S ) (9.4-3)


( dx =- pgbh
The difficult question to answer is, "What happens to these quantities at the
bump ?" The behavor of drfdx is defined by the geometry; its absolute value
decreases on the upstream side of the bump and in creases on the downstream
side. Our comments regarding r 0 and h can only be vague generalizations,
and for a start we shall assume that nei.ther h nor r 0 changes as the fluid flows
over the bump. This statement is consistent with the assumption that a hump
forms as illustrated in Fg. 9.4-1. Based on this assumption, we may wrte

dr
( dx + ToS ) > O, on the upstream side (9.4-4a)
pgbh

( dr + 'roS ) < O, on the downstream side (9.4-4b)


dx pgbh
forming the smooth backwater We now consider two cases:
are carried downstream; l. NFr > 1, supercritical ftow;
and a hydraulic jump is 2. NFr < 1, subcritical flow.
just equal to (v,.). The
in Sec. 9.5. t This statement assumes that we do not have supercritical flow on a mild or critica)
channel.
378 Open Channel Flow Chap. 9 Sec. 9.'1 Flow Over Bumps, Crests, an

Supercriticol flow

-------------------1~:..-------~bcriticolL flow
Flow over a crest

We now wish to examine the :


in Fig. 9.4-2. Water flows from a
reservoir formed by an adjustable
take a chronological approach.

Z=H ~
Z=L-----------
Flg. 9.4-1. flow over a bump.

For case 1, Eqs. 9.4-1 and 9.4-4 indicate that

dh >o on the upstream side (9.4-Sa)


dx '

dh <0 on the downstream side (9.4-Sb)


dx '
Fig. 9.4-l. Flow
Thus, for supercritical flow the profile rises up over the bump, and our
assumptions regarding h and To in the neighborhood of the bump may be Let us consider first the case
ftow occurs and the..water in the 1
correct.
For case 2, Eqs. 9.4-1 and 9.4-4 indicate that so that a small fiow takes place, t
profile will occur at the crest. F
increase the ftow rate until the crit1
dh <o on the upstream side (9.4-6a) ftow has been subcritical and c1
dx '
However, once critica! flow is rea
dh . propagated upstream, and further
-> 0 on the downstream side (9.4-6b)
dx ' out the downstream portion of th'
tion is illustrated in Fig. 9.4-3, wh
Thus, for subcritical flow our analysis indicates that the profile dips down over the crest, the critica! fiow at the c1
the bump as indicated by the dashed curve in Fig. 9.4-1. This conclusion, of of the crest.
course, violates our assumption that h and Toare constant in the neighborhood
of the jump. However, if h decreases and -r 0 increases owing to the increased The broad-chested weir
velocity, the inequalities given by Eqs. 9.4-4 still hold, and we conclude that
the profile rises over a bump when the flow is supercritical and dips when the A broad-crested weir is shown
ftow is subcritical. We must note that this discussion is very qualitative in as ftow rate measuring devices, a
nature, and the behavior of the surface profile in the neighborhood of a bump viously discussed sharp-crested v
is best understood in terms of experimental observation. The analysis is put in analysis of this ftow is identical t1
its proper place if we simply state that it does not contradict the experimental where the application of Bernoull
observation. p + lpv2
Open Channel Flow Chap. 9 Sec. 9.4 Flow Over Bumps, Crests, and Weirs 379

Flow over a crest

We now wish to examine the problem of flow over a crest, as illustrated


in Fig. 9.4-2. Water flows from a reservoir over the crest and into a second
reservoir formed by an adjustable gate. In examining this problem, we shall
take a chronological approach.

Z=H----------------~L_-------------_

Z=L----------- h

(9.4-Sa)

(9.4-Sb)
Fig. 9.4-2. Flow over a crest in a reservoir.
up over the bump, and our
~~lJLuuiul;ouu of the bump may be Let us consider first the case where the control gate is raised so that no
flow occurs and the...water in the reservoir is quiescent. If we Jower the gate
so that a small flow takes place, the flow will be subcritical anda dip in the
profile will occur at the crest. Further Iowering of the control gate will
increase the flow rate until the critica! velocity is reached. Up to this time the
(9.4-6a)
flow has been subcritical and controlled by the downstream conditions.
However, once critica{ flow is reached small disturbances can no longer be
propagated upstream, and further lowering of the control gate simply washes
wnstream side (9.4-6b)
out the downstream portion of the dip in the pro file at the crest. This situa-
tion is illustrated in Fig. 9.4-3, which shows the subcritical flow upstream of
that the profile dips down over the crest, the critica{ flow at the crest, and the supercritical flow downstream
Fig. 9.4-1. This conclusion, of of the crest.
are constant in the neighborhood
increases owing to the increased The broad-chested weir
still hold, and we conclude that
supercritical and dips when the A broad-crested weir is shown in Fig. 9.4-4. Broad-crested weirs are used
discussion is very qualitative in as flow rate measuring devices, and have a certain advantage over the pre-
in the neighborhood of a bump viously discussed sharp-crested weir (Sec. 8.5). The initial portion of the
The analysis is put in analysis of this flow is identical to that presented for the sharp-crested weir
not contradict the experimental where the application of Bernoulli's equation to any streamline gave us
p + lpv 2 + pgz = p0 + pgH (9.4-7)
380 Open Channel Flow Chap. 9 Sec. 9.4 Flow Over Bumps, Crests, and

Subcriticol flow The broad-crested weir has its grea


Critico! flow reached the critica! value. Under ti
V .
e
and Eq. 9.4-9 is written as
Supercriticol flow Ve= J2i
We may now eliminate h., solve for
rate,
Q=b~
This result ts similar in form to t
' however, in deriving this equation
Flg. 9.4-3. Flow over a crest in a reservoir. coefficieut to correct the assumpti
been able to specify the fluid depth
Over the central portion of the weir, the streamlines are straight and the The results obtained in this se
pressure is hydrostatic, flows and negligible viscous effects.
should be regarded only as estmate
+
p =Po pg(he L - z) + (9.4-8)

and the velocity in Eq. 9.4-7 is given byt


Calculatiorl of flow rate
v = J2g .Jcn- L)- h (9.4-9)
For a value of 11 L-= 0.525
This equation could be used to determine the volumetric ftow rate over a unit width, and the fluid depth on
broad-crested weir if JI and h were measured with depth gauges. Because By Eq. 9.4 -13 the volumetric fio
the streamlines are straight, the volumetric ftow rate per unit width is
q = .Jg'[i(H- L))312 =
q = h.fii .Jcn- L)- h (9.4-10) or
q=l.
Because the flow on the weir will b

q =vA= Jih.
and. sol ve for he:
he= i(H
Experiments described by Doering
conditions,
q= 1
he= O
Thus, the actual flow rate is within
Flg. 9........ Broad-crested weir. the measured fluid depth on the wei
value. In view of the simplificatio
between measured and calculated f1
t Remembcr that Bernoulli's equation neglects viscous effects; therefore, it is not
unreasonable that Eq. 9.4-9 indicates a ftat velocity profile. 8. H. A. Doeringsfeld and C. L. Bark
Open Channel Flow Chap. 9 Sec. 9.4 Flow Over Bumps, Crests, and Weirs 381

The broad-crested weir has its greatest utility when the flow at the weir has
reached the critica! value. Under these conditions, the velocity is given by
v. = .fihc (9.4-11)
and Eq . 9.4-9 is written as
v. = J2i ~(H- L)- h. (9.4-12)

ulic jump We may now eliminate h., solve for v., and obtain the total volumetric flow
rate,
(9.4-13)
This result IS similar in form to that obtained for the sharp-crested weir;
however, in deriving this equation we have not had to introduce a discharge
coefficieut to correct the assumption of a flat velocity profile, and we have
been able to specify the fluid depth at the weir.
The results obtained in this section are restricted to gradually varied
flows and negligible viscous effects. Calculations made using these equations
L- z) (9.4-8) should be regarded only as estimates.

Calculatiorl of flow rate


(9.4-9)
For a value of H - L = 0.525 ft, determine the volumetric flow rate per
the volumetric flow rate over a unit width, and the fluid depth on a broad-crested weir.
with depth gauges. Because By Eq. 9.4-13 the volumetric flow rate per unit width is given by
rate per unit width is
q = .Ji[i(H- L)]312 = ~32.2 ft/sec 2[(i)(0.525 ft)] 312
L)- h (9.4-10) or
q = 1.17 ft 3/ft-sec
Because the flow on the weir will be critica!, we may express q as
q =vA= Jg'h;h. = }.g[i(H- L)]a12
and. sol ve for h.:
h. = i(H - L) = 0.35 ft
Experiments described by Doeringsfeld and BarkerS indicate that for these
conditions,
q = 1.08 ft 3/ft-sec
h.= 0.28 ft
Thus, the actual flow rate is within 8 per cent of the calculated value, while
weir. the measured fluid depth on the weir is 20 per cent smaller than the calculated
value. In view of the simplifications made in the analyses, the agreement
viscous effects; therefore, it is not between measured and calculated flow rates is remarkable.
profile.
8. H. A. Doeringsfeld and C. L. Barker, Trans. A.S.C.E., 1941, 106:934-46.
Open Channel Flow Chap. 9 Sec. 9.5 Hydraulic Jump
382

after the jump. Expressing the left-h


9.5 Hydraulic Jump and evaluating the right-hand side, '

A hydraulic jump occurs at a transition from supercritical to subcritical p(v!)3h3 - p(v!)


flow. This transition may be brought about by a change in the slope of the
Assuming flat velocity profiles and
channel, such as the jump illustrated in Fig. 9.4-3, or it may occur when a
we have
supercritical flow is formed on a rnild channel by a sluice gate. This situation
is illustrated in Fig. 9.2-4.
In analyzing the hydraulic jump, we will consider the flow shown in Fig.
9.5-1. As shown, it consists of a supercritical flow coming off a steep channel
There are two possible solutions to
onto a mild channel. On the mild channel, the depth gradually increases and
then undergoes a sudden change, which is called an hydraulic jump. We h2 =
will, in the course of this analysis, show that the jump may be lo<-ated as ort
indicated here or up on the steep portien of the channel.

Supercriticol flow Groduoliy voried flow


Here we have excluded the negativ
is not difficult to eliminate Eq. 9.5-~
9.2-21. For a supercritical flow on
Eq. 9.2-21 will be negative, as will
that dhfdx is positive and Eq. 9.5-6
Recognizing that q 2/gh~ is the Ii
jump, Eq. 9.5-7 reduces to

ha=-~-
2
Fl. 9.5-1. Hydraulic jump.
The nature of the jump depends ve1
lnasmuch as the hydraulic jump is highly turbulent and energy losses are approaching flow; Fig. 9.5-2 illus1
large, the momentum balance provides the only route to a direct solution. brief description of these different t:
Assuming steady flow and neglecting viscous and gravitational effects, l. NFr= 1 to 3. Standing wavc
Eq. 9.1-5 reduces to undular jump. The surface is

f pVzV n dA = - f i np, dA (9.5-1)


2. N Fr = 3 to 6. Rollers develc
downstream surface remains 1
A, 3. NFr = 6 to 21. The high-Sf
when dotted with the unit vector i. The assumption of nearly straight unstable jet producing Iarge
stream1ines before and after the jump allows us to express the gauge pressure can travel for miles and cause
as 4. NFr = 21 to 80. The action ol
p, = pg(h 2 - z), and position are Ieast sensith
(9.5-2)
dissipation:t: ranges from 45 t
p, = O, h 2 < z :-: ; ; h 3
t Work done by the Bureau of Red
before the jump, and between Eq. 9.5-7 and experimental obser
p, = pg(h 3 - z), O :::;;z :::;;h 3 (9.5-3) t The major effect of an hydraulic jum
Open Channel Flow Chap. 9
Sec. 9.5 Hydraulic Jump 383

after the jump. Expressing the left-hand side ofEq. 9.5-1 in terms ofaverages,
and evaluating the right-hand side, we get

from supercritical to subcritical


(9.5-4)
by a change in the slope of the
Fig. 9.4-3, or it may occur when a Assuming flat velocity profi1es and applying the macroscopic mass balance,
by a sluice gate. This situation we have

consider the flow shown in Fig. (h2 - ha)[(h2 + ha) - 2q2


gh2h3
J= O (9.5-5)
flow coming off a steep channel
the depth gradually increases and There are two possible solutions to Eq. 9.5-5; either
is called an hydraulic jump. We
h2 =ha (9.5-6)
that the jump may be located c1s
ort
of the channel.
h = _ h2 + Jh~ + (.!L) 2h2 (9.5-7)
flow a 2 4 gh~ 2

