0% found this document useful (0 votes)
31 views

FLUID MECHANICS Electric Power Final

This document outlines the course objectives, notes, grading system, and contents for a Fluid Mechanics course at Cairo University's Faculty of Engineering. The course aims to develop students' confidence in applying fluid mechanics principles and concepts to a variety of problems. It also aims to promote interest in fluid mechanics and prepare students for more advanced study. The course contents cover topics such as fluid properties, basic principles, fluid statics, the integral forms of fundamental laws, and references. Student performance will be evaluated based on mid-term and final exams, as well as class exams and activities throughout the semester.

Uploaded by

Fady Micheal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
31 views

FLUID MECHANICS Electric Power Final

This document outlines the course objectives, notes, grading system, and contents for a Fluid Mechanics course at Cairo University's Faculty of Engineering. The course aims to develop students' confidence in applying fluid mechanics principles and concepts to a variety of problems. It also aims to promote interest in fluid mechanics and prepare students for more advanced study. The course contents cover topics such as fluid properties, basic principles, fluid statics, the integral forms of fundamental laws, and references. Student performance will be evaluated based on mid-term and final exams, as well as class exams and activities throughout the semester.

Uploaded by

Fady Micheal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 89

CAIRO UNIVERSITY

FACULTY OF ENGINEERING

Fluid Mechanics
SECOND YEAR
Electric Power Engineering

2018

1
PREFACE

Course Objectives

The objective of this course is to develop confidence in the use of basic principles,
concepts and equations by applying them in a wide variety of situations.

To develop familiarity with the main areas of knowledge of fluid mechanics


phenomena, in order to promote interest and serve as a stepping stone to more
advanced study.

An "in depth" understanding is not expected but students should acquire an awareness
of what sorts of problems can be solved, what techniques are available and where
they can go to get more information.

Course Notes

This course notes are taken and compiled directly from the references cited in the
reference section. It aims to provide a clear and concise presentation of the
fundamentals of fluid mechanics to the student. The student should read these notes
before coming to the class. The classroom time should be devoted to amplify the notes
material by discussing related issues and applying basic principles to the solution of
problems.

Course Grading System

Activities Expected Date Mark

Mid-term exam. Early April 25%

Class exams and class activities. Bi-weekly 5%

Final exam. Late May or 70%


early June

2
Contents
Chapter 1: INTRODUCTION ................................................................................................... 5
1.0 Definition of a Fluid and Fluid Mechanics................................................................ 5
1.1 Dimensions and Units ............................................................................................... 5
1.1.1. Introduction ....................................................................................................... 5
1.1.2. Systems of Dimensions ..................................................................................... 5
1.1.3. Systems of Units................................................................................................ 6
1.1.4. Preferred Systems of Units ................................................................................ 7
Chapter 2: BASIC PRINCIPLES AND CONCEPTS ............................................................. 10
2.1 Fluid as a Continuum ............................................................................................... 10
2.1.1 Density ............................................................................................................ 10
2.1.2 Specific Volume .............................................................................................. 10
2.1.3 Pressure ........................................................................................................... 11
2.1.4 Specific Weight ............................................................................................... 12
2.1.5 Specific Gravity............................................................................................... 12
2.2 Velocity Field ........................................................................................................... 12
2.3 Stress Field............................................................................................................... 13
2.4 Dynamic Viscosity of a Fluid .................................................................................... 16
2.4.1 Mathematical Modeling of Viscosity .............................................................. 16
2.4.2 Newtonian Fluid .............................................................................................. 16
2.4.3 Non-Newtonian Fluid ...................................................................................... 16
2.4.4 Kinematic Viscosity ........................................................................................ 17
2.5 Fluid Compressibility ............................................................................................... 20
2.6 Reynolds Number .................................................................................................... 21
2.7 Speed of Sound........................................................................................................ 21
2.8 Mach Number .......................................................................................................... 21
2.9 Description and classification of Fluid Motions ..................................................... 23
Chapter 3: FLUID STATICS .................................................................................................. 24
3.0 Introduction ............................................................................................................. 24
3.1 The Basic Equation of Fluid Statics ........................................................................ 24
3.2 Pressure Variation in a Static Fluid ......................................................................... 28
3.3 Forces on Plane Areas ............................................................................................. 35
3.4 Forces on Curved Surfaces ...................................................................................... 40
3.4.1 Hydrostatic Force on a Curved Submerged Surface ....................................... 41
3.4.2 Buoyancy Force............................................................................................... 43
3.4.3 Archimedes Law.............................................................................................. 43
Chapter 4: THE INTEGRAL FORMS OF THE FUNDAMENTAL LAWS ........................ 46