r-------~-~-~-~-
~~~~
1 Here we have excluded the negative root for ha as physically impossible. It
1 is not difficult to eliminate Eq. 9.5-6 as a possible solution by examining Eq.
1
1 9.2-21. For a supercritical flow on a mild channel, the right-hand side of
Eq. 9.2-21 will be negative, as will be the term (l - N~r). Thus, we conclude
that dhfdx is positive and Eq. 9.5-6 is not a possible solution.
Recognizing that q 2/gh~ is the Froude number for the flow prior to the
jump, Eq. 9.5-7 reduces to

(9.5-8)

The nature of the jump depends very strongly on the Froude number of the
turbulent and energy losses are approaching flow; Fig. 9.5-2 illustrates the four main types of jump. A
only route to a direct solution. brief description of these different types of jump is helpful.
and gravitational effects,
l. NFr = 1 to 3. Standing waves are formed and the jump is called an
undular jump. The surface is smooth.
(9.5-1) 2. NFr = 3 to 6. Rollers develop on the surface of the jump, but the
downstream surface remains smooth. The jump is called a weak jump.
3. NFr = 6 to 21. The high-speed flow entering the jump forms an
assumption of nearly straight
unstable jet producing large waves of irregular period. These waves
us to express the gauge pressure
can travel for miles and cause serious damage.
4. N~-r = 21 to 80. The action ofthis jump is well balanced and the action
(9.5-2) and position are least sensitive to the downstream depth. The energy
dissipationt ranges from 45 to 70 per cent.
t Work done by the Bureau of Reclamation (Ref. 9) indicates excellent agreement
between Eq. 9.5-7 and experimental observations.
(9.5-3)
t The majar etfect of an hydraulic jump is the dissipation of kinetic energy.
38-4 Open Channel Flow Chap. 9 Sec. 9.5 Hydraulic Jump

Undular jump
From Eq. 9.1-17, with b ~h, we o
h =
ha=
Weok iump Knowing ha and the flow rate, we
~-:, _; fluid depth, h2
....,..........- - - - " ' - ::::__-- NFr=3-6
h2-
Obviously, the jump from 2.04 to 2.6
the Froude number just upstream o

N F.,
Under these conditions, we have a1
in Fig. 9.5-2. The velocity just upst
may think of the standing wave
Steody jump
wave velocity of -9.8 ft/sec.
To determine the location of th
at the start of the mild channel is t
0.45 ft. This statement assumes tha
to a gradually varied flow that do
upstream. Application of Eq. 9.2-
beginning of the mild channel to t
If we decrease the slope of t
we obtain the conditions,
hl =
\~~~~\'\'\~~""""'0-..
Fig. 9.5-2. Various types of hydraulic jumps. h2'
ha'
5. NFr 80. The jump action is rough, and energy dissipation may be
Obviously the jump is a stronger
as high as 85 per cent. Undesirable waves are formed which travel
upstream from the jump (NFr = 2
downstream.
this case, Eq. 9.2-41 indicates tha
lt i~ important to ha ve sorne idea how to loca te the psition of a jump;
beginning of the mild channel. FUI
to outhne the method, we shall study a specific example for the flow shown
will strengthen the hydr11ulic jump a
in Fig. 9.5- 1. the start of the mild channel. Whe1
jump will occur at the beginning of
Location of the hydraulic jump channel is decreased even more, the
as Fig. 9.5-3 illustrate~. The jump
Let us consider the problem of locating the position of the hydraulic
under these conditions, because it n
jump shown in Fig. 9.5-1 for the following conditions:
both the steep and mild channels.
(a) 01 = 30; fairly well established. A detailed
(b) 02 = 5'; carried out by the Bureau of Recl~1r
(e) q = 20 fta/ft-sec;
9. "Hydraulic Design of Stilling Basi1
(d) channe1 width ~fluid depth;
graph No . 25, Hydraulics 1 aboratory f
(e) n = 0.014 ft11 6 Colorado. 1958.
Open Channel Flow Chap. 9 Sec. 9.5 Hydraulic Jump 385
Undular ump
From Eq. 9.1-17, with b ';t>h, we obtain the two uniform fluid depths

hl = 0.45 ft
ha= 2.62 ft
Knowing ha and the flow rate, we may rearrange Eq. 9.5-7 and solve for the
fluid depth, h 2
h2 = 2.04 ft
Obviously, the jump from 2.04 to 2.62 ft is nota strong one, and ifwe calculate
the Froude number just upstream of the jump we find

NF.r = 1.46
Under these conditions, we have an undular jump such as the one illustrated
in Fig. 9.5-2. The velocity just upstream from the jump is 9.8 ft/sec, and we
Steody jump may think of the standing wave f9rming the undular jump as having a
wave velocity of -9.8 ft/sec.
To determine the location of the jump, we first note that the fluid depth
at the start of the mild channel is the same as that in the steep channel, i.e.,
0.45 ft. This statement assumes that the change in channel slope gives rise
to a gradually varied flow that does not allow disturbances to propagate
upstream. Application of Eq. 9.2-51 indicates that the distance from the
beginning of the mild channel to the hydraulic jump is 700 ft.
If we decrease the slope of the mild channel from () = 5' to () = 3',
we obtain the conditions,
h = 0.45 ft
h2 = 1.71 ft
ha= 3.06 ft
and energy dissipation may be
waves are formed which travel Obviously the jump is a stronger one, although the Froude number just
upstream from the jump (NFr = 2.5) still indicates an undular jump. For
to loca te the psition of a jump; this case, Eq. 9.2-41 indicates that the jump takes place 560 ft from the
example for the flow shown beginning of the mild channel. Further decreases in the mild channel slope
will strengthen the hydraulic jump and move the position of the jump toward
the start of the mild channel. When the calculated value of h 2 is 0.45 ft, the
jump will occur at the beginning of the mild channel. lf the slope of the mild
channel is decreased even more, the jump will take place in the steep channel,
the position of the hydraulic as Fig. 9.5-3 illustrates. The jump is a good deal more difficult to analyze
conditions: under these conditions, because it may take place over a region occupied by
both the steep and mild channels. In any event, the location of the jump is
fairly well established. A detailed experimental study of hydraulic jumps
carried out by the Bureau of Rechrnation 9 has provided accurate design data.
9. "Hydraulic Design of Stilling Basins ancl f'nergy Dissipators," El{![ineeril{l{ Mono-
graph No . 25. Hydraulics Laboratory Branch, U S Rureau of Reclamation, Denver,
Colorado. 1958.
Problems
386 Open Channel Flow Chap. 9

Bz
9-3. Determine the slope of the mild
Flg. 9.5-3. Hydraulic jump. the hydraulic jump takes place at
Ans: 82 = 0.5 x 10-' radians
9-4. Apply the momentum balance to t

PRO~LEMS
the wave speed in a flowing strea1
9-1. Apply the same type of analysis used in examining the motion of a solitary 9-5. A uniform flow of water takes J
wave to deduce the surface profile for the depression wave shown in Fig. 9-1. with evenly laid unplaned timbe
Remember that the disturbance will propagate into the undisturbed region. horizontal is 6, the channel is 10
Comment on the shape of the disturbance as time progresses. the velocity, flow rate, and Frou
9-6. The surface roughness and slope
by design considerations; howeve1
ratio of the fluid depth to chan
area of flow is 100 ft 2, that will
9-7. How deep will the water flow in
channe1 if the slope is 0.001 and 1
mine the wall shear stress, T0 , an1
Ans: h = 1.3 ft, -ro= 0.065 lbtl!
9-8. Determine the volumetric flow n
flow depth is 3 ft and the slope il
9-9. U se the mechanical energy balane
the sluice gate shown in Fig. 9-9
at the gate, h2 Because the stre
traction or discharge coefficient
clearly why this is so). Show h
Flg. 9-1 . Creation of a depression wave. depth h3 could be used to deterrr
9-10. The volumetric flow rate in a 1C
9-2. The positive surge, or elevation wave, shown in Fig. 9-2 may be formed in a mnimum specific energy possible
riverbed by a cloudburst, or in an estuary by a rapidly advancing tidal front. the critica! veloeity?
In the laboratory, it can be formed by sudden changes in the position of a Ans: E 1 P = 4.6 ft, he= 3.0 ft, 1
sluice gate. The positive surge is often called a moving hydraulic jump.
Analyze the flow shown in Fig. 9-2 to obtain an expression for the velocity 9-11. Rederive the differential-macro
of the surge in terms of h1 and h2 corporating the effects of a con
Open Channel Flow Chap. 9 Problems 387

Fig. 9-2. The positive surge.

9-3. Determine the slope of the mild channel in the example in Sec. 9.5 such that
the hydraulic jurnp takes place at the beginning of the mild channel.
Ans: 82 = 0.5 X w-' radians
9-4. Apply the momentum balance' to the flow illustrated in Fig. 9.3-3 to determine
the wave speed in a flowing stream.
in examining the motion of a solitary
9-5. A uniform flow of water takes place in a rectangular channel constructed
depression wave shown in Fig. 9-1.
with evenly laid unplaned timber. If the angle between the channel and the
,.u ....at:au: into the undisturbed region.
horizontal is 6, the channel is 10ft wide, and the water is 2ft deep, compute
as time progresses.
the velocity, flow rate, and Froude number.
9-6. The surface roughness and slope (i.e., sin 8) of a rectangular channel are fixed
by design considerations; however, the width and depth are not. Calculate the
ratio of the fluid depth to channel width, subject to the restriction that the
area of flow is 100 ft 2 , that will provide a maximum discharge.
9-7. How deep will the water flow in a 15-ft wide, smooth, concrete, rectangular
channel if the slope is 0.001 and the volumetric flow rate is 87 ft 3/sec. Deter-
mine the wall shear stress, T 0 , and the Froude number.
Ans: h = 1.3 ft, T'O = 0.065 lbtlft 2 , NFr = 0.48
9-8. Determine the volurnetric flow rate in a 10-ft diameter sewer if the uniforrn
flow depth is 3 ft and the slope is 0.0002.
9-9. Use the mechanical energy balance to derive an expression for the flow through
the si u ice gate shown in Fig. 9-9 in terms of the fluid depth h1 and the depth
at the gate, h2 Because the streamlines are not parallel at the gate, a con-
traction or discharge coefficient must be included in the analysis (indicate
clearly why this is so). Show how experimental measurements of the fluid
depth h3 could be used to determine the contraction coefficient.
9-10. The volumetric flow rate in a 10-ft wide channel is 300 ft 3/sec. What is the
in Fig. 9-2 may be formed in a mnimum specific energy possible for this flow? What is the critica! depth and
by a rapidly advancing tidal front. the critica! veloeity?
sudden changes in the position of a
Ans: E 1 p = 4.6 ft, he = 3.0 ft, v 2 = 10.0 ft/sec
called a moving hydraulic jurnp.
an expression fo r the velocity 9-11. Rederive the differential-macroscopic mass and momentum balances in-
corporating the effects of a constant seepage into the channel. Express the
388 Open Channel Flow Chap. 9 Problems

9-15. Apply the momentum balance to an


to obtain a relationship between the
the bottom of the channel.
9-16. Derive an expression for the horizo
in Fig. 9-9.
h,
9-17. Water in a mild channe1 flows ata

L~~k-
from a holding basin to an irrigat
adjustment in a sluice gate at the
wave, how long will it take for this
system?
Ans: 10.4 min
Fig. 9-9. Fiow under a sluke gate. 9-18. If the amplitude of the wave in Prol
estmate the distance it will travel b
the initially symmetric wave profile
volumetric flow rate into the channel as cubic feet per second per unit length
of channel, and assume that the fluid enters the channel normal to the channel
walls.
9-12. Perform a numerical integration of Eq. 9.2-47 and compare the results with
1
Eq. 9.2-51 for the following thn:e cases: ~!
6ft
______
l. !JI = 3.0, H0 = 2.0
2. !JI = 3.0, H 0 = 0.5
3. !JI = 0.5, H 0 = 0.1
9-13. Water, ata velocity of 3 ft/sec and a depth of 2 ft, approaches a smooth rise
in a channel as shown in Fig. 9-13. Estima te the depth of the stream after the
3 in. rise.
Ans: h = 1.7 ft
Fig. 9-19. Flow thr

9-19. A rectangular channel 6ft wide hai


metric flow rate of 60 ft3/sec. A C<
trated in Fig. 9-19, and it may be as
changing the surface profile at the
varied on a mild channel.
(a) Determine the minimum value
remain undisturbed.
Fig. 9-13. flow over a rise in a channel. (b) Describe the surface profile u
striction for values of b0 greate
(e) If b0 is decreased further than th
9-14. Apply the mechanical energy balance, in conjunction with the momentum
surface both upstream and dov.
balance, to an hydraulic jump in a rectangular channel to derive an ex-
pression for the energy dissipation. Express the Ioss as a fraction of the
velocity head v2/2g and Froude number preceding the jump. Examine the
derived expression for NFr = 1 and NFr-+ co.
Open Channel Flow Chap. 9 Problems 389

9-15. Apply the momentum balance toan hydraulic jump in a 90 V-shaped channel
to obtain a relationship between the two fluid depths, h1 and h2 , measured from
the bottom of the channel.
9-16. Derive an expression for the horizontal force acting on the sluice gate shown
in Fig. 9-9.
9-17. Water in a mild channel flows ata velocity of 0.8 ft/sec and ata depth of 8ft
from a holding basin to an irrigation distributing system 2 mi away. If an
adjustment in a sluice gate at the holding basin produces a shallow-water
wave, how long will it take for this disturbance to be felt at the distributing
system?
Ans: 10.4 min
9-18. lf the amplitude of the wave in Prob. 9-17 is 6 in. and the wave length is 40ft,
estmate the distance it will travel before a pronounced skewness is formed in
cubic feet per second per unit length the initially symmetric wave profile.
the channel normal to the channel

. 9.2-47 and compare the results with

of 2 ft, approaches a smooth rise


the depth of the stream after the
---- --
Fig. 9-19. Flow through a constriction.

9-19. A rectangular channel 6ft wide has a uniform flow depth of 2ft anda volu-
metric flow rate of 60 ft 3/sec. A constriction occurs in the channel as illus-
trated in Fig. 9-19, and it may be assumed that inertial effects predominate in
in. changing the surface profile at the constriction. Treat the flow as gradually
varied on a mild channel.
1 (a) Determine the mnimum value of b0 for which the upstrearn depth will
remain undisturbed.
(b) Describe the surface profile upstrearn and downstrearn frorn the con-
striction for values of b0 greater than, or equal to, the rninirnurn value.
in conjunction with the momentum (e) If b0 is decreased further than the mnimum value, what will happen to the
,---,...- ..~ channel to derive an ex- surface both upstream and downstrearn of the constriction?
the loss as a fraction of the
preceding the jump. Examine the
Sec. 10.1 The Governing Equations for Co

Now, however, our problem has been


density as a function of temperature an
p = p(l

Compressible Flow 10 Thus, we find temperature as a new va1


introduced, which is, of course, the tot2

Conservation of energy

The fundamental postulate of con


words as
the rate at which
the time rate of ) heat is transferre
change of interna! _ to the body
and kinetic energy - (energy transfer
{ (
of a body through a rigid,
diathermal wall)
the rate
+ work is <
{ the body
forces su
Up to this point we have treated only incompressible ftows, or ftows for
The reference to heat as energy transfe~
which we could assume that the density p was a constant. Our purpose here
work as energy transfer through a mova
will be to investigate the simplest type of compressible ftow-i.e., one-
that heat and work are simply two di
dimensional flow in closed conduits. This study will provide an introduction
actually deal with "walls" or surfaces
to the use of the principie of conservation of energy in fluid mechanics
and to the phenomenon of shock waves. In presenting this material, we mechanisms.
lf we integrate Eq. 10.1-3 with res1
assume the student has had, or is presently taking, a course in thermodynamics.
Currently, the most important aspect of compressible ftow is the super- written as
fl(U + KE)
sonic motion of aircraft and rnissiles; however, this subject requires a more
detailed study than we have time for in an introductory text. For a more where =
U internal energy of the bo
comprehensive treatment, the student is referred to the two-volume work of =
KE kinetic energy of the bo
Shapiro. 1 Q= heat transferred to the b
W = work done by the body (
Equation 10.1-4 represents the first la
10.1 The Governing Equations for closed systems (i.e., matefial volumes).
Compressible Flow statement of the principie of conserva
usually encountered in thermodynami
In previous chapters the implied equation of state has been when the surroundings do work on th
between the body and the surrounding!
p = constant (10.1-1)
natural to retain the sllme sign conve
l. A. H. Shapiro, The Dynamics andThermodynamics ofCompressib/e Flow, Vols. 1 and energy transferred to the body is a ~
2 (New York: The Ronald Press Company, 1953). the transfer is accomplished by heat o
390 t Correctly speaking, Eqs. 10.1-4 and 1
the first law of thermodynamics being that
Sec. 10.1 The Governing Equations for Compressible Flow 391

Now, however, our problem has been expanded, and we must represent the
density as a function of temperature and pressure.
p = p(p, T) (10.1-2)

10 Thus, we find temperature as a new variable, and another equation must be


introduced, which is, of course, the total energy equation.

Conservation of energy

The fundamental postulate of conservation of energy may be stated in


words as
the rate at which ) (the rate at which sur-
the time rate of heat is transferred face work is done on
change of interna! _ to the body the body (energy
{ and kinetic energy} - (energy transfer + transfer through
(10.1-3)
of a body ( through a rigid, a movable
diathermal wall) adiabatic wall)

the rate at which }


+ work is done on
{ the body by body
forces such as gravity
incompressible flows, or flows for
p was a constant. Our purpose here The reference to heat as energy transfer through a rigid, diathermal wall, and
of compressible flow-i.e., one- work as energy transfer through a movable, adiabatic wall is given ~9 emphasize
study will provide an introduction that heat and work are simply two different modes of energy transfer. We
of energy in fluid mechanics actually deal with "walls" or surfaces which may transfer energy by both
. In presenting this material, we mechanisms .
taking, a course in thermodynamics. If we integrate Eq. 10.1-3 with respect to time, we obtain a result often
of compressible flow is the super- written as
' this subject requires a more d(U +
KE) = Q - W (10.1-4)
an introductory text. For a more
where U = interna! energy of the body
referred to the two-volume work of
KE = kinetic energy of the body
Q = heat transferred to the body (by the surroundings)
W = work done by the body (on the surroundings)
Equation 10.1-4 represents the first law of thermodynamicst as it applies to
c/osed systems (i.e., material volumes). The only difference between the word
statement of the principie of conservation of energy given here and the form
usually encountered in thermodynamics texts is that work is taken as positive
of state has been when the surroundings do work on the body. Since energy can be exchanged
(10.1-1) between the body and the surroundings in two ways (heat and work), it seems
natural to retain the slime sign convention for both these mechanisms. Thus
ofCompressible Flow, Vols. 1 and
energy transferred to the body is a positive quantity regardless of whether
the transfer is accomplished by heat or work.
t Correctly speaking, Eqs. 10.1-4 and 10.1-3 are definitions of the interna) energy;
the first /aw of thermodynamics being that the interna) energy is a state function.
392 Compressible Flow Chap. 10 Sec. 10.1 The Governing Equations for Co

The mathematical statement of Eq. 10.1-3 is Applying the divergence theorem to th


us to put all the terms under one inte
(a) (b) (e) (d) equation.
~~ f p(e + !v2) dV = - f q n dA + Jt<nl v dA + f pg VdV (10.1-5)
f,.(t) S'l,.(t) .r.f,.(t) f,.(t)

where e = interna! energy per unit mass Expressing gas - Vr/J, we may write
q = heat flux vector
pgv = -
The term (a) in this equation represents the time rate of change of internal
and kinetic energy of the material volume "Y m(t)-i.e., the body. Term (b) Remembering the formula for the mat
represents the heat transferred to the body. The negative sign results from
the fact that if the heat flux vector is pointing inward toward the body, the
vr,
-=--
aq,
quantity q n will be negative. Thus the negative sign in front of the integral Dt ot
is required. In terms of the symbolst used in Eq. 10.1-4, Since the gravitational potential q, is ind
may combine Eqs. 10.1-11 and 10.1-10
Q = - fqndA (10.1-6)
d,.(t)
pgv=-
Term (e) in Eq. 10.1-5 represents the rate at which work is done on the body
by surface forces. Term (d) in Eq. 10.1-5 represents the rate at which work Substitution of this result into Eq.
is done by the force of gravity. We will show shortly that this term can be equation in terms of interna!, kinetic,
represented as the material derivative of the potential energy of the body, Time rate of ehange
thereby allowing us to include it on the Ieft-hand side of Eq. 10.1-5: This of lntemal, klnetie
and potential energy
~
11
form of the energy equation may ha ve been encountered in thermodynamics per unlt volume u

texts where it would be written as


p !!._(e + tv2 + r/J) =
~(U+ KE + PE) = Q - W' (10.1-7) Dt
Here W' represents only the work done by surface forces. We will refer to this resultas the total e
If we make use of the special form of the Reynolds transport theorem show how the thermal energy equation
derived in Chap. 3, mechanica/ energy equation from Eq.
In the analysis of compressible flows
!!_fps dV =fp DS dV useful form ofthe principie of conserva tia
Dt Dt
r ,.(t) r ,.<tl to compressible flow, we will only use tJ
and express the stress vector as
and Eq. 10.1-13 must be integrated. Ti
presence of the density p outside the
t<n> = n T thfs difficulty before proceeding. Letti
Eq. 10.1-5 may be written as unit mass,

f p ~/e+iv2)dV=-f qndA+fpgvdV+f(nT)vdA and expanding the material derivative i


;r,.(t) .of,.(t) f,.(t) J1J(,.(t)

(10.1-8)
p-
an + pvVO = -
t As usual, the dot over Q indicates a time derivative. ot
Compressible Flow Chap. 10 Sec. 10.1 The Governing Equations for Compresslble Flow 393

10.1-3 is Applying the divergence theorem to the second and fourth integrals allows
us to put all the terms under one integral sign and extract the differential
(e) (d) equation.
+Jtcnl VdA+
.o/ ,.(t)
J pg VdV
j/" ,.(t)
(10.1-5) (10.1-9)

Expressing gas -Ve/>, we may write

pg v = - pv V e/> (10.1-10)
the time rate of change of internal
"Y mCt)-i.e., the body. Term (b) Remembering the formula for the material derivative, we write
body. The negative sign results from
inward toward the body, the Dc/>=ocf>+vVcf> (10.1-11)
negative sign in front of the integral Dt ot
in Eq. 10.1-4, Since the gravitational potential e/> is independent of time, ocf>fot = O and we
may combine Eqs. 10.1-11 and 10.1-lOto obtain
(10.1-6)
De/>
pgv= - p - (10.1-12)
Dt
at which work is done on the body
1-5 represents the rate at which work Substitution of this result into Eq. 10.1-9 yields the differential energy
show shortly that this term can be equation in terms of internal, kinetic, and potential energy.
of the potential energy of the body, Time ratc of changc
Jeft-hand side of Eq. 10.1-5~ This or intcrnal. klnctic
and potcntial cncrgy
Ratc of hcat
transfcr pcr
Ratc of surfacc
work pcr
been encountered in thermodynamics pcr unit volume unlt volume unlt volume

D 2
p - (e+ tv +e/>)= -V q +V (T v) (10.1-13)
(10.1-7) Dt
by surface forces. We will refer to this resultas the total energy equation. In Sec. 10.2, we will
of the Reynolds transport theorem show how the thermal energy equation may be derived by subtracting the
mechanica/ energy equation from Eq. 10.1-13.
In the analysis of compressible flows, the total energy equation is the most
p DS dV useful forro of the principie of conservation of energy. In this brief introduction
Dt to compressible flow, we will only use the macroscopic total energy balance,
and Eq. 10.1-13 must be integrated. The integration is complicated by the
presence of the density p outside the material derivative; we must remedy
thi.'s difficulty before proceeding. Letting O represent the total energy per
unit mass,
n =e+ tv2 cP + (10.1-14)
f pg v dV + f(n T) v dA and expanding the material derivative in Eq. 10.1-13, we get

(10.1-8) an
p - + pv Vil= -V q +V (T v) (10.1-15)
derivative. ot
Compressible Flow Chap. 10 Sec. 10.2 The Thermal Enery Equation ant
394

Noting that n [(opfot) +V pv] is zero by the continuity equation, we may


Adding a~d subtracting p/ p from O yic
add this term to the left-hand side of Eq. 10.1-15 to obtain a form suitable
for integration. Jp(
o
h + !v2 + tfo- ~)v n d
p
=-;: (pO) +V. (pvO) = -V q +V (T v) (10.1-16) where the integral representing the ra
dt
been replaced by Q.
Carrying out the integration over the arbitrary moving volume "Ya(t), and
As we have noted in several previ
applying the divergence theorem to the second, third, and fourth integrals,
and exits may be approximated by
we get

f ~~
1'"0 (!)
(pO) dV + J(pvO) n dA=
d (t)
- f
d
q n dA+ Jt<al v dA
4 (l) do(l)
(10.1-17)
Substitution of this expression for the 1
~> =.
0
a cancellation of terms and we obtain
Applying the general transport theorem, and noting that the convective energy
flux term will be zero except over the area A.(t), we obtain the macroscopic
total energy balance
f p(h + lv2 + 1/>)v
~t f
,V.(t)
pO dV + f f
pO(v- w) n dA=- q n dA+ Jt<al v dA
.A:,(t) .f.(t) .f.(t)
(10.1-18) If we neglect variations of h, v1, and
restrict oprselves to a system having
write Eq. 10.1-24 as
Splitting the rate of work term into area integralst over A,(t) and A.(t)
allows us to write A(h + !v1 +
~~ f pO dV
1'"0 (!)
+ f pO(v - w) n dA
.A:,(t)
=- f
da(t)
q n dA +W+ f t<al v dA
.A:,(t)
where m is the mass flow rate, and A
exit conditions (where v n is positive]
(10.1-19) Y n is negative).
Note that' the term W used throughout this text always refers to the rate of With the mass, momentum, and en
work done on the system by solid moving surfaces. In thermodynamics ready to analyze the propagation of a "
texts, this concept is often referred to as the useful work, while the work done and begin our investigation of compress
on the system at the entrance and exits (the last term in Eq. 10.1-19) is referred we should tie up sorne loose ends re
to as the ftow work. encountered in the derivation of the me
While Eq. 10.1-19 is a completely general forro of the macroscopic total section, therefore, we will derive the tl
energy balance, we shall generally restrict ourselves in this chapter to fixed the viscous dissipation does indeed apJ1
control surfaces at entrances and exits and steady state conditions. Under
these conditions, Eq. 10.1-19 reduces to
10.2 The Thermal Energy Equa
fpOv n dA
.,
=- f
.f
q n dA +W+ f
.,
t<a> v dA
for fixed control surfaces
at entrances and exits and
steady state conditions
the Entropy Equation
(10.1-20)
Our objective in this section is to p
It will be to our advantage to express this result in terms of the enthalpy per energy equations in proper perspectiv
unit mass We start the analysis with the total en
h=e+l?. (19.1-21) D
p p -(e + lv') = -V + PI Y+
t This integral is zero ov~ the fixed solid surface, A,. Dt
Compressible Flow Chap. 10 Sec. 10.2 The Thennal Enercy Equation and the Entropy Equation 395

by the continuity equation, we may Adding a~d subtracting p/ p from O yields


Eq. 10.1-15 to obtain a form suitable
fp(
~
h + tV1 + cfo- E)v n dA= Q + W + Jt1 1 vdA
p
~
(10.1-22)

-V q +V (T v) (10.1-16)
where the integral representing the rate of heat transfer to the system has
arbitrary rnoving volume "Ya(t), and been replaced by Q.
second, third, and fourth integrals, As we have noted in severa! previous examples, the stress at entrances
and exits may be approximated by

-f q n dA+
.ofoW
Jt(n) V dA
.ofoW
(10.1-17)
Substitution of this expression for the stress vector into Eq. 10.1-22leads to
(10.1-23)

a cancellation of terrns and we obtain


and noting that the convective energy
area A,(t), we obtain the macroscopic
J p(h + lv1 + cfo)v n dA = Q + W (10.1-24)

-f q n dA+
.., (t)
Jt(n)
JI(. (t)
V dA (10.1-18)
..t.
If we neglect variations of h, v1 , and cfo across the entrances and exits and
restrict ourselves to a system having one entrance and one exit, we may
write Eq. 10.1-24 as
area integra1st over A,(t) and A,(t)
(10.1-25)
-f q n dA
..,.w
+ W + f t(n) V dA
..t.< t)
where m is the mass ftow rate, and ti indicates the difference between the
exit conditions (where v n is positive) and the entrance conditions (where
(10.1-19) v n is negative).
this text always refers to the rate of With the mass, momentum, and energy balances at our disposal, we are
rnoving surfaces. In thermodynamics ready to analyze the propagation of a weak sonic disturbance (a sound wave)
the useful work, while the work done and begin our investigation of compressible ftows. However, before doing so,
(the 1ast terrn in Eq. 10.1-19) is referred we should tie up sorne loose ends regarding the viscous dissipation terrn
encountered in the derivation of the mechanical energy equation. In the next
general forrn of the macroscopic total section, therefore, we will derive the therrnal energy equation to show that
ourselves in this chapter to fixed the viscous dissipation does indeed appear as thennal energy.
and steady state conditions. Under
to
for fixed control surfaces 10.2 The Thermal Energy Equation and
at entr.ances and exits and
steady state conditions
the Entropy Equation
(10.1-20)
Our objective in this section is to put the total, therrnal, and mechanical
this result in terrns of the enthalpy per energy equations in proper perspective, and to derive the entropy equation.
We start the analysis with the total energy equation,
(19.1-21) D
p - (e + lv') = -V q + PI v +V (T v) IOtaiCMr~JequalioD (10.2-1)
surface, A . Dt
396 Compressible Flow Chap. 10 Sec. 10.2 The Thermal Energy Equation a

the stress equation of motion, temperatures, or temperature profil


studied in detail in courses in heat tr
Dv
p- = pg +VT (10.2-2) The total energy equation is gen
Dt temperature and kinetic energy are t
and the continuity equation course, for compressible flows.
We must be careful to remember
~: + V (pv) = O (10.2-3) derived from the laws of mechanics,
equations are two distinct equations.
Forming the scalar product of Eq. 10.2-2 with the velocity v yieldst under special restrictions, can dege
equation and the mechanical energy
p !!_ (!v 2 ) = pg v + v (V T) (10.2-4) regarding the role of the fundament2
Dt
The last term on the right-hand side may be expressed ast
Entropy equation
v (V T) = V (T v) - Vv : T (10.2-5)
In the next two sections, we wi
Substitution of this expression into Eq. 10.2-4 yields the mechanical energy entropy) processes. lt is therefore im
equation for compressible flow: so that we know what we mean by
that the interna! energy may be sp?.ci
= pg. v + V (T v) - Vv : T
p !!_ (!v2) mechanic:al energy equation (10.2-6)
Dt e=
Subtracting this result from the total energy equation yields the thermal we may take the material derivative
energy equation:

p-
De = -V q + Vv : T thermal energy equation ( l 0.2-7)
De =
Dt
(oe)
os p
D~
Dt
+ (oe
op
Dt
Multiplying this equation by p yields
If we write the total stress tensor as
T = - pi+ De .
"t' p- = 1p
Dt
the last term on the right-hand side of Eq. 10.2-7 may be expressed as
Remembering now that the continuit)
Vv : T = -pV v + <D (10.2-8)

Our general form of the thermal energy equation for compressible flow is
now written as we may write Eq. 10.2-12 as
De
p- = -V . q - pV V + <D (10.2-9) De
Dt p- =Tp
Dt
Thus, viscous dissipation always gives rise to an increase in the interna!
energy and, therefore, the temperature of the fluid. Substitution of pDe/Dt from Eq. 10.2-
The therma1 energy equation is most convenient to use when heat is for the rate of change of entropy
being transferred to or from the system, and we wish to know inlet and outlet Ds 1
p-=-(
t See Eqs. 7.3-2, 7.3-3, and 7.3-4. Dt T
t See Eqs. 7.3-9 through 7.3-12.
See Eqs. 7.3-13 through 7.3-17. t See Eq. 3.5-7.
Compressible Flow Chap. 10 Sec. 10.2 The Thermal Energy Equation and the Entropy Equation 397

temperatures, or temperature profiles and heat fl.uxes. This equation is


studied in detail in courses in heat transfer.
+V T (10.2-2) The total energy equation is generally used when significant changes in
temperature and kinetic energy are taking p1ace-the common situation, of
course, for compressible fl.ows.
(pv) = O (10.2-3) We must be careful to remember that the mechanical energy equation is
derived from the laws of mechanics, and that the total and thermal energy
2-2 with the velocity v yieldst equations are two distinct equations. The fact that the total energy equation,
under special restrictions, can degenerate into both the thermal energy
V + V. (V. T) (10.2-4) equation and the mechanical energy equation, has led to sorne confusion
regarding the role of the fundamental energy postulate in fluid mechanics.
be expressed ast
Entropy equation
(10.2-5)
. 10.2-4 yields the mechanical energy In the next two sections, we will be dealing with isentropic {constant
entropy) processes. 1t is therefore important to derive the entropy equation
so that we know what we mean by this term. Starting wi~h the statement
mechanicalenergy equation (10.2-6) that the interna! energy may be spP:cifi.ed in terms of the entropy and density
e = e(s, p) (10.2-10)
energy equation yields the thermal
we may take the material derivative of e to obtain

:T thermal cnergy equation (10.2-7) De


Dt
= (oe)
oso
Ds
Dt
+ (oe)
op s
Dp
Dt
= T Ds
Dt
+ .E. Dp
p2 Dt
(10.2-11)

Multiplying this equation by p yields

p De= Ip Ds + P. Dp (10.2-12)
Eq. 10.2-7 may be expressed as Dt Dt p Dt
Reme~bering now that the continuity equation may be written in the formt
{10.2-8)
Dp
equation for compressible flow is -+
Dt
pVv=O (10.2-13)
we may write Eq. 10.2-12 as
- pV V + 11> (10.2-9)
De Ds
p- = T p - - pV V (10.2-14)
rise to an increase in the interna! Dt Dt
of the fluid. Substitution of pDef Dt from Eq. 10.2-9 and rearrangement give an expression
convenient to use when heat is for the rate of change of entropy
and we wish to know inlet and outlet
Ds 1
p - =-(-V. q
Dt T
+ 11>) (10.2-15)

t Sec Eq. 3.5-7.


398 Compressible Flow Cflap. 10 Sec. 10.3 The Speed of ound

We now wish to determine the time rate of change of the entropy of a


material volume. Rearranging the right-hand side of Eq. 10.2-15 and
forming the integral, we obtain

f pD;dV=-f v.(~)dv+f (-q~T +~)dv (10.2-16)


D
7'"...<o7'".,.<l> T 7'".,.(ll T T
Using the special form of the Reynolds transport theorem given by Eq.
3.5-12, and applying the divergence theorem to the first integral on the right-
hand side, we have an expression for the rate of change of entropy.

.!!..fpsdV=-f qndA+f (-q.VT+~)dv (10.2-17)


Dt 7'".,.(l) ,,.<l> T 7'"...<t> T T
If a process is to be isentropic, the entropy of a body or of a material volume
involved in the process must be constantt therefore,

.!!_ J ps dV= O,
Dt 7'".,.<l>
for an isentropic process (10.2-18)

This condition can be satisfied as follows:


l. q n =O on the surface, i.e., the process is adiabatic;
2. VT
= O everywhere in the system, i.e., the process is "reversible";
3. ~ = Oeverywhere in the system, i.e., "frictional" effects are negligible.

In analyzing the speed of sound in Sec. 10.3 and flow in nozzles in Sec. 10.4, thus, the fluid far from the piston is
we will assume that the processes are isentropic even though viscous dissi- profiles are indicated above the tube
pation cannot be dentically zero and the temperature gradients are definite1y are comparable to the surface pro
finite. This assumption of isentropy greatly reduces the extent of the analysis shown in Fig. 9.3-1, and in what foil
while stil1 allowing us to investigate the key features of the flow. two different wave propagation phe
We will analyze this problem l
momentum balances, and because t
10.3 The Speed of Sound the x-direction, we must make the
x-direction. Figure. 10.3-2 illustrat~
used. We assume that the speed oft
We will determine the velocity ofsound by analyzing the density variations constant given by e; thus the contr
that occur when a small disturbance is created at the end of a tube, such as
that shown in Fig. 10.3-1. This process is comparable to what might happen
Mass balance
if the closed end of a cylindrical tube were hit with a mallet. If the flow were
incompressible, the density would be p0 throughout the tube and the velocity
Starting with the general macro
would everywhere be equal to the velocity of the piston at the end of the tube.

~tf pdV+ f
Since the flow is compressible, the density may increase above the value p,;
t Strictly speaking this is a sufticient but not a necessary condition. .<1)
7'" ..t,(
Compressible Flow Cl'lap. 10 Sec. 10.3 The Speed of ound 399

rate of change of the entropy of a


side of Eq. 10.2-15 and

transport theorem given by Eq.


to the first integral on the right-

r:::::::::::::::;
rate of change of entropy.

( -q. VT + ~) dV (10.2-17)
..<'> r r
of a body or of a material volume
therefore,

(10.2-18}

i.e., the process is "reversible" ;


Fi. IO.l-1. Generation of a sound wave.
"frictional" effects are negligible.

10.3 and fiow in nozzles in Sec. 10.4, thus, the fluid far from the piston is initially undisturbed. Illustrative density
'iserltr<J,pi'c even though viscous dissi- profiles are indicated above the tu be for each step of the process. The curves
temperature gradients are definitely are comparable to the surface profiles for open channe~ wave formation
reduces the extent of the analysis shown in Fig. 9.3-1, and in what follows we shall see that the analysis ofthese
key features of the fiow. two different wave propagation phenomena is similar.
We will analyze this problem in terms of the macroscopic mass and
momentum balances, and because the variable ofinterest p (or p) changes in
the x-direction, we must make the macroscopic balances differential in the
x-direction. Figure 10.3-2 illustrates the differential control volume to be
used. We assume that the speed of the density variation (or sound wave) is a
by analyzing the density variations constant given by e; thus the control volume moves with the sound wave.
at the end of a tube, such as
comparable to what might happen
Mass balance
hit with a mallet. lf the fiow were
lthr<OUJl:hOllt the tube and the velocity
Starting with the general macroscopic mass balance,
of the piston at the end ofthe tube.
may increase above the value p1 ;
a necessary condition.
~ J J
pdV+ p(v- w)ndA =O
dt ~.<O ....(1)
(10.3-1)
400 Compressible Flow Chap. 10
Sec. 10.3 The Speed of Sound

Control volume Momentum balance

We start with the general moment

-W._ /
r-
1
,
1
Density profile ~t f pv dV + f pv(v - w) 1

____ _ /___ L _ ~~ - - - - P o
r.(l) A,<t>

1 1
and note that the momentum of the
1 1 the scalar product with i to obtain
: '-~---e
1
1
L
1
___ J1 f pv,.(v- w) n
A,(l)

Remembering that viscous effects are


Eq. 10.3-7 to the moving control voh

Fl. 10.3-2. Moving control volume for analysis of sound wve. f pv.,(v., -
.A.(I)
e) dAI
o:+ Ao:
- f pv.,(v.,-
A ,(1)
we note first that since the control volume moves with a specific portion of
the wave having a fixed density, the first term is zero.
Again assuming flat velocity profile

~
dt
f pdV=O
7"'.(1)
(10.3-2)
6.x-+ O, we obtain
d
dx [pv.,(v., - o
We see that i w = e, and the remaining area integral takes the form Using Eq. 10.3-5a allows us to simpl

(10.3-3) p(v.,- e)A(~


dJ
Taking the cross-sectional area A to
While the wave velocity e may be quite large, the actual fluid velocity v., will eliminate dv.,fdx from Eq. 10.3-10 an
be small. Viscous effects may be neglected,t and the velocity profile may be
assumed to be ftat. Under these conditions, we may divide Eq. 10.3-3 by 6.x
and take the limit 6.x -+ O to obtain
Solving for e gives
-d [p(v.,- e)A] = O (10.3-4) e= v,
dx

which may be rearranged in two convenient forms, and, provided v., ~ -Jdpfdp, our expr
to
p(v., - e)A = constant (10.3-Sa)
e=
d dp
p-
dx
[(v"'- e)A] + (v.,- e)A-
dx
=O (10.3-Sb)
Note that the analysis indicates th
direction; however, we know intuiti1
t This is an experimental observation. An analytical proof would be difficult.
as Fig. 10.3-2 indicates. A rigorow
Compressible Flow Chap. JO
Sec. 10.3 The Speed of Sound
401
Control110lume
Momentum balance

We start with the general momentum balance,

~t Jpv dV +Jpv(v - w) n dA = Jr.ct>pg dV + Jten> dA (10.3-6)


r.m A.(l) .w.co
----Po
and note that the momentum of the control volume is constant. We form
the scalar product with i to obtain

J pv,.(v- w) n dA
A.(t)
=Ji 0
(1)
ten> dA (10.3-7)

Remembering that viscous effects are being neglected and i w = e, we apply


Eq. 10.3-7 to the moving control volume in Fig. 10.3-2 and obtain
for analysis of sound wave.
J pv,.(v., - e) dA'
A.(l) x+Ax
- J pv,.(v., - e) dA'
A .(1) .,
= - J p dA'
A .(t) x+A:t:
+ Jp dA'
A . (1) .,
moves with a specific portion of
term is zero. (10.3-8)
Again assuming flat velocity profiles, dividing by llx, and taking the limit
(10.3-2) llx-.. O, we obtain

area integral takes the form


~x [pv.,(v.,- e)A] = - e:)A (10.3-9)

Using Eq. 10.3-5a allows us to simplify this result to


(10.3-3)
p(v.,- e)A(~:"') = -(~:)A (10.3-10)

large, the actual fluid velocity vo: will Taking the cross-sectional area A to be constant, we may use Eq. 10.3-5b to
t and the velocity profile may be eliminate dv.,/dx from Eq. 10.3-10 and obtain
we may divide Eq. 10.3-3 by llx
(10.3-11)
Solving for e gives
(10.3-4) p
e= v., p (10.3-12)

and, provided v., ~ ,Jdp/dp, our expression for the velocity of sound reduces
to
(10.3-Sa)
e=~ (10.3-13)
(10.3-Sb) ..J-;Jp
Note that the analysis indicates that the wave may propagate in either
direction; however, we know intuitively that it will move from left to right
as Fig. 10.3-2 indicates. A rigorous proof of this fact involves extensive
Compressible Flow Chap. 10 Sec. 10.3 The Speed of Sound

analysis. Such a study has been carried out by Stoker2 indicating that a Ta
disturbance always propagates into the "region of quiet." GAS CONSTANT R ANO
If the process is adiabatic and reversible, the entropy, s, will be constant SOME (J
(i.e., the process is isentropic), and Eq. 10.3-13 is written as

e =m.= {the speed of sound} (10.3-14)


Gas

Hydrogen, H 1
For air at atmospheric pressure and normal temperatures, Eq. 10.3-14 Helium,He
Nitrogen, N,
indicates that the speed of sound is about 1100 ft/sec. Our definition of a
Oxygen, 0 2
"weak disturbance" will be that v., (in Eq. 10.3-12) is much smaller than the Carbon dioxide, C02
speed of sound. In the process under consideration, we may assume that the A ir
fluid velocity v., is equal to, or less than, the velocity of the piston; our Water vapor, H 20
analysis will therefore be correct provided the velocity of the piston is much
less than the speed of sound.
For a perfect gas, the pressure and density in an isentropic process are
This analysis has introduced us
related by
p
-=
p7
constant (10.3.15)
in the analysis of compressible fio
the velocity of sound, compressibi
velocities comparable to, or greate
where y = ef}fe.,, the ratio of specific heats. The equation of state for a unusual situations where the co
perfect gas is generally expressed as Therefore, the ratio M of the fi
pV = nf!IT (10.3-16) important measure of the compres
after the physicist, Ernest Mach. 3
where f!l is the universal gas constant and equal to 82.06 atm-cm3 /g-mole K,
and n is the number of moles. This form is generally encountered in thermo-
dynamics texts where the quantity of material is specified in terms of the
number of moles. In fluid mechanics, the mass is a more useful measure of
the quantity of material contained in a system, and Eq. 10.3-16 is written in The common designation of the fi<
the form of the Mach number are as follow1
pV= mRT (10.3-17)
where M~1,
R = _ __:f!l_,_ _
molecular weight t <M< l,
An even more convenient form results from dividing by V to yield M~ 1,
p= pRT (10.3-18) M> 1,
Here we must remember that R is different for every gas. By differentiating
Eq. 10.3-15 with respect to p, using Eq. 10.3-18 to eliminate p, and substituting Because the velocity of sound in 1
the result into Eq. 10.3-14, we can express the velocity of sound as mph), compressibility effects are 1
treatment of compressible fiows is
e= ../yRT (10.3-19)
t Remember that the circulation ve)
Values of R and y for sorne common gases are listed in Table 10.3-1. small compared to the speed of sound.
3. H. Rouse and S. Ince, History of 1
2. J. J. Stoker, Water Waves (New York: Interscience Publishers, Inc., 1957), p. 303. 1963), p. 195.
Compresslble Flow Chap. 10
Sec. 10.3 The Speed of Sound <403
2
out by Stoker indicating that a
Table 10.3-1
"region of quiet."
the entropy, s, will be constant GAS CONSTANT R ANO RATIO OF SPECIFIC HEATS y FOR
SOME COMMON GASES
. 10.3-13 is written as
y, standard
Gas conditions
(10.3-14)
Hydrogen, H 1 766.5 1.41
normal temperatures, Eq. 10.3-14 Helium, He 386.3 1.67
1100 ftfsec. Our definition of a Nitrogen, N 1 55.15 1.40
Eq. 10.3-12) is much smaller than the Oxygen, 0 2 48.29 1.39
we may assume that the Carbon dioxide, C01 35.12 1.29
A ir 53.35 1.40
the velocity of the piston; our
Water vapor, H 20 85.8 1.33
the velocity of the piston is much

This analysis has introduced us to the velocity of sound, a key parameter


(10.3.15) in the analysis of compressible flow. If the fluid velocity is much less than
the velocity of sound, compressibility effects are usuallyt small, while fluid
of state for a velocities comparable to, or greater than, the speed of sound lead to sorne
unusual situations where the compressibility effects domnate the flow.
Therefore, the ratio M of the fluid velocity to the speed of sound is an
(10.3-16) important measure of the compressibility effects; we call it the Mach number
3 after the physicist, Ernest Mach. 3
equal to 82.06 atm-cm /g-mole K,
is generally encountered in thermo-
material is specified in terms of the M=~ (10.3-20)
the mass is a more useful measure of e
system, and Eq. 10.3-16 is written in
The common designation of the flow regimes associated with various values
(10.3-17) of the Mach number are as follows:

M ~ 1, incompressible flow
t <M< 1, subsonic flow
M~ 1, transonic flow
M> 1, supersonic flow
for every gas. By differentiating
10.3-18 to elimina te p, and substituting Because the velocity of sound in liquids is approximately 5000 ft/sec (3400
the velocity of sound as mph), compressibility effects are rarely of importance in liquids, and the
treatment of compressible flows is generally restricted to gases.
(10.3-19)
are listed in Table 10.3-1. t Remember that the circulation velocities occurring in a heated pan of water are
small compared to the speed of sound.
: Interscience Publishers, Inc., 1957), p. 303. 3. H. Rouse and S. Ince, History of Hydrau/ics (New York: Dover Pub1ications, Inc.,
1963), p. 195.
Compressible Flow Chap. 10 Sec. 10.4 lsentropic Nozzle Flow

the flow in a nozzle, the derived


Speed of sound In helium and air isentropic flows in ducts of variable
important to note precisely where
We wish to compare the speed of sound in air and in helium at 72F.
For air, Eq. 10.3-19 gives

= [<L 40)(53.35 ~t-lbr)< 53 rR)(32.2lbm~ft) ]


112
Calr
lbm- R lbr-sec
r (x)
= 1130 ftjsec 0

Similarly, the velocity of sound in helium is calculated to be


cae = 3330 ftjsec
Thus, the velocity of sound in helium is approximately three times as large
as that in air. This difference gives a curious quality to the voice if one speaks
with a lung full of helium.

10.4 lsentropic Nozzle Flow

The simplest example of compressible flow is the isentropic flow from a


large reservoir through a converging nozzle, such as that shown in Fig.
Fig. 10.4-2. Control volu
10.4-1. The assumption of constant entropy requires that the process be
we will remove this restriction in
We shall now discuss the form tH
Reservoir 1 Reservoir n
applied to the differential control vo
Pressure = p 0 Bock pressure = J&
Temperoture = r
Velocity ~O Mass balance

Application of Eq. 10.3-1 to the

d
-(p(
dx
or
Flg. 10.4-1. lsentropic ftow through a nozzle. p(v., )A :

Momentum balance
adiabatic, and that q VT and <I> be negligib1e. The assumption of adiabatic
flow is reasonable, and we may neglect viscous dissipation for flow in a nozzle
without incurring much error. Because fairly large temperature changes take Starting with Eq. 10.3-6, we for
place in the nozzle, the assumption that q VT is negligible must be based on
the fact that the thermal conductivity of gases is small. Therefore, f
A.
pv!dA'
~t+&~t
- f
A.
q = -kVT~::::i O
t Here we have assumed that the dem
Although our analysis in this section will be directed toward understanding the nozzle.
Compressible Flow Chap. 10 Sec. 10.4 lsentropic Nozzle Flow
405
the flow in a nozzle, the derived equations will be generally applicable to