3
4.1 INTRODUCTION ................................................................................................... 46
4.1.1 Conservation of Mass ...................................................................................... 47
4.1.2 Newton's Second Law ..................................................................................... 48
4.1.3 First Law of Thermodynamics ........................................................................ 48
4.1.4 Second Law of Thermodynamics .................................................................... 48
4.2 SYSTEM-TO-CONTROL-VOLUME TRANSFORMATION: ............................. 49
4.2.1 Simplifications of the System-to-Control-Volume Transformation ................ 49
4.3 CONSERVATION OF MASS ................................................................................ 50
4.4 CONSERVATION OF MOMENTUM (MOMENTUM EQUATION) ................. 55
4.4.1 General Equation ............................................................................................. 55
4.4.2 Notes on Momentum Equation........................................................................ 55
4.4.3 Control Volume Moving with Constant Velocity ........................................... 61
4.4.4 Momentum Equation For Control Volume Moving with a Rectilinear
Acceleration .................................................................................................................... 63
4.5 ENERGY EQUATION ........................................................................................... 70
4.5.1 Work-Rate Term.............................................................................................. 70
4.5.2 General Energy Equation ................................................................................ 71
4.5.3 Steady, Uniform Flow ..................................................................................... 71
4.5.4 The Bernoulli Equation .................................................................................. 72
5. REFERENCES ................................................................................................................ 89

4
Chapter 1: INTRODUCTION
1.0 Definition of a Fluid and Fluid Mechanics
■ What is Fluid Mechanics?
Fluid mechanics is that branch of applied mechanics which is concerned with the
statics and dynamics of liquids and gases.

■ What is a Fluid?
- A fluid is a substance that deforms continuously under the action of
shearing stresses no matter how small the shear stress is.
- If a fluid is at rest there can be no shearing forces acting and therefore all
forces in the fluid must be perpendicular to the planes upon which they act.
- Fluids are either liquids, e.g. water, or gases, e.g. air.

■ What is the Objective of Studying Fluid Mechanics?


The objective is to determine the interactions between fluids and objects.
These interactions may take the forms of force, torque, and power.

1.1 Dimensions and Units

1.1.1. Introduction
Engineering problems are solved to answer specific questions. It goes
without saying that the answer must include units. (It makes a difference
whether a pipe diameter required is 1 meter or 1 foot!) Consequently, it is
appropriate to present a brief review of dimensions and units. We say
"review" because the topic is familiar from your earlier work in
mechanics.
We refer to physical quantities such as length, time, mass, and temperature
as dimensions. In terms of a particular system of dimensions, all
measurable quantities can be subdivided into two groups—primary
quantities and secondary quantities. We refer to a small group of
dimensions from which all others can be formed as primary quantities.
Primary quantities are those for which we set up arbitrary scales of measure;
secondary quantities are those quantities whose dimensions are expressible
in terms of the dimensions of the primary quantities.
Units are the arbitrary names (and magnitudes) assigned to the primary
dimensions adopted as standards for measurement. For example, the
primary dimension of length may be measured in units of meters, feet,
yards, or miles. These units of length are related to each other through
unit conversion factors (1 mile = 5280 feet = 1609 meters).

1.1.2. Systems of Dimensions


Any valid equation that relates physical quantities must be dimensionally
homogeneous; each term in the equation must have the same dimensions.
We recognize that Newton's second law (F * ma) relates the four
dimensions, F, M. L, and t. Thus force and mass cannot both be selected
as primary dimensions without introducing a constant of proportionality
that has dimensions (and units).
Length and time are primary dimensions in all dimensional systems in
common use. In some systems, mass is taken as a primary dimension. In

5
others, force is selected as a primary dimension; a third system chooses
both force and mass as primary dimensions. Thus we have three basic
systems of dimensions, corresponding to the different ways of specifying
the primary dimensions.
a. Mass [M], length [L], time [t], temperature [T].
b. Force [F], length [L], time [t], temperature [T].
c. Force [F], Mass [M], length [L], time [t], temperature [T].