isentropic flows in ducts of variable cross-sectional area. It will be especially
sound in air and in helium at 72F. important to note precisely where the isentropic restriction is imposed, for

1 2
(532 OR ) (32.2lbm-ft)]
lb 2
'
r-sec

is calculated to be
r

is approximately three times as large -'-----+--:----t--~f _ _ _ _ _ x


quality to the voice if one speaks

flow is the isentropic flow from a


nozzle, such as that shown in Fig.
requires that the process be Flg. 10.4-l. Control volume for analysis of nozzle flow.

we will remove this restriction in our analysis of shock waves in Sec. 10.5.
Rnervoir n We shall now discuss the form that the governing equations take when
applied to the differential control volume shown in Fig. 10.4-2.
Bock pressure = ,

Mass balance

Application of Eq. 10.3-1 to the differential control volume yieldst

!!..._ (p(v.,)A) = O (10.4-la)


dx
or
p(v., )A = constant (10.4-1 b)

The assumption of adiabatic Momentum balance


dissipation for flow in a nozzle
large temperature changes take Starting with Eq. 10.3-6, we form the scalar product with i to obtain
q. VTis negligible must be based on
gases is small. Therefore,
f
A.
pv! dAj
z+A
., - f
A.
pv! dAj
"'
= f
.si
. t<> dA (10.4-2)

t Here we have assumed that the density is constant at any given cross section of
be directed toward understanding the nozzle.
Compressible Flow Chap. 10 S.C. 10.4 lsentroplc Nozzle Flow
406

In evaluating the term on the right-hand side ofEq. 10.4-2, we must remember Provided the ve1ocity profi1e is essen
that the outwardly directed unit normal n from the control volume is given across the duct, Eq. 10.4-8 quick1y
by h + t<v.,)
in= 1, at the entrance and exit
The assumption that (v2 )
= (v.,) 2 is
i n =sin O, at the walls of the nozzle
dr 0 jdx ~ l. Further restriction ofth
where is a linear function of temperature a
nv =tan- - dro)
-
dx
1( h =e'/)
and Eq. 10.4-9 becomes
To keep the analysis reasonably simple, we must impose the restriction that c'/)T + t<v
dr 0 jdx ~ 1, so we may write
The constant may be evaluated by ,
. n n dro (10.4-3)
sm v ~ tan v = - - (v.,) =O and T = T", which gives
dx
Inasmuch as we are neglecting viscous effects, the stress vector is given by
c'/)T+ t
t(a) = -np, and the right-hand side of Eq. 10.4-2 becomes where r is called the stagnation tem
is unusual and is done to clearly de~
f
_.,
i t<n> dA= - f
A.,
p dA!.,Hz + f
A.,
p dA!.,+ (~:o) ~X f
S
p ds (10.4-4)
and not the temperature at point
pressure are the temperature and pr
if it were isentropically brought to re
Substituting Eq. 10.4-4 into Eq. 10.4-2, dividing by ~x and taking the limit it is often convenient to specify lo
~x -O, we obtain temperature and pressure and the Ma

+ 27Tr0 (d'0)p
important quantities.
t!.._ (p(v!)A) = - ci_ {(p)A) (10.4-5) We now have at our disposal the
dx dx dx
along with the equation of state for
Here p represents the mean pressure over the bounding circular strip of the sol ve for the four unknowns (T, p, f
control volume. If we now assume flat velocity profiles and a uniform
pressure across the duct (i.e., (p ) = p = p), this result reduces to
and
ci_ (p(v.,)2A) = -A (dp) (10.4-6)
(v.,) =
dx dx
which we may simplify further by using Eq. 10.4-1 b, Equation 10.4-12 can be used to expr
the local Mach number and the stag
p(v.,) d(v.,) = _ (dp) (10.4-7)
dx dx T= ro( 1
Energy balance Three assumptions led to this resul1
perature; the equation of state for
We begin with the simplified form of the energy balance given by Eq. perfect gas law; and the flow is 011
10.1-24, and note that W = Q = Oto obtain isentropic flow has not yet been imp
Since y > 1.0, Eq. 10.4-15 indic
f p(h + tv 2 + cfo)v n dA = O (10.4-1) the velocity (and therefore the Mach
of this conversion as one of thermal
Compressible Flow Chap. 10 S.C. 10.4 lsentroplc Nozzle Flow

ofEq. 10.4-2, we must remember Provided the velocity profile is essentially ftat and the enthalpy does not vary
n from the control volume is given across the duct, Eq. 10.4-8 quickly reduces to
h + t(v.,)2 = constant (10.4-9)
The assumption that (v2 ) = (v.,)2 is reasonable in light of the restriction that
walls of the nozzle
dr0 fdx ~ l. Further restriction ofthe analysis to gases for which the enthalpy
is a linear function of temperature allows us to write
_ dr0 )
h = c'IJT + constant (10.4-10)
dx
and Eq. 10.4-9 becomes
we must impose the restriction that c'IJT + ! (v.,) 2 = constant (10.4-11)
The constant may be evaluated by applying this result to reservoir 1 where
dr 0 (10.4-3)
=-- (v.,) =O and T = r', which gives
dx
(10.4-12)
the stress vector is given by
0
. 10.4-2 becomes where T is called the stagnation temperature. The use of "o" as a superscript
is unusual and is done to clearly designate r as the stagnation temperature
p dA!., + (~:0 ) llx J dsp (10.4-4)
and not the temperature at point "o." The stagnation iemperature and
pressure are the temperature and pressure that a fluid particle would attain
S
if it were isentropically brought to rest. In the analysis of compressible ftows,
dividing by !lx and taking the limit it is often convenient to specify local conditions in terms of the stagnation
temperature and pressure and the Mach number; thus r and p 0 are especially
important quantities.
(10.4-5) We now have at our disposal the mass, momentum, and energy equations
along with the equation of state for a perfect gas, p = pRT. We may now
the bounding circular strip of the sol ve for the four unknowns (T, p, p, (v.,)). Noting that
velocity profiles and a uniform
yR
p), this result reduces to C'IJ=-- (10.4-13)
y-1
and
-A(~:) (10.4-6)
(v.,)1 = MlyRT (10.4-14)

Eq. 10.4-1b, Equation 10.4-12 can be used to express the local temperature as a function of
the local Mach number and the stagnation temperature.

-e:) (10.4-7)
T = To ( 1 + y ~ rl
1 M2 (10.4-15)

Three assumptions led to this result: enthalpy is a linear function of tem-


perature; the equation of state for the gas can be approximated by the
the energy balance given by Eq. perfect gas law; and the ftow is one-dimensional. Thus, the restriction to
obtain isentropic ftow has not yet been imposed.
Since y > 1.0, Eq. 10.4-15 indicates that the temperature decreases as
)v n dA= O (10.4-1) the velocity (and therefore the Mach number) increases. Intuitively we think
of this conversion as one of thermal energy to kinetic energy.
Sec. lOA lsentropic Nozzle Flow
Compressible Flow Chap. 10
-408
Using the perfect gas law (Eq. 10.3-
Our next step in the analysis requires the restriction of isentropic flows. of temperature and pressure, and rep
By a straightforward analysis, we can show that the temperature and
pressure in an isentropic process are related by

_p_ = constant (10.4-16)


p ; >-1
we may write the mass flow rate as
and we may use this result in conjunction with Eq. 10.4-15 to obtain an
expression for the pressure m= A~
p = po(1 + y~ 1 M2r1-y (10.4-17)
Substitution of T and p from Eqs. lO
This equation is only valid for isentropic flows. We will obtain an interesting
result if we use the binomial theorem4 to expand the term in parentheses.

p =Po[+ _Y_ Y-1M2+ 1


1 -y 2
~y ~y~:
2! 4
2
(y- 1) M4 + ...]
(10.4-18)
So far, we have determined the temp
terms of T 0 , p 0 , and M, and we nee
Making use of Eq. 10.4-14 and simplifying the terms somewhat, we obtain flow completely. Substitution of Eq.
result:
-
p- p
o[ - ~ (v,l +
2 RT
y(2y- 1 )M4
8
+ ...J (10.4-19)

Application of the perfect gas law allows us to obtain a form somewhat


reminiscent of Bernoulli's equation in the absence of gravitational effects Thus, T, p, m, and (v.,) are known a
and p 0 are generally specified, the
po[1 + y( 2 y - l) M 4
8
+ ...] (10.4-20) where in the duct to specify the flow ,

[ + ( ~ 1) JljY-1
back pressure in reservoir Il is spe1
= p + ! p < v., > 2 1 y M
2 necessary to determine the tempera1
tributions. We may do so by assumin
Let us imagine now that we have a free stream of gas at a pressure p to calculate the Mach number in t
moving at a velocity (v.,), which impinges on a Pitot tube such as that illus- determine the pressure distribution b~
trated in Fig. 8.5-7. It is not a bad assumption that the flow along the at the exit of the nozzle is different
stagnation streamline is isentropic ; thus, the pressure at the nose of the Pilot we assume a new value of mand rep
0
tube is the stagnation pressure, p 0 If the flow were incompressible p would Rather than doing so, we will ex
0
be given by specifying the pressure ratio pfp and
p0 = p + !P (v.,)2 (10.4-21) m/ A throughout the nozzle. By worl
unit area, mfA, we avoid specifying
However, Eq. 10.4-20 would certainly give a more accurate determination.
example we have assumed that the g~
This simple example provides us with an estima te of the effects of compressi-
is 300K. The results are shown in F
bility, for if M< t. Eqs. 10.4-20 and 10.4-21 will agree to within 1 per cent.
We want to think of the pressure
Returning now to the analysis of the nozzle, we designate the mass flow
pressure divided by the stagnation pre~
rate as m and write Eq. 10.4-1 b as
the. points on the curve represent the
m = p(v., )A (10.4-22) nozzle shown in Fig. 10.4-l. Let us ir
4. H. B. Dwight, Tables of Integrals and Other Mathematical Data (New York: The
Macmillan Company, 1961), p. l.
Sec. 10.4 lsentropic Nozzle Flow 409
Compressible Flow Chap. 10

the restriction of isentropic flows. Using the perfect gas law (Eq. 10.3-18) to express the density as a function
show that the temperature and of temperature and pressure, and representing the velocity as
by
(10.4-23)
( 10.4-16)
we may write the mass flow rateas
10.4-15 to obtain an

fy
(10.4-17)
m=Ap
viiT M (10.4-24)

flows. We will obtain an interesting Substitution of r and p from Eqs. 10.4-15 and 10.4-17 gives us
to expand the term in parentheses.
. f*'( + - -
m = Apo -
Rro
1
y - 1 2)( y+l) 2(1-y)
2
M M (10.4-25)

So far, we have determined the temperature, pressure, and mass fl.ow rate in
0
terms of ro, p , and M, and we need only specify the velocity to define the
the terms somewhat, we obtain flow completely. Substitution of Eq. 10.4-15 into Eq. 10.4-23 gives us this
result:
(10.4-19)
(10.4-26)
us to obtain a forro somewhat
absence of gravitational effects
Thus, r, p, m, and (v.,) are known as functions of p 0 , ro, and M. While ro
0

.. .] (10.4-20) and p are generally specified, the Mach number must be determined every-
where in the duct to specify the flow completely. For practica! purposes, the

( _2 1) M2
[1 +_Y_ ]1/Y-1 back pressure in reservoir 11 is specified and a trial-and-error solution is
necessary to determine the temperature, pressure, and Mach number dis-
tributions. We may do so by assuming a mass fl.ow rate and using Eq. 10.4-25
free stream of gas at a pressure p
to calculate the Mach number in the duct. Once we know M, we may
on a Pitot tube such as that illus-
determine the pressure distribution by Eq. 10.4-17. Ifthe calculated pressure
that the fl.ow along the
at the exit of the nozzle is different from the back pressure in reservoir 11,
the pressure at the nose of the Pilot
we assume a new value of m and repeat the calculation.
fl.ow were incompressible p 0 would
Rather than doing so, we will examine the results of these equations by
specifying the pressure ratio p/p 0 and determining the distribution of M and
(10.4-21)
m/ A throughout the nozzle. By working in terms of the mass flow rate per
give a more accurate determination. unit area, m/A, we avoid specifying the geometry of the nozzle. For this
estima te of the effects of compressi- example we have assumed that the gas is air and the stagnation temperature
4-21 will agree to within 1 per cent. is 300K. The results are shown in Fig. 10.4-3.
nozzle, we designate the mass flow We want to think of the pressure ratio p/p0 as being physically the back
pressure divided by the stagnation pressure. When we loo k at it in this fashion,
(10.4-22) the points on the curve represent the values of M and mfA at the exit of the
nozzle shown in Fig. 10.4-l. Let us imagine that the back pressure is initially
Mathematical Data (New York: The
-410 Compressible Flow Chap. 10 Sec. 10.4 lsentropic Nonle Flow

mlA

p/pO
Fig. 10.4-3. Variation of M and rii/A for isentropic ftow.

equal to the stagnation pressure and that no flow takes place between the
two reservoirs. In Fig. 10.4-3, then, p/p0 = 1 and M= m/ A =O. Now let
us lower the pressure in reservoir 11, which according to Fig. 10.4-3, causes
an increase in both the Mach number and mfA entirely in accord with our
intuition. When the pressure ratio reaches a value of about 0.5, the mass
flow rate per unit area reaches a maximum and begins to decrease. As the
back pressure tends to zero (p/p 0 - 0), the flow rate tends to zero while the
Mach number steadily increases to values greater than 10. This behavior is
certainly not in accord with our intuition, and we had better reserve judgment
on the validity of these curves.
Experimental evidence is in order here, and the results indicate that mfA flg. 10.4-4. Flow issui
is constant at the maximum value even though the back pressure is con- Photograph
tinually decreased. The Mach number at the exit of the nozzle is also Center, Va.
constant at a val1,1e of 1.0, and the flow is termed critica/. This situation is
similar to the critica! flow we encountered in the study of open channel flow
in that the flow rate is not controlled by the back pressure (downstream -----~--(al
conditions) for sonic (critical) or supersonic (supercritical) flows.
~--(b)
Figure 10.4-4 shows a Schlieren photograph of an air jet issuing from a ~......--(e)
convergent nozzle. The Mach number at the exit is 1.0 and the pressure at
the exit is 105 times the back pressure. Under these conditions, there is a
sudden expansion of the air stream as it lea ves the nozzle and the flow is no ---(d)
longer one-dimensional. lt has overexpanded and a shock wave has formed
in the jet downstream from the nozzle exit. The performance curves for a X

convergent nozzle are shown in Fig. 10.4-5 and indicate the constant flow
rate achieved for values of the pressure ratio less than the critica! value. Flg. 10.<4-5. Performance
Compressible Flow Chap. 10
Sec. 10.4 lsentropic Nozzle Flow 411
mlA

/A

m/A for isentropic flow.

no flow takes place between the


= 1 and M= m/A =O. Now let
according to Fig. 10.4-3, causes
and m/A entirely in accord with our
a value of about 0.5, the mass
and begins to decrease. As the
the ftow rate tends to zero while the
greater than 10. This behavior is
, and we had better reserve judgment

and the results indicate that mjA Fla. 10.4-4. Flow issuing from a converging nozzle.
though the back pressure is con- Photograph courtesy NASA Langley Research
at the exit of the nozzle is also Center, Va.
is termed critica/. This situation is
in the study of open channel flow
by the back pressure (downstream --4---(o)
(supercritical) flows.
~+--(b)
of an air jet issuing from a m t------.
-------(e)
at the exit is 1.0 and the pressure at A
Under these conditions, there is a
leaves the nozzle and the flow is no ----(d)
and a shock wave has formed
exit. The performance curves for a X
o 0.5
.4-5 and indicate the constant flow p/po
ratio less than the critica! value.
Fla. 10.4-5. Performance curves for a converging nozzle.
412 Compressible Flow Chap. 10 Sec. 10.4 lsentropic Nczzle Flow

Choke flow As we begin our progress throug


increases because of the decrease in
The condition of maximum ftow rate observed in a converging nozzle is because the decrease in area accelerat
called choke jlow, and all engineers who work with fluids should recognize pressure ratio, the Mach number is l.
its existence as a fact of life. lt would be possible for an engineer, given a are to continue along the performan
pressure ratio less than the critica! value, to "plug and chug" with the appro- mf A is going to decrease. In our pre
priate equations and calculate a mass flow rate for converging nozzle that is intuition told us that points on the E
less than the value occurring at the choke flow condition. were inadmissible. Experiments con
There are two other ways in which choke flow may occur. While these choke flow was introduced. In this e
mechanisms are too complex for us to analyze in this limited treatment of imagining an increased cross-sectio
compressible ftow, they deserve a comment. In a converging nozzle, we decrease in m/ A, but what about the M
accelerated the flow by decreasing the area of the duct; however, there are continuing to increase while the are2
other ways in which the flow can be accelerated provided the flow is opposition to our intuition, but if we
compressible. find the velocity does increase in t
Let us consider first the case of compressible, adiabatic flow in a long velocities are obtained.
duct of constant cross-sectional area. The decrease in pressure gives rise to
a decrease in the density, which in turn forces the velocity to increase. lf Subsonic and supersonic flow
the pressure drop is suffi.ciently large, the velocity may reach sonic velocity
at sorne point in the duct and choke flow will occur because the downstream The increasing velocity in the divet
conditions can no longer influence the flow rate. phenomenon which we need to analyz
Choke flow may also be brought about by heating the gas. Such heating density decreasing faster than the area
causes a decrease in density and an increase in velocity; thus, sonic velocity process in concrete terms. Starting
may be attained and a choked condition occurs. lt also follows that cooling carry out the differentiation and divi~
the gas m:ay prevent the occurrence of choke flow.
l(dp) + (v.,)1d.d
; dx
Converging-diverging nozzles
Differentiation of the energy balance 1
We now consider isentropic flow through the converging-diverging
dh
nozzle shown in Fig. 10.4-6. The equations derived at the beginning of this dx + (vz)
section apply to this system; however, in examining the performance curves
in Fig. 10.4-3 we will proceed in a which we may simplify further if we
slightly different fashion. In this case,
Reservoir O let us imagine that p 0 and PB are dh = T
Reservoir 1 Pressure =p8 fixed, thus fixing the mass flow rate.
Po T" The pressure ratio p/p0 decreases as For isentropic fiows, we may write
we proceed through the nozzle from
reservoir 1 to reservoir 11, and we ask (~~).-
ourselves the question, "Can we
specify a value of A for each value Substitution into Eq. 10.4-28 yields a
of p/p0 such that both the perfor- equation:
mance curves in Fig. 10.4-3 and our (vz) (d(v"'))
Flr. 10.4-6. Converging-diverging nozzle. intuition are satisfied ?" dx s
Compressible Flow Chap. 10
Sec. 10.4 lsentropic Nczzle Flow
413

As we begin our progress through the nozzle, p/p0 decreases and m/A
increases because of the decrease in area. The Mach number also increases
observed in a converging nozzle is because the decrease in area accelerates the flow. When we reach the critica!
work with fluids should recognize pressure ratio, the Mach number is l.O and m/ A is a maximum. Now, if we
be possible for an engineer, given a are to continue along the performance curve, the area A must increase if
to "plug and chug" with the appro- m/A is going to decrease. In our previous example of a converging nozzle,
rate for converging nozzle that is intuition told us that points on the performance curve for supersonic flow
flow condition. were inadmissible. Experiments confirmed our intuition, and the idea of
choke flow may occur. While these choke flow was introduced. In this case intuition does not restrict us from
analyze in this limited treatment of imagining an increased cross-sectional area giving rise to the predicted
In a converging nozzle, we decrease in m/ A, but what about the Mach number (and therefore the velocity)
area of the duct; however, there are continuing to increase while the area increases? This process is surely in
accelerated provided the flow is opposition to our intuition, but if we again turn to experimental evidence we
find the velocity does increase in the diverging section and supersonic
~mpressit>Ie,
adiabatic flow in a long velocities are obtained.
decrease in pressure gives rise to
forces the velocity to increase. If
Subsonic and supersonic flow
velocity may reach sonic velocity
will occur because the downstream
rate. The increasing velocity in the diverging portion of the nozzle is a curious
by heating the gas. Such heating phenomenon which we need to analyze further. Certainly it results from the
in velocity; thus, sonic velocity density decreasing faster than the area increases, but we need to examine the
occurs. lt also follows that cooling process in concrete terms. Starting with the mass balance (Eq. 10.4-1a), we
flow. carry out the differentiation and divide by p<v,.>A to obtain

l(dp\ + _1 d(v.,) +!_dA= 0 (10.4-27)


p dx} (v.,) dx A dx
Differentiation of the energy balance (Eq. 10.4-9) yields
through the converging-diverging
derived at the beginning of this dh + (v.,) d(v.,) = 0 (10.4-28)
examining the performance curves dx dx
Fig. 10.4-3 we will proceed in a which we may simplify further if we use the thermodynamic relationship
different fashion. In this case,
us imagine that p0 and Pn are dh = Tds +! dp
thus fixing the mass flow rate. (10.4-29)
p
pressure ratio p/p0 decreases as For isentropic flows, we may write
proceed through the nozzle from
rp ~,,.rvm'r I to reservoir 11, and we ask
the question, "Can we (~~).=~e~). (10.4-30)

a value of A for each value Substitution into Eq. 10.4-28 yields an interesting special form of the energy
0
p/p such that both the perfor- equation:
curves in Fig. 10.4-3 and our
are satisfied ?"
(v.,) (d(v.,))
dx s
= _l(dp)
p dx .
(10.4-31)
Compressible Flow Chap. 10 S.C. 10.4 lsentrople Nozzle Flow

This result is identical to the momentum equation (Eq. 10.4-7) ifthe entropy incrbses with increasing cross-sectio
is constant. Thus, one-dimensional isentropic fiows may be solved by means cross-sectional area. We must keep i
of the mass and total energy equations, the momentum equation yielding no isentropic fiow that may not occur i
new information. Directing our attention now to Eq. 10.4-27, we specify ducts. The other fiow properties a
that the fiow be isentropic and write the density gradient as, changes from subsonic to supersonic;
In studying the performance cu

(~:). = (:;).(~:). (10.4-32) tinuously decreasing pressure ratio p


that lead to a converging-diverging
and use Eq. 10.4-31 to eliminate the pressure gradient results for isentropic fiow from still
~ that A(x) is speci.fied so that the no

( dx}.
~\ = -p(OP\ (v.,>(d(v.,)) (10.4-33) 10.4-6. If we specify the stagnation ~
op}. dx derived equations to determine the 1
starting at reservoir 1 where p = p 0
Substitution of this result into Eq. 10.4-27, and use of Eq. 10.3-14 to express
(opfop). in terms of the speed of sound yield p = PB Such calculations lead to

- .!.2 + _1_)2 (v.,) d(v.,) + .!_dA = 0 (10.4-34)


Throot
( c (v.,) dx A dx
Expressing this result in terms of the Mach number, and rearranging, we
obtain
d(v.,) = - [~ (1 - M~J-1 (10.4-35)
dA (v.,)
From this result we see that if M < 1, the velocity increases if A decreases-
i.e., the fiow is accelerated in a converging section and decelerated in a
diverging section. However, if the fiow is supersonic, M > 1 and the velocity

Table 10.4-1
SUBSONIC ANO SUPERSONIC FLOW IN
CONVERGINO ANO DIVEROING CONDUITS

o
. 1. 1~
0 1vergmg
Fl1. 10.4-7. Pressure ratio in

first two curves, a and b, are for Iow


always less than 1.0. Here, the nozzl
Parameter Subsonic Supersonic Subsonic Supersonic being low where the velocity is high
the diverging section. The pressure
Velocity increase decrease decrease increase downstream cross-sectional area is s
Pressure decrease increase increase decrease (a), (b), and (e) indicate increasing
Temperature decrease increase increase decrease causing a Mach number of 1.0 to be
Density decrease increase increase decrease
Entropy constant constant constant constant gives the maximum flow rate for the
and the solution for the pressure is
Compressible Flow Chap. 10 S.C. 10.4 lsentropk Nozzle Flow 415

equation (Eq. 10.4-7) if the entropy incrbses with increasing cross-sectional area and decreases with decreasing
flows may be so1ved by means cross-sectional area. We must keep in mind that these conclusions hold for
the momentum equation yielding no isentropic flow that may not occur in long reaches of converging-diverging