In system a. force [F] is a secondary dimension and the constant of


proportionality in Newton's second law is dimensionless. In system b,
mass [M] is a secondary dimension, and again the constant of
proportionality in Newton's second law is dimensionless. In system c,
both force [F] and mass [M] have been selected as primary dimensions.
In this case the constant of proportionality, 𝑔𝑐 in Newton's second law
(written 𝐹̅ = 𝑚𝑎 ̅/𝑔𝑐 ) is not dimensionless. The dimensions of 𝑔𝑐 must
in fact be [ML/Ft2] for the equation to be dimensionally homogeneous. The
numerical value of the constant of proportionality depends on the units of
measure chosen for each of the primary quantities.

1.1.3. Systems of Units


There is more than one way to select the unit of measure for each primary
dimension. We shall present only the more common engineering systems
of units for each of the basic systems of dimensions.
a) MLtT
SI, which is the official abbreviation in all languages for the Système
International d’Unités, is an extension and refinement of the traditional
metric system. More than 30 countries have declared it to be the only
legally accepted system.
In the SI system of units, the unit of mass is the kilogram (kg), the unit of
length is the meter (m), the unit of time is the second (sec), and the unit
of temperature is the Kelvin (K). Force is a secondary dimension, and its
unit, the Newton (N), is defined from Newton's second law as
1 N ≡ kg ∙ m/sec 2
In the Absolute Metric system of units, the unit of mass is the gram, the unit
of length is the centimeter, the unit of time is the second, and the unit of
temperature is the Kelvin. Since force is a secondary dimension, the unit of
force, the dyne, is defined in terms of Newton's second law as
1 dyne ≡ 1 g ∙ cm/sec2

b) FLtT
In the British Gravitational system of units, the unit of force is the pound
(lbf), the unit of length is the foot (ft), the unit of time is the second, and
the unit of temperature is the Rankine (R). Since mass is a secondary
dimension, the unit of mass, the slug, is defined in terms of Newton's
second law as

1 slug ≡ 1 lbf ∙ sec2/ft

c) FMLtT

6
In the English Engineering system of units, the unit of force is the pound
force (lbf), the unit of mass is the pound mass (lbm), the unit of length is
the foot, the unit of time is the second, and the unit of temperature is the
Rankine. Since both force and mass are chosen as primary dimensions.
Newton's second law is written as
𝑚𝑎̅
𝐹̅ =
𝑔𝑐

A force of one pound (1 lbf) is the force that gives a pound mass (1 lbm) an
acceleration equal to the standard acceleration of gravity on Earth, 32.17
ft/sec2. From Newton's second law we see that (to three significant figures)
1 𝐼𝑏𝑚 𝑥 32.2 𝑓𝑡/𝑠𝑒𝑐 2
1 𝐼𝑏𝑓 =
𝑔𝑐
Or
gc = 32.2 ft ∙lbm/lbf ∙ Sec 2

The constant of proportionality, gc, has both dimensions and units. The
dimensions arose because we selected both force and mass as primary
dimensions; the units (and the numerical value) are a consequence of our
choices for the standards of measurement.
Since a force of 1 lbf accelerates 1 lbm at 32.2 ft/sec2, it would accelerate
32.2 lbm at 1 ft/sec2. A slug also is accelerated at 1 ft/sec2 by a force of 1
lbf. Therefore,
1 slug = 32.2 lbm

1.1.4. Preferred Systems of Units


In this text we shall use both the SI and the British Gravitational systems of units.
In either case, the constant of proportionality in Newton's second law is
dimensionless and has a value of unity. Consequently, Newton's second law is
written as F = ma. In these systems, it follows that the gravitational force (the
"weight"4) on an object of mass, m, is given by W = mg.
SI units and prefixes, together with other defined units and useful
conversion factors, are summarized in Tables 1.1 and 1.2 at the end of this
section.

7
EXAMPLE 1.1
A mass of 100 kg is acted on by a 400-N force acting vertically upward and 600-N
force acting upward at a 45° angle. Calculate the vertical component of the
acceleration. The local vertical component of the acceleration of gravity is 9.81
m/s2.

SOLUTION OF EXAMPLE 1.1


The first step in solving a problem involving forces is to draw a free-body diagram
with all forces acting on it. It appears as follows:

Newton's second law relates the net force acting on a mass to the acceleration,
expressed as

∑Fy = ma y

Using the appropriate components, we have

400 + 600 sin 45° - 100 x 9.81 = 100a y

ay = -1.57 m/s 2

8
Table 1.1 FUNDAMENTAL DIMENSIONS AND THEIR UNITS

Table 1.2 DERIVED UNITS

9
Chapter 2: BASIC PRINCIPLES AND CONCEPTS
2.1 Fluid as a Continuum

 All fluids are composed of molecules in constant motion. We are, in most


engineering applications, interested in average or macroscopic effects of
many molecules.