now to Eq. 10.4-27, we specify ducts. The other flow properties also behave differently when the flow
density gradient as, changes from subsonic to supersonic; these are indicated in Table 10.4-1.
In studying the performance curves in Fig. 10.4-3, we specified a con-
(10.4-32) tinuously decreasing pressure ratio p/p0 , and found that values of A(x) exist
that lead to a converging-diverging nozzle. We now wish to look at the
results for isentropic flow from still another point of view. Let us assume
that A(x) is specijied so that the nozzle resembles the one shown in Fig.
(10.4-33) 10.4-6. If we specify the stagnation properties, A(x), and m, we can use the
derived equations to determine the pressure at every point in the nozzle,
, and use of Eq. 10.3-14 to express starting at reservoir 1 where p = p 0 and continuing to reservoir 11 where
yield p = p 8 Such calculations lead to the curves shown in Fig. 10.4-7. The

1 dA Throat
+--=0 (10.4-34)
Adx
number, and rearranging, we
Ps=.q,
(10.4-35) Ps=Pc

velocity increases if A decreases-


section and decelerated in a
Forbidden
supersonic, M > 1 and the velocity region

0
r.l~
X

. Fl1. 10.4-7. Pressure ratio in a converging-diverging nozzle.


D 1vergmg

first two curves, a and b, are for low flow rates where the Mach number is
always less than 1.0. Here, the nozzle acts like a Venturi meter, the pressure
being low where the velocity is high and a pressure recovery taking place in
the diverging section. The pressure recovery is incomplete because the
decrease increase downstream cross-sectional area is smaller than the upstream area. Curves
increase decrease (a), (b), and (e) indicate increasing mass flow rates with the latter flow
increase decrease
increase decrease causing a Mach number of 1.0 to be reached at the throat. This condition
constant constant gives the maximum flow rate for the nozzle at these stagnation conditions,
and the solution for the pressure is double valued and may follow either
416 Compressible Flow Chap. 10 Sec. 10.5 Shock Waves

curve (e) or curve (d). In actual practice, the pressure distribution at this
flow rate would be determined by whether the back pressure were equal to
Peor Pd
But what of the "forbidden region" between Pe and Pa? Our analysis
indicates that this condition simply cannot be reached, yet we know that
experimentally we may adjust the back pressure to take on values between
Pe and Pd Once again, experimental evidence is in order, we find that if the
back pressure is adjusted to sorne value between Pe and pd, a shock wave
develops and there is a sudden transition from supersonic to subsonic flow
accompanied by an increase in entropy. The pressure distributions for such
flows are indicated in Fig. 10.4-8. This phenomenon is similar to the sudden

Throot
Flg. 10.5-1. Control
of a shock wave.

increases. We will now reformulate


equations for the control volume sho
shock wave) without imposing the res
that in the previous section the mo
specify the flow; however, when we
that the momentum equation is requt

Mass balance

Since p(v.,)A is a constant, the m


o X

Fig. 10.4-8. Formation of shock waves in a converging-diverging nozzle.

transition that occurs at a hydraulic jump in open channel flow. Note that or by Eq. 10.4-24 we may write
these shock waves are advancing into the flowing stream at speeds greater
than the speed of sound just as the hydraulic jump advances into a flowing M (Y"_
stream at speeds greater than Jih-i.e., the velocity of propagation for a Pt \) "R'T:
small disturbance.
Momentum balance
10.5 Shock Waves
To obtain the momentum balan
In attacking this problem, we will make use of the experimental obser- x 1 to x 2 to obtain
vation that shock waves occur over very thin regionst and show that a transi- m( (v.,)2 - <v.,)
tion from supersonic to subsonic flow is permissible, provided the entropy
t The thickness of a shock wave depends on the Mach number, temperature, and Here we have assumed that the sho
pressure and is on the order of 10- in. (see Ref. 1, Vol. 1, p. 134). be neglected relative to variations in
Compressible flow Chap. 10 Sec. 10.5 Shock Waves 417

the pressure distribution at this


the back pressure were equal to

" between Pe and pd? Our analysis


ot be reached, yet we know that
pressure to take on values between
is in order, we find that if the
between Pe and pd, a shock wave
from supersonic to subsonic flow
The pressure distributions for such
phenomenon is similar to the sudden

Flg. 10.5-1. Control volume for the analysis


of a shock wave.

increases. We will now reformulate the mass, momentum, and energy


equations for the control volume shown in Fig. 10.5-1 (which contains the
shock wave) without imposing the restriction of constant entropy. Remember
that in the previous section the momentum equation was not required to
specify the flow; however, when we allow the entropy to vary we will find
that the momentum equation is required.

Mass balance

Since p(v.)A is a constant, the mass balance simply reduces to


in a converging-diverging nozzle. (10.5-1)
in open channel flow. Note that or by Eq. 10.4-24 we may write
the flowing stream at speeds greater
ic jump advances into a flowing M [Y'_ M Jy
, the velocity of propagation for a P1 1\} R"J; - P2 2\} ~ (10.5-2)

Momentum balance

To obtain the momentum balance, we simply integrate Eq. 10.4-6 from


use of the experimental obser- x 1 to x 2 to obtain
thin regionst and show that a transi-
(10.5-3)
is permissible, provided the entropy
on the Mach number, temperature, and Here we have assumed that the shock is so thin that variations in A(x) may
. 1, Vol. 1, p. 134). be neglected relative to variations in (v.,) and p .
<418 Compressible Flow Chap. 10 Problems

Energy balance Substitution of this result into Eq. 10.5

s2 - s1 =(el>- iR) In
We derived the energy balance in the previous section without the
restriction of constant entropy, and we need only note that the temperature We may now use Eq. 10.5-4 to express
is given by Eq. 10.4-15; thus, the ratio of temperatures is Mach numbers, and we obtain
1 +y-1M2
2 1+
T2 1
s2 s1 = (cP- iR) ln -
(10.5-4) -
T 1 +y-1M2 (
1+
2 2
Entropy change Now, the second law ofthermodynamio
process; therefore, Eq. 10.5-12 indicatt
Starting with Eq. 10.2-10, we can quickly derive the thermodynamic M 2 =M,
relationship or
1
dh = Tds dp +- (10.5-5) M 2 <M,
p Our analysis indicates that only transit
Rearranging and integrating ds between x 1 and x 2 gives are possible. The analysis does not in
2 2 2 nor does it indica te that the transition si

s2 - S = f f f d~
1
ds =
1
d: -
1 p
(10.5-6)
1t only states that such a transition is
increases.
From the analysis of isentropic fl
Restricting our ana1ysis to fluids that may be treated as perfect gases, we may conditions (i.e., a back pressure in tH
carry out the integration to obtain the possibility of an isentropic flow.
might ask, "Why isn't there a smooth
s2 - s1 =el> In (i) - R In (~) (10.5-7) sonic to subsonic flow ?" The answer
able at present. We can only state that
The ratio T 2/T1 is already available in Eq. 10.5-4, and we need only find flow occur abruptly.
f2fp 1 to determine the entropy change across the shock. It is easiest to do so
by means of the mass balance and the perfect gas law. Rearranging Eq.
PR.OB
10.5-1 and neglecting the change in area from x 1 to x 2 , we obtain
P2 (v.,)t 10-l. Integrate Eq. 10.1-19 with respect
-=-- (10.5-8) the first law of thermodynamics
P1 (v.,)2
meaning of each term.
Expression of the velocities in terms of the Mach number (Eq. 10.4-23)
10-2. Starting with Eq. 10.2-15, derive
before and after the shock wave yields for a control volume, "Ya(t).
112 10-3. Derive Eq. 10.4-7 by integrating ti
-P2 = M1(T1) (10.5-9)
p1 M2 T2 cross-sectional area A(x)-i.e.,
and the perfect gas law may be used to express the pressure ratio as
12
f
A(z)
i p [:; JdA
+ v Vv = - j
A(
P2 = P2T2 = M1(T2) 1 (10.5-10)
p1 p1T1 M2 T1 t Thus, we could have the situation M
Compressible Flow Chap. 10 Problems 419

Substitution of this result into Eq. 10.5-7 and rearrangement give

previous section without the s2 - s1 = (e~ - iR) In (~) - R In (~:) (10.5-11)


need only note that the temperature
of tempera tu res is We may now use Eq. 10.5-4 to express the temperature ratio in terms of the
Mach numbers, and we obtain

(10.5-4) s2- s1 =(e~- iR) In(+~ M:) -R In (~1) (10.5-12)


1 r___! M2 2
+ 2 2

Now, the second law ofthermodynamics tells us that lls 2 Ofor any adiabatic
process; therefore, Eq. 10.5-12 indicates either
derive the thermodynamic
M 2 =M, lls =O
or
(10.5-5) M 2 < M1 , lls > O
Our analysis indicates that only transitions that decrease the Mach number
x 1 and x 2 gives
are possible. The analysis does not indicate that such transitions do occur,

J -f
2 2 nor does it indica te that the transition should be from supersonic to subsonic.t
dh dp (10.5-6) lt only states that such a transition is possible, and if it occurs, the entropy
T pT increases.
1 1
From the analysis of isentropic fiows we know that there are certain
be treated as perfect gases, we may conditions (i.e., a back pressure in the "forbidden region") which disallow
the possibility of an isentropic fiow. With this information in hand, one
might ask, "Why isn't there a smooth, nonisentropic transition from super-
(10.5-7)
sonic to subsonic fiow?" The answer to this question is apparently unavail-
Eq. 10.5-4, and we need only find able at present. We can only state that transitions from supersonic to subsonic
the shock. lt is easiest to do so fiow occur abruptly.
perfect gas law. Rearranging Eq.
from x 1 to x 2 , we obtain PROBLEMS

(10.5-8) 10-l. Integrate Eq. 10.1-19 with respect to time to obtain the traditional form of
the first law of thermodynamics for open systems. Carefully explain the
meaning of each term.
10.4-23)
10-2. Starting with Eq. 10.2-15, derive the general macroscopic entropy balance
for a control volume, '"~'"a(t).
(10.5-9) 10-3. Derive Eq. 10.4-7 by integrating the x-direction equation of motion over the
cross-sectional area A(x)-i.e.,
the pressure ratio as
J
AW
i p [:; + v VvJdA = - Ji
AW
Vp dA + Ji pg dA
AW
+ p. J
i V 2v dA
AW
(10.5-10)
t Thus, we could have the situation M 1 > M 2 > 1.
420 Compressible Flow Chap. 10

Make the same simplifying assumptions as in Sec. 10.4, and use the Leibnitz
rule (see Prob. 3-5) for interchanging integration and differentiation.
10-4. For a given value of T o, show that an upper limit exists for the velocity
given by

(v., ) < ..J~


y-=-} ~ Flow Around
10-5. A perfect gas at T1 and p in a large reservoir flows isentropically through a
convergent nozzle of area A2 into the atmosphere where the pressure is p 0 lmmersed Bodies
Derive expressions for the mass flow rate mtaking into account the phenom-
enon of choke flow. (This type of calculation would be a key step in the
design of a safety valve for a high-pressure chemical reactor.)
10-6. If ata shock such as that shown in Fig. 10.5-1, M 1 = 3.0, T1 = 1000R and
p 1 = 0.2 atm, find the Mach number and pressure after the shock. Assume
1' = 1.5.

10-7. For isentropic flow in a divergent channel, the density decreases 14 per cent
(p2 = 0.86p1 ) for a 10 per cent increase in the cross-sectional area (A 2 =
1.10 A1). Calculate the change in velocity between these two points.
10-8. A stream of air flows ata Mach number of 0.8. lf a Pitot tube is used to
measure the velocity, compute the error incurred by the use of the incom-
pressible form of Bernoulli's equation.
10-9. Air ftowing in a duct enters a nozzle at a velocity of 5000 ft/sec, a pres-
sure of 100 psia, and a temperature of 800F. The nozzle is designed to The objectives of this chapter are to
increase the pressure isentropically to 125 psia. If the mass ftow rate is 1O ena associated with ftow around imm
lbm/sec, what is the downstream cross sectional area of the nozzle? a rational explanation of the drag
10-1 O. Air is contained in a tan k at 200 psia and 7rF. lf the flow of a ir through the shown in Fig. 8.2-6 in Chap. 8. Bou
outlet valve is approximated as isentropic and one-dimensional, what is the thus giving sorne practice in setting U
maximum flow rate into the surrounding atmosphere? Take the cross- balances. We shall also use boundary
sectional area at the valve as 1.0 in. 2 and the ambient pressure as 14.7 psia. performing an order of magnitude am
What would the ftow rate be if the pressure in the tank were increased to
400 psia?
11.1 Descri ption of Flow

The study of ftow around imme


subject to discuss with any degree of
the fact that there are three distinct,
inftuence the force exerted by the ftui
the ftow past a thin airfoil at a zero a
number. At high Reynolds numbers,
except in a thin region close to the
outside this thin region is governed a
called irrotational. As the name impli
4
Compressible Flow Chap. 10

as in Sec. 10.4, and use the Leibnitz


integration and differentiation.
an upper limit exists for the velocity

Flow Around
reservoir flows isentropically through a
atmosphere where the pressure is p 0 Immersed Bodies 11
rate mtaking into account the phenom-
calculation would be a key step in the
chemical reactor.)
10.5-1, M 1 = 3.0, T1 = 1000R and
and pressure after the shock. Assume

the density decreases 14 per cent


in the cross-sectional area (A 2 =
between these two points.
of 0.8. If a Pitot tube is used to
error incurred by the use of the incom-

at a velocity of 5000 ft/sec, a pres-


of 800F. The nozzle is designed to
to 125 psia. lf the mass ftow rate is 10 The objectives of this chapter are to introduce the student to the phenom-
sectional area of the nozzle? ena associated with ftow around immersed bodies and to provide him with
a rational explanation of the drag coefficient-Reynolds number curves
and nF. If the flow of air through the
shown in Fig. 8.2-6 in Chap. 8. Boundary layer theory will be introduced,
and one-dimensional, what is the
thus giving sorne practice in setting up and solving differential-macroscopic
atmosphere? Take the cross-
and the ambient pressure as 14.7 psia. balances. We shall also use boundary layer theory to explore the method of
pressure in the tank were increased to performing an order of magnitude analysis.

11.1 Description of Flow

The study of ftow around immersed bodies is an enormously difficult


subject to discuss with any degree of exactness. The difficulty results from
the fact that there are three distinct, and fairly complex regions of ftow that
inftuence the force exerted by the fluid on the body. Figure ll.l-1 illustrates
the flow past a thin airfoil at a zero angle of inclination and a high Reynolds
number. At high Reynolds numbers, inertial effects predominate everywhere
except in a thin region close to the surface of the solid body. The ftow
outside this thin region is governed almost entirely by inertial effects and is
called irrotational. As the name implies, fluid elements in this region undergo
-421
Flow Around lmmersed Bodies Chap. 11 Sec. 11.2 The Suddenly Accelerated Flat Plac

Fluid extends

-oo~--c=========~
t sO, the plote is stoti
t >O, the veiocity of ti
Fl. 11.2-1. The suddenly acc
lrrotationol
the area of fluid mechanics. The systern
Fl. 11.1-1. Flow regimes around an immersed body. Fig. 11.2-1.
We start with the two-dimensiona1

translation but do not rotate; therefore, there is no shear deformation and . p-+v-+v--
(ov~ OV~ ov.,)
viscous effects are negligible. A great deal is known about irrotational flow, ot ~ ax ay - 11

because in the absence of viscous effects the equations of motion can be


solved by fairly simple mathematical techniques. While the study of ir-
ovtl ovtl ov") = - 1
p (- + V~- + V
rotational flow is well within the student's capabilities, it will be treated only
at ax ay
11 -
,
qualitatively here, and we will concentrate our attention on the boundary and the continuity equation
layer. The wake is extremely difficult to treat theoretically, and again we will OV~ +~
restrict ourselves to qualitative comments regarding this region. ox
Near the solid surface, viscous effects become important, and this flow
Since the plate is infinite in the x-directi
region is called the boundary layer. The equations of motion here reduce to
and p are not functions of x. The conl
what are known as the Prandtl boundary layer equations. In the next
to
section, we willlay the ground work for both the derivation and the solution
of these equations. ovtl =
oy
and application of the boundary condi1
11.2 The Suddenly Accelerated Flat Plate B.C. 1: V"= O,
leads us to the conclusion that V11 is zero
As an introduction to the analysis of boundary layers we will examine the the equations of motion reduce to
fluid motion that results when an infinitely wide, infinitely long, flat p1ate is
suddenly accelerated from rest to a constant velocity, u0 We may rightly ov.,)
ask, "Why study a hypothetical flow that can never be realized in practice ?"
p (a =
The answer is that by performing the analysis we will gain intuition-i.e.,
we may improve our skill at guessing the flow topology. In addition, the o=-(~~
equations we will solve are identical to those describing transient heat and
mass transfer into a semi-infinite medium; thus, the analysis has value beyond We need solve only the first ofthese te
Flow Around lmmersed Bodies Chap. 11 Sec. 11.2 The Suddenly Accelerated Flat Plate -423

Fluid extends to y =oo

-oo~--~c:::::::::::::::~:::I::::x::~:::Jr---~+oo
t sO, the piote is stotionory
t >O, the velocity of the plote is o constont, u0
Flg. 11.2-1. The suddenly accelerated flat plate.

the area of fluid mechanics. The system under consideration is illustrated in


Fig. 11.2-1. .
We start with the two-dimensional fonp. of the Navier-Stokes equations
there is no shear deformation and
(otav., + v-+v- . op (o 2v.,2 02v.,)
ov., 11 ov.,) --'--+,u-+- (11.2-1)
is known about irrotational flow,
p--
"' ox oy - ax ox oy2
the equations of motion can be
techniques. While the study of ir- p(avti + v., ovti + vti ovti) = - op- pg + .u(J2v112+ J2v11) (11.2-2)
capabilities, it will be treated only ot ox oy oy ox oy2
our attention on the boundary and the continuity equation
treat theoretically, and again we will
regarding this region. (11.2-3)
become important, and this flow
equations of motion here reduce to Since the plate is infinite in the x-direction, we can argue logically that v.,, v 1,
layer equations. In the next and p are not functions of x. The continuity equation immediately reduces
both the derivation and the solution to
ovti =o (11.2-4)
oy
and application of the boundary condition
Flat Plate
B.C. 1 : 1
v = O, y= O (11.2-5)

boundary layers we will examine the 11


leads us to the conclusion that v is zero everywhere. Under these conditions,
the equations of motion reduce to
wide, infinitely long, flat plate is
eonsta111t velocity, u0 We may rightly 2
P(ov.,) =,u (o v.,) (11.2-6)
can never be realized in practice ?" ot ol
analysis we will gain intuition-i.e.,
the flow topology. In addition, the
those describing transient heat and
o=-(~~)- pg (11.2-7)

; thus, the analysis has value beyond We need solve only the first ofthese to determine the velocity as a function
Flow Around lmmersed Bodies Chap. 11 Sec. 11.2 The Suddenly Accelerated Flat Plat

of y and t. Defining the dirnensionless velocity Uz as At this point we can see that setting a
sirnplifying our equation. Substituting
U = Vz (11.2-8)
z
Uo
setting a = 1, we get
we rnay write Eq. 11.2-6 in the forrn
(~~z)bytb-1 =
oUZ ='V
ot ol
(o Uz)
2
(11.2-9)
If we could sornehow elirninate y and
which we rnust solve subject to the following boundary conditions would be reduced to solving an ord
B.C. 2: Uz =O, t =O, O ::;;y::;; oo (11.2-10) accornplish this by setting b = -t, ther
B.C. 3: Uz = 1, t >O, y= O (11.2-11) 1] = yt
B.C. 4: Uz =O, y= oo, t O (11.2-12) Equation 11.2-16 reduces to

The boundary condition B.C. 2 is often referred toas an initial condition since it (:~x) = - ~
specifies the velocity field at sorne initial time. The boundary condition B.C. 3
indicates that for times greater than zero the plate is rnoving at a constant veloc- Putting the boundary conditions in
ity, thus an infinite acceleration is required. While this is physically irnpossible, B.C. 2': Uz= O,
it provides us with sorne insight as to how disturbances originating at salid
surfaces are propogated into the surrounding fluid. B.C. 3': uz =1,
Equation 11.2-9 is a partial differential equation. Until now such B.C.4' : Uz= O,
equations have been considered beyond the scope of this text; however, we
can solve this one without expending too rnuch effort, and in the process we
Here we see that boundary conditions E
can becorne acquainted with a fairly general rnethod of solving partial
terrns of the transformation of variables h
differential equations. A solution rnay be obtained by rneans of Laplace
transforrns, but here we will use a technique known as a similarity solution. We solve the differential equation t
Thus, we define a new variable 1J in terrns of powers of y and t,
P= -~
(11.2-13) ~
where a and b are as yet undeterrnined constants, and we seek a solution of where
the forrn P=
Uz(y, 1) = Uz[1J(y, !)] (11.2-14)
If we find that Uz can, indeed, be represented in terrns of the single variable Separating variables and integrating gi1
rJ, we will have been successful. lf we cannot represent Uz in terrns of 1], we
rnust try other rnethods of solution. lnP =-
We note first that the two partial differentials in Eq. 11.2-9 rnay be written
or
oUz = (dUz) 01J = (dUz) byatb-1 (11.2-15a) au.,_ e
ot drJ ot d11 dr2 - 1

oUz = d Uz (01])
2 2 2
+ dUz (d 1J)
2

Forrnation of the indefinite integral giv


ol drJ 2
ay drJ ol
2 (11.2-15b)
d Uz
=-(aya- 1 b 2
t) + dUz
-[a(a- 1)y a-2 b
t]
d1] 2 d1J
Flow Around lmmersed Bodies Chap. 11
Sec. 11.2 The Suddenly Accelerated Flat Plate 425
velocity U., as
At this point we can see that setting a = 1 will eliminate one term, thus
simplifying our equation. Substituting Eqs. 11.2-15 into Eq. 11.2-9, and
(11.2-8)
setting a = 1, we get
2
dU"')bytb-t =
( dr 11 (ddrU"')t
2
2
b (11.2-16)
(11.2-9)
If we could somehow eliminate y and t from this equation, our problem
boundary conditions
would be reduced to solving an ordinary differential equation. We can
o :S:: y :S:: 00 (11.2-10) accomplish this by setting b = -!, thereby defining r as,
y=O (11.2-11) r = yr-t /2 (11.2-17)
tO (11.2-12) Equation 11.2-16 reduces to

rred to as an initial condition since it


dUx) = _ 2v(d
( dr
2
U.,) (11.2-18)
r dr 2
time. The boundary condition B.C. 3
the plate is moving at a constant veloc- Putting the boundary conditions in terms of r, we find
While this is physically impossible,
B.C. 2': U.,=O, r = 00 (11.2-19)
how disturbances originating at solid
fluid. B.C. 3': U.,= 1, r=O (11.2-20)
equation. Until now such B.C.4': U.,=O, r = 00 (11.2-21)
the scope of this text; however, we
much effort, and in the process we
general method of solving partial Here we see that boundary conditions B.C. 2' and B.C. 4' become identical in
be obtained by means of Laplace terms of the transformation of variables from y and t to 11
known as a similarity so/ution. We solve the differential equation by first reducing the order to obtain
of powers of y and t,
(11.2-13)
p = _ 2v(dP) (11.2-22)
r dr
where
P=dU.,
(11.2-14) dr
L""~~~~-u in terms of the single variable
represent U., in terms of r, we Separating variables and integrating gives
2

trerentals in Eq. 11.2-9 may be written


In P = - !l._ + C1
4v
or
(~~"') byatl>-
1
(11.2-15a) dU.,- C' e
-"14
- (11.2-23)
d172 - 1

Formation of the indefinite integral gives


(11.2-15b)
f
~
dU.,
-[a(a - 1)ya-2 tb )
dr u"' = e; e-r"t 4 d-r + c2 (11.2-24)
o
Sec. ll.l The Suddenly Accelerated Flat Plate
Flow Around lmmened Boclia Chap.

where -r is the dummy variable of integration. It will be convenient to define


a new dummy variable Eby
T
E=- ---A~
#
allowing us to write Eq. 11.2-24 as
~rv4.
u;l: = c~J e-~~ dE+ C 2 (11.2-25)
u.
o
By boundary condition 3', C2 = 1, and application of boundary condition 2'
leads to the result

(11.2-26) o
Flg. 11.1-1. Velocity profil
accelerated ftat plate.

The definite integral in Eq. 11.2-26 is known to be equa1 to .;;.2, and our
expression for the velocity becomes
Approximate solution
~v'4v

U:1:= 1 -2- f e-~ 2 dE (11.2-27)


J;,o We will now solve this problem agai
compare the results with the exact s~h
The latter term in this equation is known as the error function, abbreviated
"Why solve a problem wit~ an approxtm