 1 m 3 volume of air at STP (P = 101.3 kPa, T=15 °C ) contains 2.5 x


1025 molecules. i.e. a volume of 1 mm 3 of air at STP contains 2.5 x 1016
molecules.

 Always the number of involved molecules of the fluid is immense and


the separation between them is negligible by comparison to the distances
involved in a particular situation. That is why fluids are usually considered
as a continuum (a hypothetical continuous substance). The fluid is regarded as a
continuum in the macroscopic region.

2.1.1 Density

Defined as mass per unit volume, and may be expressed as:


𝛿𝑀 𝑑𝑀
𝜌= lim =
𝛿𝑉→𝛿𝑉 ′ 𝛿𝑉 𝑑𝑉

Figure 2.1 Definition of density at a point

2.1.2 Specific Volume

Defined as the inverse of density:

1
𝑣𝑠 =
𝜌

10
2.1.3 Pressure

We call normal stress a pressure.


Pressure: normal force per unit

δF dF
P = lim =
δA→0 δA dA

Dimensions : ML-1 T-2

Units : 1 atmosphere = 101.3 kPa

1 bar = 10 5 Pa = 10 5 N/m 2

Figure 2.2 Gage pressure and absolute pressure

Pabsolute = Patmospheric + Pgage

If Pgage is –ve, we call it Pvacume , and :

11
Pabsolute = Patmospheric - Pgage

2.1.4 Specific Weight


Specific weight is the weight per unit volume i.e. 𝛾 = 𝜌𝑔
Dimensions ML-2 T-2
Units : N/m3
For air 𝛾 = 12.07 N/m 3
For water 𝛾 = 9810 N/m3

2.1.5 Specific Gravity


Specific gravity, SG (or relative density) is defined as the ratio of the mass
density of a substance to some standard substance; water, for instance.
𝜌
𝑆𝐺 =
1000
𝑑𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑢𝑏𝑠𝑡𝑎𝑛𝑐𝑒
𝑖. 𝑒. 𝑆𝐺 =
𝑑𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑤𝑎𝑡𝑒𝑟
Units : SG has no units.

2.2 Velocity Field


Complete representation of velocity is:


𝑉 = ⃗ (𝑥, 𝑦, 𝑧, 𝑡)
𝑉

For steady flow ⃗


𝑉 = ⃗ (𝑥, 𝑦, 𝑧)
𝑉

Thus in steady flow, one property may vary from point to point in the field, but
all properties remain constant with time at every point.

One - Two -, and Three - Dimensional Flows


Flow is classified as one -, two -, or three - dimensional depending on the
number of space coordinates required to specify the velocity field.

Example of one - dimensional flow

𝑟 2
𝑢 = 𝑈𝑚𝑎𝑥 [1 − (𝑅) ]

Figure 2.3 Example of one-dimensional flow

12
Example of two- dimensional flow

Figure 2.4 Example of two-dimensional flow

Figure 2.5 Example of uniform flow at a section

Uniform flow means that the velocity is constant at any section normal to
the flow direction.

2.3 Stress Field

There are two types of forces ( or stresses ) encountered in fluid study. These are:

a) Surface Forces : act on the boundaries of a medium


through direct contact.

b) Body Forces : developed without physical contact and


distributed over the volume of fluid, i.e.
gravitational and electromagnetic forces.

𝛿 𝐹𝑛
𝜎𝑛 = lim Figure 2.6 The concept of stress in a continuum.
𝛿𝐴𝑛 →0 𝐴𝑛
𝛿

and
13
𝛿 𝐹𝑡
𝜏𝑛 = lim
𝛿𝐴𝑛 →0 𝛿 𝐴𝑛

Figure 2.7 (a) Force components and (b) stress


components, on the element of area 𝜹𝑨𝒙 .

𝛿 𝐹𝑋
𝜎𝑋𝑋 = lin
𝛿𝐴𝑋 →0 𝛿 𝐴𝑋
𝛿 𝐹𝑌
𝜏𝑋𝑌 = lim
𝛿𝐴𝑋 →0 𝛿 𝐴𝑋
𝛿 𝐹𝑍
𝜏𝑋𝑍 = lim
𝛿𝐴𝑋 →0 𝛿 𝐴𝑋

Figure 2.8 Notation for stress

𝜎𝑋𝑋 𝜏𝑋𝑌 𝜏𝑋𝑍


[𝜏𝑌𝑋 𝜎𝑌𝑌 𝜏𝑌𝑍 ]
𝜏𝑍𝑋 𝜏𝑍𝑌 𝜎𝑍𝑍
14
15
2.4 Dynamic Viscosity of a Fluid
Dynamic viscosity, 𝜇 is that property of a fluid which determines the
amount of resistance to a shearing force.
Viscosity is due primarily to interaction between fluid molecules .

2.4.1 Mathematical Modeling of Viscosity


Consider the behavior of a fluid element between the two infinite plates shown
in Figure 2.9. The upper plate moves at constant velocity 𝛿 𝑢 , under the
influence of a constant applied force 𝐹𝑋 . The shear stress, 𝜏𝑌𝑋 is given by:

Figure 2.9 Deformation of a fluid element

𝛿 𝐹𝑋 𝑑 𝐹𝑋
𝜏𝑌𝑋 = lim =
𝛿𝐴𝑦 →0 𝛿 𝐴𝑦 𝑑 𝐴𝑦
𝛿𝛼 𝑑𝛼
Deformation rate = lim =
𝛿𝑡→0 𝛿 𝑡 𝑑𝑡

Distance, 𝛿 𝑙 = 𝛿 𝑢 𝛿 𝑡
Or

𝛿𝑙= 𝛿𝑦𝛿𝛼
𝛿𝛼 𝛿𝑢
∴ =
𝛿𝑡 𝛿𝑦
𝑑𝛼 𝑑𝑢
Taking the limits, i.e. =
𝑑𝑡 𝑑𝑦

𝑑𝑢
∴ 𝜏𝑌𝑋 𝛼
𝑑𝑦

2.4.2 Newtonian Fluid


𝑑𝑢
𝜏𝑦𝑥 = 𝜇 𝑑𝑦

2.4.3 Non-Newtonian Fluid


𝑑𝑢 𝑛
𝜏𝑦𝑥 = 𝑘 (𝑑𝑦)

16
Figure 2.10 (a) Shear stress, 𝝉, and (b) apparent viscosity, 𝝅, as a function of a
deformation rate for one-dimensional flow of various non-Newtonian fluids.

2.4.4 Kinematic Viscosity


𝜇
𝛾= 𝛿

17
EXAMPLE 2.1: Viscosity and Shear Stress in Newtonian Fluid

An infinite plate is moved over a second plate on a layer of liquid as shown. For
small gap width, d, we assume a linear velocity distribution in the liquid. The liquid
viscosity is 0.65 x 10-3 kg/(m · s) and its specific gravity is 0.88. Calculate:
(a) The kinematic viscosity of the liquid, in m2/s.
(b) The shear stress on the lower plate, in Pa.
(c) Indicate the direction of each shear stress calculated in part (b).

SOLUTION OF EXAMPLE 2.1

GIVEN: Linear velocity profile in the liquid between infinite parallel plates as shown

𝜇 = 0.65 𝑥 10−3 𝑘𝑔/(𝑚. 𝑠)

𝑆𝐺 = 0.88

Find: (a)  in units of m2/s

(b)  on lower plate in units of Pa.


(c) Direction of stress in part (b).

Solution:

Basic Equation

𝑑𝑢 𝜇
𝜏𝑦𝑥 = 𝜇 𝑑𝑦 Definition:  = 𝜌
Assumption: (1) Linear velocity distribution

(2) Steady flow

(2) 𝜇 = constant
𝜇 𝜇 0.65 𝑥 10−3 𝑘𝑔 𝑚3
(a)  = = = × (0.88) 1000 𝑘𝑔
𝜌 𝑆𝐺𝜌𝐻2 𝑂 𝑚 ∙𝑠

 = 7.39 × 10−7 𝑚2 /𝑠

18
𝑈 𝑘𝑔 𝑚 1
(b) τlower = 𝜇 = 0.65 × 10−3 × 0.3 × = 0.650 𝑁/𝑚²
𝑑 𝑚 ∙𝑠 𝑠 0.3 𝑥 10−3 𝑚

(c) Direction of shear stress on upper and lower plates.

The upper plate is a negative y surface, so


positive 𝜏𝑦𝑥 acts in the negative x direction.

The lower plate is a positive y surface, so


positive 𝜏𝑦𝑥 acts in the positive x direction.