erf(r/../4v). Tabulated values of the error fl,lnction are available, 1 and our is already available-espectally when ~he
final expression is any real situation ?" Aside fr?m keepmg
that by comparing an approxtmate result
U;l: = 1- erf ~ (11.2-28)
'Y 4vt sorne insight into the validity o~ th.e
This result is plotted in Fig. 11.2-2. The key point of this analysis is that U,. it provides an interesting apphcatton
reaches 99 per cent of the undisturbed value for y/../ 4vt = 1.8, and we can balance.
The control volume for the approx1
use this fact to define a "boundary layer" thickness () as
l1.2-3. The upper boundary is located
() = 3.6J;i (11.2-29) y-direction. We start w~th the complet
For values of y larger than (), the velocity is, for all practica} purposes, zero. and dot it with i to obtam
We may also interpret Eq. 11.2-29 by saying that the distance the disturbancc
penetrates into the quiescent fluid is proportional to the square root of time t,
and the kinematic viscosity, v. This result agrees with our intuition, which
~~ f pvz dV +f pvz(v
f 0
(1) A,W
-1
should tell us that a large viscosity leads toa rapid propagation of a disturbancc,
and a zero value of the viscosity would not allow the disturbance to be fclt We again impose the restriction that the v
anywhere in the fluid. of x, and make the approximation (A) t
1. M. Abramowitz and l. A. Stegun, eds., Handbook of Mathematical FunctiolfS
A. 1:
(Washington, D.C.: National Bureau of Standards, 1964).
Flow Around lmmersed Bodia Chap. 1 , Sec. 11.2 The Suddenly Accelerated Flat Plate 427

1---'....----i---- Exact solution t-----i


- - - Approximate solution

(11.2-25)

application of boundary condition 2'

o
(11.2-26) o

Fls. 11.2-l. Velocity profiles for the suddenly


accelerated flat plate.
known to be equal to J;2, and our
Approximate solution
(11.2-27)
We will now solve this problem again, by an approximate method, and
as the error function, abbreviated compare the results with the exact solution. The student might well ask,
ft,mction are available, 1 and our "Why solve a problem with an approximate method when the exact solution
is already available--especially when the problem is somewhat removed from
any real situation ?" Aside from keeping idle minds busy, the answer must be
(11.2-28) that by comparing an approximate result with the exact solution, we can gain
sorne insight into the validity of the approximate method. In addition,
key point of this analysis is that U. it provides an interesting application of the macroscopic momentum
value for y/../ 4vt = 1.8, and we ca: balance.
thickness ~ as The control volume for the approximate solution is illustrated in Fig.
11.2-3. The upper boundary is located at y = ~(t) and is moving in the
(11.2-29) y-direction. We start with the complete macroscopic momentum balance
is, for all practica! purposes, zero. and dot it with i to obtain
that the distance the disturbance
to the square root of time t,
agrees with our intuition, which ~~ Jpv., dV +Jpv.,(v - w) n dA = J i t<n> dA (11.2-30)
a rapid propagation of a disturbance, f.w A.w -"'.w
not allow the disturbance to be felt We again impose the restriction that the velocity and pressure are independent
of x, and make the approximation (A) that
Handbook of Mathematical FunctioiU
1964). A. 1: v., =o, y ;;>8(t) (11.2-31)
428 Flow Around lmmersed Bodies Chap. JI Sec. 11.2 The Suddenly Accelerated Flat Plate

r:---------~ We now make the assumption tha1


allowing us to express the velocity as

L.
6(f) 1 1
A.2: v.,(y, t) = v,
J 1 1 Note that this equation is comparable to
-oo-1 ~ --; --+00
1 1 7} = .l
1 1 !5(
"----L ----.4
We further assume that the functional
Flg. 11.2-l. Control volume for the suddenly accelerated ftat plate.
represented by a polynomial.

Of course, we know this approximation is very reasonable on the basis of the


exact solution. A little thought will indicate that all the momentum flux
A.3: v., = a + b(~) + ce
terms are zero, and Eq. 11.2-30 reduces to To determine the constants in the po .
information regarding the velocity profil
L L

dt ff pv., dy dx =f [r
6(1)

E_ 11 .,, - r ., Jdx
11 (11.2-32) 1
v,.= O, y = d(t) (an intuit1
11=6(t) 11=0
o o o ov.. _ o y= c5(t) (continuit
Substitution of Newton's law of viscosity for r 11., yields
oy- ,
6 1 v,. = u 0, y=O (a bonafi
d JLJ d d JL{(ov.,). (ov.,). }d
dt- o o pVz y X = ..t o -oy
<> - - These three equations allow us to deten
1/=6(1) oy v- o X (11.2-33)
Eq. 11.2-37. We could obtain more
derivatives of v,. to be zero at y = !5; h1
By approximation 1, we have specified v., = O for y ~ d(t). If we require the
regarding the velocity profil~ ca~ be ?bu
derivative (ov.,joy)
to be a continuous function, then approximation 1 Examining Eq. 11.2-1, keepmg m nund
naturally leads us to the condition,t
we quickly conclude that

(!:"') =o, y= d(t) (11.2-34)


2

oyz- o,
0 Vz - y =o (from the

and Eq. 11.2-33 reduces to Equations 11.2-38 may be used to deter:


6(1) obtain
~ J pv., dy =
dr o
-..t (ov.,)
oy 11= 0
(11.2-35) V =U
..
[1- ~(~6.
o 2
Defining dimensionless variables as
Here we have used the fact that v., is independent of x ; thus, the integration
with respect to x can be dropped. This result states that, the force exerted v.,
U,.=- an
by the plate on the fluid is equal to the time rate of change of the momentum Uo
of the fluid.
we see that the velocity is given by
tOne can also argue that all higher derivatives are zero at y = <S(t). u.. = 1- !l
Flow Around lmmersed Bodies Chap. 11 Sec. 11.2 The Suddenly Accelerated Flat Plate 429

We now make the assumption that the velocity profiles are similar,
allowing us to express the velocity as

L. --,...+00
A.2: v.,(y, t) = v.,[..L]
!5(t)
Note that this equation is comparable to Eq. 11.2-14 if we designate r as

r =
y
!5(t)
(11.2-36)

the suddenly accelerated flat plate.


We further assume that the functional dependence of v., on yjt5(t) can be
represented by a polynomial.

is very reasonable on the basis of the A.3: (11.2-37)


indicate that all the momentum flux
to To determine the constants in the polynomial, we require sorne specific
information regarding the velocity profile, which is given as

T 11 .,,
1/=d(t)
- T 11 .,
11=0
Jdx (11.2-32) 1
v.,=O, y= !5(t) (an intuitive approximation) (11.2-38a)
ov.,_ o y= !5(t) (continuity of stress) (11.2-38b)
for T 11., yields oy- ,

v., = u0 , y=O (a bonafide boundary condition) (11.2-38c)


ov.,). (av.,). }d (11.2-33) These three equations allow us to determine only three of the constants in
- : ; 1/=d(t) - ay 11-o x
Eq. 11.2-37. We could obtain more equations by requiring the higher
deriva tives of v., to be zero at y = !5; however, a key piece of information
v., = O for y !5(t). If we require the
regarding the velocity profile can be obtained from the differential equation.
function, then approximation 1
Examining Eq. 11.2-1, keeping in mind that p and v., are independent of x,
we quickly conclude that

y = !5(t) (11.2-34) J2v., =O y=O (from the differential equation) (11.2-38d)


ol '
Equations 11.2-38 may be used to determine the constants a, b, e, and d to
obtain
-.t- av.,)
(ay (11.2-35) (11.2-39)
11=0

Defining dimensionless variables as


~""'''~"""u''"' of x; thus, the integration
result states that, the force exerted U - v., and Y= .l (11.2-40)
time rate of change of the momentum . , - Uo ~

we see that the velocity is given by


U., = 1 - ! Y+ ! Y3 (11.2-41)
S.C. 11.3 The Boundary Lay~r on a Flat Plate
-430 Flow Around lmmersed Bodies Chap. 11

Putting Eq. 11.2-35 into dimensionless form yields


Ua;

~(oJlu.,
dt
dY) =- ~(du., )
OdY Y=O
(11.2-42)
o --"'

By substituting Eq. 11.2-41 and carrying out the differentiation and integra-
tion, we obtain a first-order differential equation for the boundary layer _....
thickness.
--"'

do=(~) dt (11.2-43)
--"'

Separating variables, integrating and imposing the boundary condition --"'

B.C. 1: <5 =O, t=O (11.2-44)


we ha ve an expression for the boundary layer thickness as a function of time,
<5 = 2.8J;i (11.2-45)
While the boundary layer thickness predicted by the approximate solution is
very different from the exact solution, the result is notas bad as it might seem.
The comparison of velocity profiles is given in Fig. 11.2-2, and certainly the Fil. 11.3-1. Boundary layer
two profiles are in good agreement for values of y/#f less than 1.0. We
conclude that this type of approximate analysis should yield reasonably
will be given and compared with an
satisfactory results and we are now ready to attack the more complex
understood that the order of magnitude at
boundary layer flow illustrated in Fig. 11.3-1.
analysis and has applications far beyond
W e start with the two-dimensional
continuity equation for steady, incomprc
11.3 The Boundary Layer on a Flat Plate
v.,(ov.,) + v~~(ov.,) =-!
ax
1

The flow we wish to consider in this section is illustrated in Fig. 11.3-1. oy P


lt consists of an infinite stream flowing past a thin flat plate of length L. The
plate is infinitely wide and the flow is uniform at a velocity of U 00 This flow v.,(ovll) + v1 (ovll) = _!
has sorne of the same characteristics of the suddenly accelerated flat plate, ax oy P
especially if we view the problem as a flat plate moving at a velocity u 00
through an initially quiescent fluid . Under these conditions, the time t in the ov., +OVIl
previous problem would be comparable to xfu 00 where x is measured from ax oy
the leading edge of the plate. For this reason we may use Eq. 11.2-29 to We use the free stream velocity and
estmate the boundary layer thickness as dimensionless variables
<5 ~ 3.6-../vxfuoo (11.3-1)
Vz
Our first objective in this section will be to perform an order of magnitude
Uz=-,
Uoo
analysis of the equations of motion to obtain the Prandtl boundary layer
X
equations. We will not solve the equations because. the mathematical X=-,
techniques required are beyond the scope of this text ; however, the result L
Flow Around lmmersed Bodies Chap. 11 S.C. 11.3 The Boundary Lay~r on a Flat Plate 431

form yields

v(dU,.) (11.2-42)
=-~ df Y =O

out the differentiation and integra-


equation for the boundary !ayer

(11.2-43)

imposing the boundary condition


t=O (11.2-44)
layer thickness as a function of time,
2.8J;i (11.2-45)
by the approximate solution is
the result is notas bad as it might seem.
given in Fig. 11.2-2, and certainly the Fl. 11.31. Boundary layer fl.ow over a fl.at plate.
values of y/#f less than 1.0. We
analysis should yield reasonably
ready to attack the more complex will be given and compared with an approximate analysis. It should be
11.3-1. understood that the order of magnitude analysis is a general tool of engineering
analysis and has applications far beyond the development presented here.
We start with the two-dimensional Navier-Stokes equations and the
a Flat Plate continuity equation for steady, incompressible flow.

v.,(ov.,) + v (ov.,) = _ !(op) + v(o v., + 0 v.,)


2 2
(11.3-2)
section is illustrated in Fig. 11.3-1.
pasta thin flat plate of length L. The
ox oy 11
p ox ox ol 2

v.,(ov + v (ov = _ !(op) + v(o v + ov


2 2
uniform at a velocity of uro . This flow 11
)
11
)
11 11
) (11.3-3)
of the suddenly accelerated flat plate, ox oy 11
p oy ox ol 2

a flat plate moving at a velocity uro


these conditions, the time t in the ov., +OVIl= o (11.3-4)
to x/uro where x is measured from ax oy
reason we may use Eq. 11.2-29 to We use the free stream velocity and the length of the plate to form the
dimensionless variables
(11.3-1)
U= V:z P- Po
P=--
be to perform an order of magnitude :z
Uro
' pu!,
to obtain the Prandtl boundary layer
equations because. the mathematical X {)' = tJ(x)
X=-,
of this text; however, the result L L '
<432 Flow Around lmmersed Bodies Chap. 11 Sec. 11.3 The Boundary layer on a Flat Plat

and write the origir.al equations as

(1) (1) (<5') (~) (1') ( 1)


<5'2

~X (o U.,~
3 3
(11.3-5)
+ 3! axa)
(1) (<5') (<5') (1) ( <5')
(~) U.,lxo-AX = U.,lxo- ~x(~
u au" + u au" = _ aP + _1_ (a u"2 , a u")
2 2

"'ax "aY ay N Re,L ax ' aY 2


(11.3-6) _~X3! (ooXU.,)
3 3

3
1

(1) (1)
Subtracting Eq. 11.3-12 from Eq. 11.3
au., + au" = 0
(11.3-7)
deriva ti ve,
ax aY
where the quantities in parentheses over each term indicate the order of OU "') = UziXo+AX - Uzlxo-AX _
magnitude of that term. Here we ha ve used what is called a length Reyno/ds ( oX Xo 2~X
number, noting this fact by giving the Reynolds number as N Re L
An order of magnitud e analysis of a complex set of differenal equations
can be an extremely profitable undertaking. With this technique an engineer Provided the higher order terms are
may often reduce an essentially unsolvable problem to one which still approximation is
contains the pertinent elements of the real world yet is amenable to analysis. oU., 1 ~ U.,lxo+
The technique requires skill, daring, and luck, and the reward may be great. oX Xo
We start with the x-component of the dimensionless velocity, U.,. Its value
ranges between O at Y = O and 1 for Y <5', and we write the order of By adding Eqs. 11.3-11 and 11.3-12, w
magnitude of U., as for the second derivative.
u.,= 0(1) (11.3-8)
oU.,
2
U .,lxo+AX -
--~
The generally accepted meaning of this expression is that U., is closer to 1.0 oX 2

than it is to either 0.1 or 10.0.


An order of magnitude is perhaps best looked upon as an estima te of the From these two equations, we conc1ude
average value of a function over the region under consideration. Viewing it for the derivatives may be expressed as
in this way, we might express Eq. 11.3-8 as au.,
-=
6
ax
U.,= 0{~
l ff ~., dYdX] = 0(1)
c5 o o
(11.3-9) azu.,
oX 2 -
To treat the estimate of a derivative, we need to expand the function in a
Taylor series and develop sorne finite difference expressions. Expanding U., Letting X 0 = t and ~X = t. we find
around the point X = X 0 gives
au.,
u =UI +(X-X)au.,l +(X-Xo)2a2u.,l
2 ax
"'X "'Xo 0
ax Xo 2! aX Xo
oU
2
--"'-
+ (X - X o? aau., 1 ... (11.3-10) oX2
3! aX 3 Xo +
Flow Around lmmersed Boclies Chap. 11
Sec. 11.3 The Boundary Layer on a Flat Plate ..33

Letting (X- X0) equal +~X and -~X, respectively, gives


(1') ( 1)
15'2
U.,lx.+~X = U.,lx. + ~X (o 1 U.,)
~X (o U.,) 1
oX x. + T! (JX2 x.
2 2

(11.3-5)
(11.3-11)

(b') (~)
+ _1_ (o U oU
2 2
11 , 11)
(11.3-6)
Nlle,L oX2 T (Jy2
(11.3-12)

(11.3-7) Subtracting Eq. 11.3-12 from Eq. 11.3-11 yields an expression for the first
derivative,

oU.,) = U.,lxoHX- U.,lxo-AX- .!~r(oau.,) 1


over each term indicate the order of
used what is called a length Reynolds ( oX Xo 2~X 3! oX 3 Xo
Reynolds number as N Re L
a complex set of differenal equations + higher order terins (11.3-13)
With this technique an engineer Provided the higher order terms are not excessively large, a 'reasonable
problem to one which still approximation is
real world yet is amenable to analysis.
luck, and the reward may be great. oU., 1 ~ U:eiXoHX - U:elxo-AX (11.3-14)
dimensionless velocity, U.,. lts value oX Xo 2~X
Y b', and we write the order of By adding Eqs. 11.3-11 and 11.3-12, we may also obtain an approximation
for the second derivative.
(11.3-8) 2
expression is that U., is closer to 1.0 0 U.,
--~
U :elxo+AX - 2U :elxo + U,,lxo-AX ( 11.3-15)
oX 2 ~x 2
looked u pon as an estima te of the From these two equations, we conclude that the order of magnitude estimates
under consideration. Viewing it for the derivatives may be expressed as
as
oU., = 0(U.,)

dYdx) ~ 0(1)
(11.3-16a)
oX ~X
(11.3-9)
oU., _
2
O(U.,)
(11.3-16b)
need to expand the function in a oX 2 - ~X 2
tltlterence expressions. Expanding U., Letting X 0 = ! and ~X = !, we find that

+(X- Xo)2 iJ2U., 1 au., = 0(1)


1Xo 2! oX 2 Xo ax
+o oo (11.3-10) (J2U., = 0(1)
1X o oX 2
Flow Around lmmersed Bodles Chap. 11 S.C. 11.3 The Boundary Layer on a Flat

In the second case, we are stretching our definition, for ~X2 could also be On ~the basis of the order of J
written expect these equations to be valid o
~XI = ! = O(lo-1) express this restriction in terms of a
however, our final result would not be altered if we made this change.
To determine the order of magnitud~ of U11 , we note first that !5' = ~L ~3.6~(:
U X 00

aull = _ au., = 0(1) (11.3-17) or


oY ax NRe,a: ~
Using a finite difference formula in conjunction with Eq. 11.3-17, we find that because xjL is always 1ess than one
layer equations to be valid if the len
oU" = UviY-6' - UvJY=o = 0(1) (11.3-18) The solution of Eqs. 11.3-21 a1
ay <5'
possess at the moment; however, th
and we conclude that difficulty and the result is compare
u"= <5'[0(1)1 = 0(<5')
The other order of magnitude estimates indicated in Eqs. 11.3-5 1.0
1

and 11.3-6 readily follow by similar techniques. In examining Eqs. 11.3-5


and 11.3-6, we see that the corresponding terms in Eq. 11.3-6 are of order of
6' less than those in Eq. 11.3-5. Thus, if 6' is about I0-2 , we can neglect the
y-direction momentum equation and also (o 2 U.,joX2) relative to (o 2 U.,foY2),
which allows us to simplify the equations tot
0.8
~
,.. .)

au., au.,
11
dP 1 (J2 U.,)
u.,ax + u oY = - dX+N
Re,L
oY
au., aull = 0
2
Prandtl
boundary
!ayer
(11.3-19)

equations (11.3-20)
0.6

0.4 ,'
" ./,

ax +ay
These equations representan indeterminant set of equations, for we have three
unknowns-U.,, U11 , and P-and only two equations. However, for this type
0.2 1
11 "
of ftow we know that I'Eqs. 11, 3-21
dP =O
J ond 11.3-22 -
Y;;::: 6'
dX ' 1 1
and by our order of magnitude analysis we conclude that variations of P o 1.0 2.0 3.C
throughout the boundary layer must be small. This statement suggests that "'""Y.
dPjdX within the boundary layer must be nearly zero, and we are led at last
to the Prandtl boundary layer equations for ftow pasta ftat plate: FIJ. 11.3-l. Comparison of ex
profiles for ftow past a ftat plate.
au., au., 1
u.,ax +u~~ aY =-:-
(J2 U.,)
ayz (11.3-21)
ae,L The calculated and experimental v
(11.3-22) function of the single variable 1J = )
differential equations are transform
t This step relies very heavily on the plausible intuitive hypothesis that "small causa The length Reynolds number for the
produce small effects." 105 to 7.28 x 105, large enough tosa
Around lmmersed Bodles Chap. 11 S.C. 11.3 The Boundary Layer on a Flat Plate -435

definition, for ~X2 could also be On Hhe basis of the order of magnitude analysis presented here, we
expect these equations to be valid only if 15' :::;; 1Q-1 By Eq. 11.3-1, we may
express this restriction in terms of a length Reynolds number as
altered if we made this change.
of U11 , we note first that 15' = ~ ~ 3.6 :;-(~) = ~ (~) :::;; 10-2 (11.3-23)
L \) ;:::;: L JN Re,z L
(11.3-17) or
Nae,z :2:: (3.6)1 X 10' (11.3-24)
because x/Lis always less than one. Thus, we expect the Prandtl boundary
layer equations to be valid ifthe length Reynolds number is greater than 10'.
(11.3-18) The solution of Eqs. 11.3-21 and 11.3-22 requires more skill than we
possess at the moment; however, they can be solved without a great deal of
difliculty and the result is compared with experimental data in Fig. 11.3-2.

indicated in Eqs. 11.3-5 1.0


c~;J:um.uc::;.
,....a-
In examining Eqs. 11.3-5 ~
terms in Eq. 11.3-6 are of order of
15' is about IQ-2 , we can neglect the
(o 2 U.,/oX2) relative to (o 2 U,fo Y2),
tot
0.8
9
,.) V

JI/
Prandtl
boundary
(11.3-19)
u.
0.6
;;r
!ayer
NRe,.r =~
equations (11.3-20)

'
V
0.4 + 1.08xl0 5

set of equatins, for we have three


)1 < 1.82xl0 5

equations. However, for this type


0.2 1 o 3.64xl0 5
5.46xl05

ll
J
" t--Eqs. 11 . 3-21
ond 11.3-22

1 1
e 7.28xl05

we conclude that variations of P o 1.0 2.0 3.0 4.0 5.0 6.0 7.0
small. This statement suggests that '/}:ay u;;
.-v;-
be nearly zero, and we are led at last
for ftow pasta ftat plate: FIJ. 11.3-l. Comparison of experimental and theoretical velocity
profiles for ftow past a ftat plate.

- Nae,L
1 (ooYU)
- - - __.,
2

2
(11.3-21)
The calculated and experimental values of the velocity are plotted as a
(11.3-22) function of the single variable r = yV u 00 f'vx, which arises when the partial
differential equations are transformed to an ordinary differential equation.
intuitive hypothesis that "small causa The length Reynolds number for the experimental points ranges from 1.08 x
106 to 7.28 x 105, large enough to satisfy the restriction given by Eq. 11.3-24.
436 Flow Around lmmersed Bodies Chap. JI Sec. 11.3 The Boundary Layer on a Flat ~

We conclude that our order of magnitude analysis and simplification of the


t-,o-dimensional Navier-Stokes equations have been successful. The result
indicates that the boundary layer thickness is given by

5 = 4.9 r;;
'.};: (11.3-25)

which is not too different from the original estmate given by Eq. 11.3-1.
An interesting comparison between theory and experiment is the measured
and calculated drag coefficients shown in Fig. 11.3-3. It becomes obvious
that theory and experiment are not in agreement for values of NRe,z less
than lOS.

0.5
Fig. 11.3-<t. Control vol
1 1
~ For a fixed control volume an
0.2
o - Theory
takes the form
o
J
D Experiment

Co
0.1

0.05
" ~ o

~ ~.0
.........
A,
(pv)v n dA

Applying the x-component of this eq


Momentum ftux
atx + Ax
=

Momentu
at"
1

0.02

50 10,z 5xl0 2
NRe,L
103
' ~
2xl0 3
...............
5xl0 3

We ha ve made the following assumpti


Flg. 11.3-3. Comparison of experimental and theoretical drag coeffi-
cients for ftow pasta ftat plate. (Experimental data are taken from NACA analysis that led to the Prandtl bou
Tech. Mem. 1316, 1951.) A. 1 : The viscous stress on the
may be stated
Approximate solution
which is comparable to sp
Although we have left the Prandtl boundary layer equations unsolved,
we can analyze the ftow with an approximate solution. The control volume
to be used for the differential-macroscopic momentum balance is shown in a
Fig. 11.3-4. In this analysis, we will start directly with the macroscopic A. 2: The pressures at x and x
momentum balance; however, we could integrate the Prandtl boundary contribute to the moment
layer equations to obtain the same differential macroscopic momentum from the facts that (opfo
balance. This problem will be left as an exercise for the student. outside the boundary laye
Flow Around lmmersed Bodies Chap. 11 Sec. 11.3 The Boundary Layer on a Flat Plate 437

analysis and simplification of the


have been successful. The result
is given by

(11.3-25)

estimate given by Eq. 11.3-l.


and experiment is the measured
in Fig. 11.3-3. It becomes obvious
agreement for values of NRe,z less

r.---------- L --------~

Fig. ll.l...C. Control volume for approximate solution.

For a fixed control volume and steady ftow, the momenturn balance
Theory
takes the forrn
o Experiment
J
A,
(pv)v n dA = J
~
pg dV + J
~
t 1a 1 dA (11.3-26)

Applying the x-cornponent of this equation to the control volurne yields


Momentum flux Momentum ftux Momentum ftux
atx + Llx atx aty ~a

Surface force
aty =O

We have rnade the following assurnptions in accord with the order of magnitude
analysis that led to the Prandtl boundary !ayer equations.
A. 1 : The viscous stress on the x-surfaces is neglected. Thts assurnption
rnay be stated
r.,., <.{ r 11 .,
which is comparable to specifying that
boundary layer equations unsolved,
oxtmate solution. The control volume 02V 02V
-"'2 <.{ -"'
rnornenturn balance is shown in ox ol
directly with the rnacroscopic A. 2: The pressures at x and x +
tl.x are equal, and therefore do not
integrate the Prandtl boundary contribute to the rnornenturn balance. This assurnption is derived
differential rnacroscopic momenturn frorn the facts that (opfoy) is negligible and that the pressure
exercise for the student. outside the boundary !ayer is constant.
08 Flow Around lmmersed Bodies Chap. 11 Sec. 11.3 The Boundary Layer on a Flat 1

A. 3: The viscous stress at y = r5 is zero, which results from specifying Carrying out the integration and difl
that the velocity gradient (iJv,jiJy) must be continuous and is
therefore zero at the edge of the boundary layer. 39 (dr5~
280 dx/
Dividing Eq. 11.3--27 by ~x and taking the limit ~x-+ O, we obtain
Separating variables, integrating, an
6

t!_
dx
J pv: dy- pu~ (dr5)
dx
+ pU 00 V11 1
11~6
= -t-t iJv., 1
iJy ~~~o
(11.3-28) B. c. 1: r5 =o,
o we get an expression for the bounda
where the shear stress has been represented in terms of the velocity gradient
at y = O. Once again, we represent v., as a polynomial in {y/r5) =

(11.3-29) which is in excellent agreement witl


velocity profiles for the exact and a
which is subject to the restrictions 11.3-5, and the agreement is seen to
v.,= O, y=O (a bonafide boundary condition) (11.3-30a) agreement between the exact and a
accelerated flat plate.
V.,= U 00 , y=r5 (an intuitive approximation) (11.3-30b)
1.0 1
iJv., =O y=r5 (continuity of stress) {11.3-30c)
iJy '
iJ2v., = O y=O (from the differential equation) (113-30d) 0.8
iJl '
These conditions allow us to determine the four constants in Eq. 11.3-29, 0.6
and the velocity is given by A~"'
Ux
v., = u 00 {iY- !Y3) (11.3-31) r'
0.4 1..1
where
Y=~
r5 0.2
, ~
To determine v11 at y = r5, the continuity equation may be integrated
/r''
V
v~~j = -f6(iJv.,) dy (11.3-32a)
o 1.0 2.C
;
11~6 iJx
o
Fl1. 11.3-5. Velocity prc
and substitution of Eq. 11.3-31 for v., gives

3u"' (dr5) (11.3-32b) To gain sorne insight into the nat1


v11 1y~6 = -8- dx
the following example. For the flat F
Substitution of Eqs. 11.3-31 and 11.3-32b into Eq. 11.3-28, and the changing following conditions:
of the variable of integration from y to Y, give
u"'= 100 mph
p = 0.075 lbm/ft3 (e
(11.3-33) fl = 3.8 x I0-7 lbr
L = 6 in.
Around lmmersed Bodies Chap. 11 Sec. 11.3 The Boundary Layer on a Flat Plate -439

which results from specifying Carrying out the integration and differentiation yields
must be continuous and is
boundary layer. 39 (di}) 3( V ) (11.3-34)
280 dx = 215 u..,
the limit !l.x--+ O, we obtain
Separating variables, integrating, and applying the boundary condition

..,v, = -p, ov., 1 (11.3-28) B. c. 1: 15 =O, X= o (11.3-35)


v~6 oy !1=0
we get an expression for the boundary layer thickness
in terms of the velocity gradient
a polynomial in (y/15) IJ = 4.64 ;;
\};: (11.3-36)

(11.3-29) which is in excellent agreement with that given by the exact solution. The
velocity profiles for the exact and approximate solutions are shown in Fig.
11.3-5, and the agreement is seen to be excellent-far better, in fact, than the
(11.3-30a) agreement between the exact and approximate solutions for the suddenly
accelerated flat plate.
(11.3-30b)
1.0
(11.3-30c) ::.:::.0 ..:::
...J ~

0.8 . /~
(113-30d)
~~
the four constants in Eq. 11.3-29, 0.6
1'
~....
u. A
- tP) (11.3-31) ~('
0.4
~ Exoct solution

0.2
~"~"' ---Approximote solution
~-

equation may be integrated ./


/
(11.3-32a)
o 1.0 2.0 30 4.0 5.0
y VU(SJ/VX
FIJ. 11.3-5. Velocity profile for ftow past a ftat plate.

(:~) (11.3-32b) To gain sorne insight into the nature of the boundary !ayer, let us consider
the following example. For the flat plate shown in Fig. 11.3-1, we assume the
into Eq. 11.3-28, and the changing following conditions:
Y, give
u..,= 100 mph
p = 0.075 lbm/ft3 (density of air)
5(dt}) 3( V ) (11.3-33) p, = 3.8 X I0-7 lbr-sec/ft2 (viscosity of air)
S dx = - 215 """
L = 6 in.
Flow Around lmmersed Bodies Chap. 11 Sec. 11.-4 Externa! Flows and Wakes

The length Reynolds number is 6 x 105, and the boundary layer thickness region, a wake region, and a stagnatio
at x =Lis on the plate results almost entirely fron
back surfaces of the plate. Under th
!::::J(S) 6in.
proportional to the density times the
../6 X 105
:::::, 3.86 x 10- 2 in.
The point here is that the boundary layer thickness is in general quite small This situation is comparable to the je
for conditions that might be encountered for air flowing past solid objects. and the rough-pipe region of flow (
Since the kinematic viscosity v of water is about a factor of 10 less than that plate placed perpendicularly to the fn
of air, we can also expect boundary layers for the flow of water around solid jected area and the kinetic energy pe
objects to be very thin.
--d
en-(1
11 A Externa! Flows and Wakes Both mathematicaJ2- 5 and experimen
available for a flat plate placed pa
We are now in a position to understand the nature of the boundary layer These results are shown in Fig. 11.4
flow, and we must go on to the examination of the flow outside the boundary large difference in the drag coefficient
layer-i.e., the external flow-and the characteristics of the wake region. In For Reynolds numbers less than
the absence of a chapter on irrotational flow, we are forced to treat this drag coefficients differ by less than
problem in a qualitative manner. This will be done by giving a qualitative intuition that the drag coefficient fo
description of the flow about various solid bodies and examining the drag flow is the larger of the two. For Re
coefficients for these bodies. When a thin flat plate is placed parallel to the effects predominate for the perpend'
flow field, the drag exerted by the fluid is almost entirely from viscous effects, constant. The behavior of the parall
and we can use Eqs. 11.3-31 and coefficient continues to decrease with
11.3-36 to calculate the drag coeffi- keep in mind that the drag force is 1
cientt not increasing as fast as pu~. These n
L
distinct types of drag forces to be ce
2b J1-' (ov.,)
oy
dx there were two distinct mechanisms
e --..l!o_ _ _ __
v=O
rough pipes.
D- (2Lb)(}pu!,) While the drag coeffi.cient for the
1.293 2. Y. H. Kno, "On the Flow of an Inco
(11.4-1)
Moderate Reynolds Numbers," J. Math. Ph;
.JNRe,L 3. S. Tomotika and T. Aoi, "The Stea<
with considerable accuracy, provided Cylinder and a Flat Plate at Small Reynol
the boundary layer flow remains 1953, 6:290.
laminar and NRe L ~ 103 However, 4. B. A. Boley and M. B. Friedman, "Or
of a Flat Plate," J. Aerospace Sci. 1959, 26:L
if the plate is plaed perpendicular to
5. H. Schlichting, Boundary Layer Theor.
the flow, an entirely different situa- lnc., 1955), Chap. 7.
tion, depicted in Fig. 11.4-1, is 6. Z. Janour, "Resistance of a Plate in
Fl1. 11.4-1. Flow past a plate at high created. For this case, the flow may NASA Tech. Note 1316, Nov., 1951.
Reynolds numbers. be divided into an irrotational flow 7. S. Dhawan, "Direct Measurements of
8. S. F. Hoerner, Aerodynamic Drag (p
t The exact solution gives a coefficient of 1.328. New Jersey), Chap. 3.
Around lmmersed Bodies Chap. 11 Sec. 11.4 Externa! Flows and Wakes 441

region, a wake region, anda stagnation region in front ofthe plate. The drag
on the plate results almost entirely from the pressure difference on the front and
back surfaces of the plate. Under these conditions, the drag is very nearly
proportional to the density times the square of the free stream velocity.

drag oc pu~
thickness is in general quite small This situation is comparable to the jet impinging on the flat plate (Sec. 7.8)
for air flowing past solid objects. and the rough-pipe region of flow (Sec. 8.2). The drag coefficient for a flat
about a factor of 10 less than that plate placed perpendicularly to the free stream is defined in terms of the pro-
for the flow of water around solid jected area and the kinetic energy per unit volume of the free stream.

e - drag force
(11.4-2)
D - (Lb)(!pu~)
Both mathematical 2- 5 and experimental 6- 8 values of the drag coefficient are
available for a flat plate placed parallel and perpendicular to the flow.
the nature of the boundary !ayer These results are shown in Fig. 11.4-2, and, as we would expect, there is a
of the fiow outside the boundary large difference in the drag coefficients for the two configura~ions.
of the wake region. In For Reynolds numbers less than one, viscous effects predominate and the
fiow, we are forced to treat this drag coefficients differ by less than 50 per cent. lt is consistent with our
be done by giving a qualitative intuition that the drag coefficient for the plate placed perpendicular to the
bodies and examining the drag flow is the larger of the two. For Reynolds numbers greater than 10, inertial
fiat plate is placed parallel to the effects predominate for the perpendicular plate and the drag coefficient is
almost entirely from viscous effects, constant. The behavior of the parallel plate is quite different, and the drag
we can use Eqs. 11.3-31 and coefficient 'continues to decrease with increasing Reynolds number. We must
1.3-36 to calculate the drag coeffi- keep in mind that the drag force is not necessarily decreasing; it is simply
not increasing as fast as pu!,. These results indicate clearly that there are two
L
distinct types of drag forces to be considered for immersed bodies, just as
e __
Jp, (ov.,)
2b
ay
J!.o_ _ _~~~o
__
dx there were two distinct mechanisms of drag to be considered for flow in
rough pipes.
D- (2Lb)(!pu~) While the drag coefficient for the perpendicular plate remains constant
1.293 2. Y. H. Kno, "On the Flow of an Incompressible Viscous Fluid Past a Flat Plate at
(11.4-1)
.J NRe,L Moderate Reynolds Numbers," J. Math. Phys. 1953, 32:83.
3. S. Tomotika and T. Aoi, "The Steady Flow of a Viscous Fluid Past an Elliptic
considerable accuracy, provided Cylinder and a Flat Plate at Small Reynolds Numbers," Quart. J. Mech. Appl. Math.,
boundary !ayer flow remains 1953, 6:290.
and NRe L ~ 103 However, 4. B. A. Boley and M. B. Friedman, "On the Viscous Flow Around the Leading Edge
the plate is plaed perpendicular to of a Flat Plate," J. Aerospace Sci. 1959, 26:453.
5. H. Schlichting, Boundary Layer Theory (New York: McGraw-Hill Book Company,
fiow, an entirely different situa- Jnc., 1955), Chap. 7.
depicted in Fig. 11.4-1, is 6. Z. Janour, "Resistance of a Plate in Parallel Flow at Low Reynolds Numbers,"
For this case, the fiow may NASA Tech. Note 1316, Nov., 1951.
divided into an irrotational fiow 7. S. Dhawan, "Direct Measurements of Skin Friction," NACA Rept. 1121, 1953.
8. S. F. Hoerner, Aerodynamic Drag (pub. by author, 148 Busteed, Midland Park,
New Jersey), Chap. 3.
Flow Around lmmersed Bodies Chap. 11 Sec. 11.4 Externa! Flows and Wakes

1 1 1 11 1 1 1 1 1 erratic manner and then follows a s11


5.0 ' :-...
!"" :-...~
~ 1.1 Flot plote perpendicular to flow - the curve for the 1ower Reyno1ds nur

2.0
1' .- ~~-
in the ftow, and in view of our knov
turbu1ent ftow in smooth tubes, it st
1.0 " r~
to the onset of turbu1ence.
Both experiment and theory 6 ve
0.50 [".. 3 x 105 , and a schematic drawing o
/ Flot plot& porollel to flow
Co
0 .20 K
~ ........ Tron
1.-...J r-....
0.10

'
1-----Lominor____...,
......
0.05
't-.. ~"""-~
0.02
!"'
0 .0 1
0.1 0.2 0 .5 1.0 2.0 5.0 10 20 50 lo2 2 5 10\3 2 5 104
NRe
Fl1. 11.4-2. Drag coefficients for thin, fiat plates.

for Reyno1ds numbers higher than those in Fig. 11.4-2, an important change
in the drag characteristics of the paralle1 p1ate occurs. This change is illus-
trated in Fig. 11.4-3, where we see that the drag coefficient for a ftat p1ate
behaves much like the friction factor for ftow in a pipe. At a Reyno1ds
number of about 3 x 105, the drag coefficient suddently increases in an
Fl1. 11.4-4. Boundary ll
l{.r~
1 1
Q
~
1 1 11 As in the case of pipe ftow, the tra
1
T.
UUil~~IIVII
variety of factors. Schubauer and S
6 ~

t-~~~~ intensity of turbu1ence resu1ted in a


and by careful1y eliminating the free 1
~ ~~
-
Tur~ule1
t---. number of 3 x 108 was reached.

~~l'
-~ fl_

""" ~
On the basis of our qualitative e
eD

N _., ~
/ '1'.
~

-- - ~

'r-.
p1aced both paralle1 and perpendicu1
three distinct types of drag forces: "fo
acting on the front and back (or lel
"friction force"; and a turbu1ent "fr
mnor these phenomena can be obtained o
analysis and an extended survey of 1
description presented here does toucl
1(}3
in this chapter. .
The next step in the process of sh
NRa, L
9. G. B. Schubauer and H. K. Skramst1
Fl1. 11.4-l. Drag coefficient for a fiat plate-the transition to turbulence.
l Transition on a Flat Plate." NACA Rlpt. 9
Sec. 11.4 Externa! Flows and Wakes

erratic manner and then follows a smooth curve decreasing less rapidly than
the curve for the lower Reynolds numbers. Obviously a change has occurred
in the flow, and in view of our knowledge of the transition from laminar to
turbulent flow in smooth tubes, it seems reasonable to attribute the change
to the onset of turbulence.
Both experiment and theory 5 verify the onset of turbulence at N Re 40 ~
3 x 105, and a schematic drawing of the flow field is shown in Fig. 11 ..4-4.

Tronsition

for thin, ftat plates.

in Fig. 11.4-2, an important change


plate occurs. This change is illus-
the drag coefficient for a flat plate
flow in a pipe. At a Reynolds
oe1mc1.em suddently increases in an
Fl1. 11.4-4. Boundary layer formation on a ftat plate.

As in the case of pipe flow, the transition Reynolds number depends on a


variety of factors. Schubauer and Skramstad8 found that an increase in the
intensity of turbulence resulted in a transition at a lower Reynolds number,
and by carefully eliminating the free stream turbulence, a transition Reynolds
number of 3 x l ()8 was reached.
On the basis of our qualitative examination of the flow past a flat plate
placed both parallel and perpendicular toa uniform stream, we are aware of
three distinct types of drag forces: "form force" from the difference in pressure
acting on the front and back (or leading and trailing) surfaces; a laminar
"friction force"; anda turbulent "friction force." A clear understanding of
these phenomena can be obtained on1y with a more detailed mathematical
analysis and an extended survey of experimental results; however, the brief
description presented here does touch on the key points we wish to bring out
in this chapter. .
The next step in the process of sharpening our intuition is to examine the
9. G. B. Schubauer and H. K. Skramstad, "Laminar Boundary Layer Oscillations and
Transition on a Flat Plate." NACA hpt. 909, 1949.
Flow Around lmmersed Bodies Chap. 11 Sec. 11.4 Externa! Flows and Wakes

5.0

2 .0
--
Um l]o r- 1- Flat plate
10.0
8 .0
6 .0
4.0
.,... t-..
1.0 Cylinder
2 .0 ~
1' ""..
0 .50 r-.
e, ~QJD f-t Coi.O
0.20 1 0.8
' 11 0.6
0 .10
0.4
0 .05

0.2 1

0 .02
0 .01 0 .1 11111 1
0 .1 0 .2 0.5 1.0 2.0 5.0 10 20 50 102 2 10 2 4 6 8102 2 4 6 8103 2
NRe
Flg. 11.4-5. Drag coefficients for a cylinder anda ftat plate. Flg. 11.4-6. Drag coefficient vers
cylinder.
drag coefficients for a cylinder and a flat plate shown in Fig. 11.4-5. At low
Reynolds numbers, the drag coefficients are nearly the same, with the cylinder movies on drag, said, "Well, this is
showing a somewhat higher value for CD It is not at all obvious that the curious phenomenon we must exa
drag on these two rather different objects should be nearly identical in the low separation for both laminar and tur
Reynolds number region. However, it appears that the degree of deformation separation point when the boundary
suffered by the fluid as it flows past the plate and the cylinder is about the to the abrupt decrease in the drag e
same, which leads to similar values for the drag coefijcient. The drag on the layer separation, we will present a b
flat plate results entirely from form drag even at low Reynolds numbers, field outside both the wake and the t
while the total drag on the cylinder at low Reynolds numbers is evenly
divided between form and friction drag.3 From the results shown in Fig. The externa! flow
11.4-5, we conclude that the form drag for the cylinder must always be less
than that for the flat plate, and that streanilining reduces the total drag at Both analysis and experiment i
high Reynolds numbers. boundary layer can be described r
Recalling that the boundary !ayer on the flat plate become turbulent at a neglected entirely. For steady flow, t
high Reynolds number yielding an abrupt increase in the drag coefficient, we
might expect a similar change in the drag coefficient-Reynolds number curve pv Vv =
for the cylinder. The complete curve is shown in Fig. 11.4-6, and, indeed,
there is an abrupt change in the drag coefficient. But while the drag increases and the fl.ow is termed irrotational. Ec
for the flat plate, the drag on the cylinder undergoes an abrupt decrease! of what is called a perfect fluid, or an
Here it seems appropriate to quote Shapiro, 10 who, at this point in his classic fluid has a zero value of the viscos
11.4-3 may also be obtained by postul
10. A. H. Shapiro "The Fluid Mechanics of Drag," distributed by Educational Services,
Inc., 47 Galen St., Watertown 72, Mass. A written version of the experiments and commen- to any surface, requiring the stress te1
tary contained in the film is available in the paperback book, A. H. Shapiro, Shape and T=
Flow- The Fluid Dynamicsof Drag(Garden City, New York : Anchor Books, Doubleday
and Co. Inc., 1961). and the stress equations of motion qt
ow Around lmmersed Bodies Chap. 11 Sec. 11.4 Externa! Flows and Wakes 445

10.0
8.0
6.0
4.0

2.0 ~ .......

eno c.DI.O
~~
1--

0.8
1\
0.6
0.4

0.2

0 .1
10 2 4 6 8102 2 4 6 8103 2 4 6 8 lo4 2 4 6 8~ 2 4 6 81()6

NR
Fl. 11.4-6. Drag coefficient versus Reynolds number for a circular
cylinder.
plate shown in Fig. 11.4-5. At low
nearly the same, with the cylinder movies on drag, said, "Well, this is certainly unexpected." To explain this
D It is not at all obvious that the curious phenomenon we must examine the process of boundary !ayer
should be nearly identical in the Iow separation for both laminar and turbulent flow, for it is the shifting of the
that the degree of deformation separation point when the boundary layer becomes turbulent that gives rise
plate and the cylinder is about the to the abrupt decrease in the drag coefficient. Before considering boundary
drag coefl].cient. The drag on the !ayer separation, we will present a brief, qualitative description of the flow
even at low Reynolds numbers, field outside both the wake and the boundary !ayer.
low Reynolds numbers is evenly
3 From the results shown in Fig.
The externa! flow
for the cylinder must always be less
reduces the total drag at Both analysis and experiment indicate that the region outside the
boundary !ayer can be described reasonably well if viscous effects are
the flat plate become turbulent at a neglected entirely. For steady flow, the equations of motion reduce to
increase in the drag coefficient, we
coefficient-Reynolds number curve pv Vv = -Vp + pg (11.4-3)
shown in Fig. 11.4-6, and, indeed,
1 1,,1'-'ll" But while the drag increases
.. and the flow is termed irrotational. Equation 11.4-3 also describes the motion
undergoes an abrupt decrease! of what is called a perfect fluid, or an inviscid fluid. By definition, an inviscid
who, at this point in his classic fluid has a zero value of the viscosity coefficient; thus, fl = O. Equation
11.4-3 may also be obtained by postulating that the stress always acts normal
" distributed by Educational Services,
to any surface, requiring the stress tensor to be given by
of the experiments and commen-
book, A. H. Shapiro, Shape and T= -pi (11.4-4)
New York: Anchor Books, Doubleday
and the stress equations of motion quickly reduce to Eq. 11.4-3.
Flow Around lmmersed Bodies Chap. 11 S.C. 11.4 Externa! Flows and Wakes

Now, inviscid fluids do not exist in the real world. Yet there is much to be force on an immersed body, for expe
learned from the study of such fluids, just as we learned a great deal from the bodies experienced a drag at high
study of the suddenly accelerated, infinite, flat plate, even though it is a flow were presumably negligible.
which cannot be realized in practice. Here, we will examine the solution of The two points of maximum p1
Eq. 11.4-3 for flow pasta cylinder, only to obtain a limiting condition for referred to as the forward (O = O)
the form drag. The results of one possible solution ofEq. 11.4-3 for flow past terminology is appropriate, because
a cylinder are shown in Fig. 11.4-7. The streamlines are shown in Fig. the fluid is inviscid and "slips" past
11.4-7a and the pressure distribution over the surface of the cylinder in Fig. rear stagnation pressure approachc
11.4-7b. The analytic expression for the dimensionless pressure on the surface sometimes called the pressure recove1
of the cylinder11 is recovery is complete.
We now have sorne knowledge of
P - Pao = (1 - 4 sin 2 O) (11.4-5)
!pu!, i.e., it is zero for the flow illustrated in
description of inviscid flow somewha
Remembering that an inviscid fluid exerts only a normal force, we may use into the nature of the flow of real
Eq. 10.4-5 to express the drag coefficient as illustrated in Fig. 11.4-7 is just one o

f 11.4-3 subject to the boundary cond


21r

CD = Frorm =A n(1 - 4 sin 2 O) dO (11.4-6) velocity on the surface of the cylinde


(!pu!,XbD) Thete is a class of solutions for in
o
as Helmholtz flows,t or free streamli
Inasmuch as A n = -cos O, the drag coefficient is zero,
shown in Fig. 11.4-8 for flow past a
b

CD = - J(1- 4 sin O) cosO dO= O


2
(11.4-7)
o
This result, often known as D'Alembert's paradox,11 caused considerable
consternation among early researchers in fluid mechanics. Even in the
absence of viscous forces, it was felt that a streaming fluid would exert sorne
8='1TI2

(o)
Flg. 11.4-8. Helmho

-3
wake region was taken to be the sa
o '11'/2 1T
~) ~ o pressure in the wake is taken to be s
FIJ. 11.4-7. Flow of an inviscid ftuid past a cylinder. separation moves toward the rear <
progressively smaller. When the prc
11. L. M. Milne-Thomson, Theoretical Hydrodynamics, 4th ed. (New York: The forward stagnation pressure, the wa
Macmillan Company, 1960), p. 157. collapses to the continuous flow.
12. H. Rouse and S. Ince, History of Hydraulics (Ncw York: Dovcr Publications, Inc.,
1963), p. 102. t See Ref. 12, p. 201, and Ref. 11, Cha
Around lmmersed Bodies Chap. 11 S.C. 11.4 Externa! Flows and Wakes +47

force on an immersed body, for experience certainly indicated that immersed


as we learned a great deal from the bodies experienced a drag at high Reynolds numbers where viscous effects
flat plate, even though it is a flow were presumably negligible.
we will examine the solution of The two points of maximum pressure shown in Fig. 11.4-7 are often
to obtain a limiting condition for referred to as the forward (O = O) and rear (O = 1r) stagnation points. The
solution of Eq. 11.4-3 for flow past terminology is appropriate, because the velocity is zero at the points even if
streamlines are shown in Fig. the fluid is inviscid and "slips" past solid surfaces. The degree to which the
the surface of the cylinder in Fig. rear stagnation pressure approaches the forward stagnation pressure is
~rn~emolllleis pressure on the surface sometimes called the pressure recovery. For the illustrated case, the pressure
recovery is complete.
We now have sorne knowledge of the limiting behavior of the form drag,
4 sin 2 O) (11.4-5)
i.e., it is zero forthe flow illustrated in Fig. 11.4-7. We can carry this qualitative
description of inviscid flow somewhat further and perhaps gain more insight
only a normal force, we may use
into the nature of the flow of real fluids past immersed bodies. The flow
illustrated in Fig. 11.4-7 is just one of an infinite number of solutions to Eq.
11.4-3 subject to the boundary condition that the normal component of the
(11.4-6) velocity on the surface of the cylinder is zero.
There is a class of solutions for inviscid flow past immersed bodies known
as Helmholtz flows,t or free streamline flows. A possible flow of this type is
shown in Fig. 11.4-8 for flow pasta cylinder. The pressure in the stagnant

(11.4-7)
Helmholtz flow
paradox,11 caused considerable ,/'
in fluid mechanics. Even in the \ _..,Continuous 1
a streaming fluid would exert sorne --t. "( inviscid /
\ flow 1

,,_. ,
\ 1
\
~
3

o 7T/2 ,
8
(al (b)
Flg. 11.4-8. Helmholtz flow past a cylinder.

wake region was taken to be the same as the free stream pressure. If the
(b) O rr/2
pressure in the wake is taken to be sorne value larger than Poo the point of
0
fluid past a cylinder. separation moves toward the rear of the cylinder and the wake becomes
progressively smaller. When the pressure in the wake is taken to be the
forward stagnation pressure, the wake disappears and the Helmholtz flow
collapses to the continuous flow.
(New York: Dover Publications, Inc.,
t See Ref. 12, p. 201, and Ref. 11, Chap. 12.
448 Flow Around lmmersed Bodies Chap. 11 Sec. 1 1.-4 Externa! Flows and Wakes

Inasmuch as the pressure distribution in the Helmholtz flow is not


symmetric, a positive drag force results ; the flow is therefore in accord with
our experience. In addition, the stagnant wake region also bears sorne
resemblance to reality, although in actual fact it is unstable and a series of
oscillating vortices are formed.
Our treatment here has been extreme! y qualitative ; however, from
the Helmholtz and continuous flow pressure distributions we may draw the
following conclusion : When the .flow separa tes from the so/id surfa ce the
pressure recovery is not complete and the form drag increases.

Boundary !ayer separation


Seporotion point

When a real fluid flows pasta cylinder, a boundary layer forms, separating
from the cylinder as shown in Fig. 11.4-9. A detailed sketch of the boundary
Fig. 11.4-10. Separation of a lamn
!ayer separation is shown in Fig. 11.4-9, and a qualitative explanation of the

Laminar flow at intermediate which indicates that the sign () o


Boundary !ayer Reynolds numbers
whether the pressure gradient is positr
of th,e boundary !ayer on a flat pla
o "\ gradient does not vary across the bou
'\ \ 1
1 opfox at the solid surface by examinin
\ 1 layer.
\ 1 Over the front surface of the cylind
-3 / ' ,_.., /
Continuous inviscid flow
by Bernoulli's equation, then, the pr
inertia and the pressure gradient are
o 1rl2 1T against the effects of viscosity. A decr
{) a favorable pressure gradient, which
(al bl remains attached to the sol id surface. '
Fig. 11.4-9. Flow of a viscous fluid pasta cylinder.
op.
OX
separation phenomenon follows. We imagine a rectangular coordinate and Eq. ll.4-ll indicates that
system located on the surface of the cylinder as shown in Fig. 11.4-1 O, and
write the x-direction equation of motion for steady ftow and negligible a2v.,) < o
(ol
gravitational effects.

ov.,
p ( v., ox + Vu
ov., )
oy = -
op
OX
(ov ov., )
2

+ f-t OX + oy
2
.,
2

2
(11.4-8)
This condition is illustrated by profile
Let us now consider the possibilit:
did not take place, and that the ft
At the solid surface, v, and vu are both zero and Eq. 11.4-8 is simplified to accurately described by the continuot
For values of () > 1r/2, the pressure
(11.4-9) adverse pressure gradient wou\d be e:
forces would thereby be acting against
Around lmmersed Bodies Chap. JI Sec. 11.4 Externa! Flows and Wakes 449

-ffi-
in the Helmholtz flow is not

wake region also bears sorne (b)

fact it is unstable and a series of

qualitative; however, from Seporoted boundory loyer


distributions we may draw the
from the so/id surface the
form drag increases.

Woke

Seporotion point

r, a boundary layer forms, separating


. A detailed sketch of the boundary
Fig. 11.4-10. Separation of a laminar boundary !ayer from a cylinder.
and a qualitative explanation of the

Laminar flow at intermediate which indicates that the sign () of the second derivative .depends upon
RP.ynolds numbers
whether the pressure gradient is positive or negative. Following the analysis
/ "' of the boundary layer on a flat plate, we shall assume that the pressure
~ 1
o \ 1 gradient does not vary across the boundary layer; we can therefore determine
p-p,
__
a> -1 '\
\ opjox at the solid surface by examining the inviscid flow outside the boundary
t pu! \ !ayer.
\ 1 Over the front surface of the cylinder, the velocity increases as x increases;
-2
-3 / ' ,_.., /
Continuous inviscid flow
by Bernoulli's equation, then, the pressure must be decreasing. Both fluid
inertia and the pressure gradient are tending to propel the fluid forward
o rr/2 TT against the effects of viscosity. A decreasing pressure leads to what is called
() a favorable pressure gradient, which gives rise to a boundary layer that
(b) remains attached to the solid surface. When the pressure gradient is favorable,
fluid past a cylinder.
op < o
ox
imagine a rectangular coordinate and Eq. ll.4-11 indicates that
as shown in Fig. 11.4-1 O, and

(oolv.,)
2
for steady flow and negligible < 0
At the solid surface for a
(ll.4-10)
favorable pressure gradient

)v2"' )2v.,) This condition is illustrated by pro file (a) in Fig. 11.4-1 O.
OX
+.u-+-
( ox oy2 2
(11.4-8) Let us now consider the possibility that separation and wake formation
did not take place, and that the flow outside the boundary !ayer was
zero and Eq. 11.4-8 is simplified to accurately described by the continuous, inviscid flow shown in Fig. 11.4-7.
For values of () > TT/2, the pressure would be increasing and a so-called
(11.4-9) adverse pressure gradient would be established. Both viscous and pressure
forces would thereby be acting against the inertial force ofthe fluid, and there
<450 Flow Around lmmersed Boclles Chap. 11 Sec. 11.-4 Externa! Flows and Wakes

is a possibility of a backflow occurring and a wake forming. The point at


which the backflow starts is called the separation point. By Eq. 11.4-9 we
know that an adverse pressure gradient leads to
2

(oolvlll) >0 At the solid suace for an


(11.4-11)
adverse pressure gradient

Toe change in sign of the second derivative indicates an inflection point


such as that shown in profile (e) in Fig. 11.4-10. This qualitative analysis
indicates that the separation point should occur where the pressure is a
minimum. The continuous inviscid flow solution would therefore indicate a
separation point atO = 90, although the actual separation point for laminar
boundary layers13 is at O~ 80. For all the approximations, the agreement
is not bad.
We shall conclude this discussion of flow around immersed bodies by
examining all the flow regimes associated with uniform flow pasta cylinder.
In practice, the engineer will encounter a variety of shapes of immersed
bodies, and limiting our discussion to cylinders certainly does not do justice
to the subject. However, the pertinent aspects of the flow around immersed
Fl1. 11.4-11. Streamlines
bodies are contained in this one example, and it should be a satisfactory
introduction to the subject.
In Fig. 11.4-11, a series of flow fields is illustrated based on the obser- phenomenon. The Krmn vortex trru
vations and calculations of a number of investigators. 13- 17 At Reynolds range of from about 60 to 5000. At '
numbers less than one there is no wake and the flow follows the curved tends to become more turbulent in
surface of the cylinder in much the same manner as the inviscid continuous obscured.
flow. At a Reynolds number of about 4, a visible wake appears. As the As the Reynolds number increases
Reynolds number increases, the wake grows in size and the separation point where a sudden decrease in the drag e~
moves forward. At a Reynolds of about 40, the flow tends to become of this section is to try to provide a
unstable, and the two vortices behind the cylinder are alternately shed to phenomenon, and by now we have su
form what is known as Krmn's vortex trail. This flow pattern is illustrated From our examination of the bo
in Fig. 11.4-12. The alternate shedding ofvortices produces periodic transverse that a transition to turbulence takes :
forces on the cylinder that may cause severe damage if shedding frequency is
comparable to the resonant frequency of the cylinder. The "singing" of
electrical transmission wires in a high wind is a familiar example of this
13. A. S. Grove, F. H. Shair, E. E. Petersen, and A. Acrivos, "An Experimental
Investigation of the Steady Separated Flow Past a Circular Cylinder," J. Fluid Mech.,
1964, 19:60.
14. L. Prandtl and O. G. Tietjens, Applied Hydro and Aeromechanics, (New York:
Dover Publications, Inc., 1957), p. 302.
15. l. Proudman and J. R. A. Pearson, "Expansions at Small Reynolds Numbers for
Flow Pasta Sphere anda Circular Cylinder," J. Fluid Mech., 1957, 2:237.
16. A. Thom, "The Flow Past Circular Cylinders at Low Speeds," Proc. Roy. Soc.
(London) Ser. A., 1933, 141:651.
17. D. J. Tritton, "Experiments on the Flow Pasta Circular Cylinder at Low Reynolds
Numbers," J. Fluid Mech., 1959, 6:547. FIJ. 11.4-12. The
Around lmmersed Bodles Chap. 11 Sec. 11.-4 Externa! Flows and Wakes 451

and a wake forming. The point at


separation point. By Eq. 11.4-9 we
leads to

At the solid surface for an


(11.4-11)
adverse pressure gradient

indicates an inflection point


11.4-10. This qualitative analysis
occur where the pressure is a
solution would therefore indicate a
actual separation point for laminar
the approximations, the agreement

flow around immersed bodies by


with uniform flow pasta cylinder.
a variety of shapes of immersed
certainly does not do justice
of the flow around immersed
Fil. 11.4-11. Streamlines for ftow pasta cylinder.
and it should be a satisfactory

is illustrated based on the obser- phenomenon. The Krmn vortex trail is observed in the Reynolds number
of investigators. 1s-17 At Reynolds range of from about 60 to 5000. At higher Reynolds numbers, the motion
and the flow follows the curved tends to become more turbulent in nature and the distinct vortices are
manner as the inviscid continuous obscured.
4, a visible wake appears. As the As the Reynolds number increases further, it eventually reaches the point
in size and the separation point where a sudden decrease in the drag coefficient takes place. The main purpose
40, the flow tends to become of this section is to try to provide a rational explanation for this curious
the cylinder are alternately shed to phenomenon, and by now we have sufficient knowledge to do so.
trail. This flow pattern is illustrated From our examination of the boundary !ayer on a flat plate, we know
vntuc: produces periodic transverse that a transition to turbulence takes place for a length Reynolds number of
damage if shedding frequency is
of the cylinder. The "singing" of
wind is a familiar example of this
and A. Acrivos, "An Experimental
a Circular Cylinder," J. Fluid Mech.,

~xp;msons at Small Reynolds Numbers for


Fluid Mech., 1957, 2:237.
at Low Speeds," Proc. Roy. Soc.

r Past a Circular Cylinder at Low Reynolds


FIJ. 11.4-12. The Krmn vortex trail.
452 Flow Around lmmersed Bodies Chap. 11 Problems

about 3 x 105 lf we are not cautious and apply the result for a flat plate to
the cylinder, we may write PROB

11-1. Obtain an approximate solution fo


UooS) = 3 X 105 (11.4-12)
( v transition Sec. 11.2) using a fourth order poi)
c2v
where s is the distance from the forward stagnation point to the point where -"'=(
ay2
the transition toa turbulent boundary layer takes place. If 01 is the angle at
which the transition takes place, we may rewrite Eq. 11.4-14 as Compare the calculated boundary 1
exact solution.
(
uoo D)
V
(_!!__)
360
() t
= 3 X 105 (11.4-13) Ans: 6 = 3.65 y;
or 11-2. Integrate the Prandtl boundary laye
expressing v., as a polynomial in
Hint: Start by forming the integral
d
Thus, if the Reynolds number is greater than 3 x 10 5, we might expect the av.,
turbulent boundary to existo ver a portion of the forward half of the cylinder. Jo (v., ax +vil
The fact that the sudden reduction in the drag coefficient occurs ata Reynolds
number of about 3 x 105 certainly leads us to believe that this abrupt change and expressing v., as
is associated with the appearance of a turbulent boundary layer.
Careful experimental studies of this phenomenon18 - 20 have clearly
demonstrated that the appearance of the turbulent boundary layer causes
the abrupt decrease in the drag coefficient, by the following mechanism. 11-3. Obtain an approximate solution ~
When the boundary layer becomes turbulent, the separation point moves to velocity as
the rear of the cylinder, perhaps as much as 10. This change reduces the
area of the low pressure region, altering the pressure distribution around the
cylinder so that the pressure at the rear stagnation point increases. Both
these occurrences tend to reduce the form drag, and since it accounts for Compare the calculated boundary
practically all the drag at these high Reynolds numbers, the drag abruptly those given by the solution of the
decreases when the boundary layer becomes turbulent. Ans: 6 = 3.47 Vvxfuoo
The description of flow around immersed bodies presented in this chapter
has been brief and very qualitative. However, we have touched upon the 11-4. Air at 72F and 1 atm ftows past
essential elements-laminar and turbulent boundary layers, externa! flows, velocity of30 ft/sec. What is the bo
and wakes. Extensive experimental studies have been conducted for a large the leading edge?
variety of solid bodies, for both compressible and incompressible flow. None 11-5. Consideran imrnersed body ofma
of these, however, will be discussed here, for to do so would add little to our an infinite, quiescent fluid at a ve!
knowledge of flow phenomena if we could not pursue a more rigorous power required to move the bod
mathematical attack. characteristic area used in defining

18. See Ref. 14, p. 96. 11-6. The turbulent boundary layer t
19. A. Fage and J. H. Warsap, "The Effects of Turbulence and SurfaceRoughness on determined approximately if we
the Drag of a Circular Cylinder," Aero. Res. Council Rept. 1283, Gr. Brit., 1930. ftow in smooth tubes. The velocit
20. A. Roshko, "Experiments on the Flow Past a Circular Cylinder at Very High
Reynolds Numbers," J. Fluid Mech., 1961, 10:345.
Flow Around lmmersed Bodies Chap. 11 Problems 453

apply the result for a fiat plate to


PROBLEMS

11-1. Obtain an approximate solution for the suddenly accelerated flat plate (see
(11.4-12)
Sec. 11.2) using a fourth order polynomial and the additional condition
o2v
stagnation point to the point where oy2, = o, y = o(r)
!ayer takes place. If 01 is the angle at
rewrite Eq. 11.4-14 as Compare the calculated boundary !ayer thickness with that obtained by the
exact solution.
(11.4-13) Ans: o = 3.65 Vrl
11-2. Integrate the Prandtl boundary !ayer equations between y = O and y = o(x),
expressing v, as a polynomial in yfo(x), to derive an expression for o(x).
Hint: Start by forming the integral
d d
than 3 x 105, we might expect the
of the forward half of the cylinder.
drag coefficient occurs ata Reynolds
f ( :::z +
o
V:z Vy :;:z) dy = V f :;:z
o
dy

us to believe that this abrupt change and expressing v, as


boundary !ayer.
phenomenon18 - 20 have clearly
turbulent boundary !ayer causes
by the following mechanism. 11-3. Obtain an approximate solution for flow past a flat plate, expressing the
the separation point moves to velocity as
as 10. This change reduces the
the pressure distribution around the o s.; y s.; o
stagnation point increases. Both
drag, and since it accounts for Compare the calculated boundary !ayer thickness and drag coefficient with
...... v.u-> numbers, the drag abruptly those given by the solution of the Prandtl boundary layer equations.
turbulent.
Ans: o = 3.47 Vvxfu 00
bodies presented in this chapter
nuJPv"r, we have touched upon the 11-4. Air at nF and 1 atm flows past a flat plate with a uniform free stream
boundary layers, externa! fiows, velocity of30 ft/sec. What is the boundary !ayer thickness ata point 5 in. from
have been conducted for a large the leading edge?
and incompressible fiow. None 11-5. Consideran immersed body of mass M (such as a submarine) moving through
for to do so would add little to our an infinite, quiescent fluid at a velocity w = :Au0 Derive an expression for the
could not pursue a more rigorous power required to move the body in terms of eD and u0 Let A be the
characteristic area used in defining eD
11-6. The turbulent boundary !ayer thickness on a smooth, flat plate can be
of Turbulence and Surface Roughness on determined approximately if we make use of experimental data for turbulent
Rept. 1283, Gr. Brit., 1930. flow in smooth tubes. The velocity profile may be approximated by
Past a Circular Cylinder at Very High
(see Sec. 6.3)
454 Flow Around lmmersed Bodies Chap. 11

and the shear stress at the wall by

To = (o p (z )2

(see Sec. 8.2)

thus leading to
lndex
To = ({)pu!o
1= 0.316c<5 VUoo rl/4
Here we have replaced the average velocity in a tube with the free stream
velocity, and the tu be diameter with 2<5. We can use the momentum equation
to determine the turbulent boundary !ayer thickness. Consider carefully the
boundary condition to be imposed on the derived differential equation for <5,
Use these results to calculate a drag coefficient that may be compared with
the experimental curve for CD shown in Fig. 11.4-3.

Acceleration, 82, 166


Adiabatic flow, 398
Ambient pressure in the momentum
balance, 257
Angular momentum principie, 7
Apparent viscosity, 19, 68
Archimedes' principie, 53
Atmospheric pressure, 43
Averages
area and volume, 5, 61, 109
time, 187
Fl. 11-7. Suddenly accelerated joumal bearing.
Barometers, 43
Bemoulli's equation, 230
11-7. A bearing, shown in Fig. 11-7, is initially at rest. At time t = O it is suddenly
Bingham model fluid, 19
accelerated to a constant angular velocity, w. If the viscosity of the oil in the Blasius, H., 300
bearing is 100 centipoise and the annular gap is 0.10 in., estimare the time Borda mouthpiece, 247
required to obtain a steady velocity profile. Boundary conditions, 42, 173
Boundary layer flow, 422, 430
Boundary layer thickness, 426, 431
Boundary layer separation, 448
Buoyancy force, 53

Cavitation, 22
Cavitation number, 22
Center of pressure, 49
Center of stress, Prob. 2-14
Olokc flow, 412
Flow Around lmmersed Bodies Chap. 11

(see Sec. 8.2)


0.316Nil'' Index

velocity in a tube with the free stream


215. We can use themomentumequation
!ayer thickness. Consider carefully the
the derived differential equation for d.
coefficient that may be compared with
in Fig. ll.4-3.

Acceleration, 82, 166 Coefficient of expansion, 13


Adiabatic ftow, 398 Coefficient of viscosity, 14
Ambient pressure in the momentum Colebrook, C. F., 300
balance, 257 Compressibility, 13
Angular momentum principie, 7 Compressible ftow, 3, 390
Apparent viscosity, 19, 68 Conservation of energy, 8, 391
Archimedes' principie, 53 Conservation of mass, 8, 92
Atmospheric pressure, 43 Conservation of momentum, 8
Averages Constitutive equations, 9, 128
area and volume, 5, 61, 109 Continuity equation, 93, 189
time, 187 Continuity of stress, 42, 174
Continuity of velocity, 42, 176
Barometers, 43 Continuum, 1
Bernoulli's equation, 230 Contraction coefficient, 334
at rest. At time t = O it is suddenly Control surface, 33
Bingham model fluid, 19
w. lf the viscosity of the oil in the Control volume, 33, 88
Blasius, H., 300
gap is 0.10 in., estimate the time Borda mouthpiece, 247 Convective acceleration, 82
pro file. Boundary conditions, 42, 173 Couette viscometer, 15, 177
Boundary !ayer ftow, 422, 430 Critica! ftow, 359
Boundary !ayer thickness, 426, 431 Critica! depth, 362
Boundary !ayer separation, 448 Critica! Reynolds Number, 5, 195
Buoyancy force, 53 Critica! velocity, 362
Cylinder, ftow around, 305, 445
Cavitation, 22
Cavitation number, 22 D'Alembert's paradox, 446
Center of pressure, 49 Darcy-Weishbach formula, 310
Center of stress, Prob. 2-14 Deformation, 135
Cbokc flow, 412 Density, 2
455
456 lndex lndex

Derivatives (see Material derivative, Total Ideal gas Jaw, 9 Partial derivative, 79
derivative, and Partial derivative) lncompressible flow, 3 Path lines, 97
Differential-macroscopic balance, 273, lndex notation, 26 Perfect gas, 9, 402
355, 405, 436 lrrotational flow, 151, 445 Piezometric head, 308
Diffuser, 312 lsentropic flow, 398, 404 Pipe flow:
Dilatant fluid, 20, 68 Isotropy, 141 laminar, 57, 169, 294
Dimensional analysis, 158, 291 turbulent, 196, 295
Discharge coefficient, 333 Kinematic viscosity, 16 Pipeline networks, 318
Divergence of a tensor, 123 Kinetic energy, 224 Pitot-static tube, 337
Divergence of a vector, 84 Kronecker delta, 130 Potential energy function, 41, 393
Divergence theorem, 84 Power-law fluid, 20, 63
Drag coefficients, 304 Laminar boundary !ayer, 430 Prandtl boundary !ayer equations, 434
Drag force, 286 Laminar flow, 4, 56, 169, 294 Prandtl mixing length theory, 200
Dummy index, 27 Laminar sublayer, 197 Pressure:
Dyad, 112 Laws of Mechanics, 7 atmospheric, 43
Linear momentum principie, 8, 36 gauge,45
Eddy viscosity, 200 Local acceleration, 82 vapor, 22
Energy equation, 391 Loss coefficient, 311, 316 Pressure head, 308
Enthalpy, 394 Projected areas, 37, 53, 111
Entrance length, 171, 369 Mach number, 403 Pseudoplastic fluid, 20
Entrance Joss, 315 Manning formula, 352
Entropy equation, 398 Manometer, 44, 323 Rate of strain, 14, 135, 147
Equation of state, 9 Mass balance, 214 Rate of strain tensor, 134, 138
Equations of fluid statics, 36 Material coordinates, 76 Reducer, 313
Equations of motion, 154, 193 Material derivative, 78 Relative roughness, 293
Material position vector, 76 Relative velocity, 156
First law of thermodynamics, 391 Material surface, 33 Reynolds, 0., 4
Flow meters, 331 Material volume, 32 Reynolds number, 5, 160
Flow nozzle, 335 Mechanical energy balance, 228 Reynolds stresses, 194
Free index convention, 27 Mechanical energy equation, 224, 396 Reynolds transport theorem, 92
Friction factors, 287, 293, 352 Minor loss, 308 Rotation, 151
Friction Joss, 308 Mixing Jength, 202 Roughness factor, 293
Froude number, 165, 359 Momentum balance, 218, 254
Moody chart, 298 Schedule number, 302
Gas constant, 402 Moving control volumes, 261 Separation, 448
Gauge pressure, 45 Moving reference frames, 154 Sharp-crested weir, 339
Gauss' theorem, 84 Shear rate, 14
Gibbs notation, 23 Shock waves, 416
Navier-Stokes equations, 154
Gradient of a scalar, 40 Net stress vector, 257 Slip flow, 3
Gradually varied flow, 355 Solitary wave, 370
Newton's law of viscosity, 16, 133, 139
Gravitational constant, 11 Sonic velocity, 3, 398
Newton's second law of motion, 7
Gravitational potential function, 40 Spatial coordinates, 76
Noncircular r.onduits, 303
Non-Newtonian fluids, 16, 63 Spatial position vector, 76
Normal vector, 35, 46 Specific energy, 360
Hagen-Poiseuille flow, 61 Specific gravity, 55
Hardy Cross method, 319 Nozzle flow, 247, 335
Speed of sound, 402
Head loss, 308 Stagnation pressure, 407
Helmholtz flows, 447 One-dimensional flow, 6 Stagnation temperature, 407
Hydraulic jump, 382 Order of magnitude analysis, 329 Steady flow, 6
Hydraulic radius, 159, 303, 310, 352 Orifice meter, 334 Strain rate, 14
Hydrometer, 55 Oscillating systems, 323 Streak lines, 97
Hydrostatic stress, 36 Ostwald-de Wael model fluid, 20, 63 Stream function, 103
lndex lndex 457
Ideal gas law, 9 Partial derivative, 79 Streamlines, 97
Incompressible flow, 3 Path lines, 97 Stress equations of motion, 121, 131 , 219
Index notation, 26 Perfect gas, 9, 402 Stress tensor, 112
Irrotational flow, 151, 445 Piezometric head, 308 Stress vector, 35, 108
n ...,ntrt)nir flow, 398, 404 Pipe flow: Supercritical flow, 359
lsotropy, 141 laminar, 57, 169, 294 Surface profiles, 364
turbulent, 196, 295 Surface tension, 21
Kinematic viscosity, 16 Pipeline networks, 318
Kinetic energy, 224 Pitot-static tube, 337 Tangent vector, 49, 98, 230
Kronecker delta, 130 Potential energy function, 41, 393 Thermal energy equation, 396
Power-law fluid, 20, 63 Time a verages, 187
Laminar boundary !ayer, 430 Prandtl boundary !ayer equations, 434 Time derivatives (see Material derivative,
Laminar flow, 4, 56, 169, 294 Prandtl mixing length theory, 200 Total deriva ti ve, and Partial derivative)
1'-"''uu. mu sublayer, 197 Pressure: Torque, 7, 48, 179
of Mechanics, 7 atmospheric, 43 Total derivative, 77
momentum principie, 8, 36 gauge, 45 Transition region, 197
acceleration, 82 vapor, 22 Transport theorem, 88
coefficient, 311 , 316 Pressure head, 308 Turbulent core, 197
Projected areas, 37, 53, 111 Turbulent flow, 4, 186
Mach number, 403 Pseudoplastic fluid, 20 Turbulent stress tensor, 194
Manning formula , 352 Turbulent velocity profile, 196, 203, 242
Manometer, 44, 323 Rate of strain, 14, 135, 147 Two phase flow, 173
Mass balance, 214 Rate of strain tensor, 134, 138
Material coordinates, 76 Reducer, 313 Uniform flow, 349
Material derivative, 78 Relative roughness, 293 Uniformly accelerated flow, 166
Material position vector, 76 Relative velocity, 156 Units, 10
Material surface, 33 Reynolds, O., 4 Unit tensor, 129
Material volume, 32 Reynolds number, 5, 160 Universal gas constant, 402
Mechanical energy balance, 228 Reynolds stresses, 194 Unsteady flow, 6, 261, 320
Mechanical energy equation, 224, 396 Reynolds transport theorem, 92 U-tube manometer, 44, 323
Minor loss, 308 Rotation, 151
Mixing length, 202 Roughness factor, 293 Vapor pressure, 22
Momentum balance, 218, 254 Vectors, 22
Moody chart, 298 Schedule number, 302 Velocity head, 308
control volumes, 261 Separation, 448 Velocity of sound, 402
reference frames, 154 Sharp-crested weir, 339 Vena contracta, 334
Shear rate, 14 Venturi meter, 331
Shock waves, 416 Viscosity, 14, 134
Slip flow, 3 Viscous dissipation, 223, 396
Solitary wave, 370 Viscous stress tensor, 130
Sonic velocity, 3, 398 Vorticity tensor, 138, 152
Spatial coordinates, 76 Vorticity vector, 152
Spatial position vector, 76
Specific energy, 360 Wall shear stress, 205, 289, 351
Specific gravity, 55 Waves:
Speed of sound, 402 gravity, 374
Stagnation pressure, 407 ripples, 374
One-dimensional flow, 6 Stagnation temperature, 407 shallow-water, 370

~
der of magnitude analysis, 329
Steady flow, 6 Wave speed, 373
rifice meter, 334 Strain rate, 14 Weir:
cillating systems, 323 Streak lines, 97 broad-crested, 379
twald-de Wael model fluid, 20, 63 Stream function, 103 sharP-crested, 339

S-ar putea să vă placă și