19
2.5 Fluid Compressibility
All materials are compressible i.e. the volume of a given mass will be reduced to
(V - ∂V) when a force is exerted uniformly all over its surface.
If pressure, P changes to P+∂P the relation between the change of pressure and
change of volume depends on the bulk modulus of the material
change in pressure
Bulk Modules =
volumetric strain
Since a positive increase in pressure makes the volume decrease, therefore,

1
(1/∆𝑃) 𝛼
(∆𝑉/𝑉)

1 1
𝑖. 𝑒. 1⁄∆𝑃 = − ×
𝐾 (∆𝑉/𝑉)

Figure 2.11 Schematic drawing to model fluid compressibility

Therefore, the bulk's modulus of elasticity is given by


−𝑑𝑃
𝐾 = 𝑙𝑖𝑚∆𝑉→0 − ∆𝑃/(∆𝑉/𝑉) = (𝑑𝑉/𝑉)

It may be written in terms of the mass density as follows

𝑉𝑠 = 1⁄𝜌 𝜌𝑉𝑠 = 1

Differentiating both sides we get dV + Vd = 0  dV = - V d


Hence,
−𝑑𝑉 𝑑𝜌
=
𝑉 𝜌
Therefore the bulk's modulus of elasticity of a fluid may be written as
𝑑𝜌
𝐾= (𝑁/𝑚2 )
(𝑑𝜌/𝜌)
- The more compressible the fluid the lower its bulk's modulus value.
- Dimensions : ML-1 T-2
- Units : N/m 2 Pascal

20
2.6 Reynolds Number
Reynolds number, Re is defined as:

𝜌𝑉̅ 𝐷
𝑅𝑒 =
𝜇

 : density

𝑉̅ : average flow velocity

D : Pipe diameter ( or a characteristics length)

 : Viscosity

If : Re ≤ 2300 the flow is laminar

If : Re ≥ 4100 the flow is turbulent

in between, the flow is transitional

2.7 Speed of Sound


Speed of sound, C is defined as:

𝜕𝑃
𝐶= √ )
𝜕𝜌 𝑠

For liquids and solids

𝐾
𝐶= √ K : bulk modulus
𝜌

For gases

𝐶 = √𝛾𝑅𝑇

Where,

 : specific heat ratio


R : gas constant
T : absolute temperature

2.8 Mach Number

Mach number, M is given by:

𝑉 𝑆𝑝𝑒𝑒𝑑
𝑀= =
𝐶 𝑙𝑜𝑐𝑎𝑙 𝑠𝑝𝑒𝑒𝑑 𝑜𝑓 𝑠𝑜𝑢𝑛𝑑
For M  0.3, the maximum density variation of a gas is less than 5 %

21
 M  0.3, gases can be treated as incompressible flow
( i.e.  = constant)

22
2.9 Description and classification of Fluid Motions

Coninuum Fluid
Mechanics

Inviscid
viscous
=0

Laminar Turbulent

Compressible Incompressible Internal External

Figure 2.12 Possible classifications of continuum fluid mechanics.

23
Chapter 3: FLUID STATICS
3.0 Introduction
By the definition of a fluid, shear deformation must increase
continuously when a shear stress of any magnitude is applied. The
absence of angular deformation implies the absence of shear stresses.
Therefore, fluids either at rest or in "rigid-body" motion are able to sustain
only normal stresses. Analysis of hydrostatic cases is thus appreciably
simpler than that for fluids undergoing angular deformation.
Mere simplicity does not justify our study of a subject. Normal forces
transmitted by fluids are important in many practical situations. Using the
principles of hydrostatics, we can compute forces on submerged objects,
develop instruments for measuring pressures, and deduce properties of the
atmosphere and oceans. The principles of hydrostatics also may be used to
determine forces developed by hydraulic systems in applications such as
industrial presses or automobile brakes.
In a static, homogeneous fluid, or in a fluid undergoing rigid-body motion, a
fluid particle retains its identity for all time, and fluid elements do not
deform. We may apply Newton's second law of motion to evaluate the
forces acting on the particle.

3.1 The Basic Equation of Fluid Statics


Our primary objective is to obtain an equation that will enable us to
determine the pressure field within a static fluid. To do this, we apply
Newton's second law to a differential fluid element of mass dm =  d  ,
with sides dx, dy, and dz, as shown in Figure 3.1. The fluid element is
stationary relative to the stationary rectangular coordinate system shown.

Figure 3.1 differential fluid element and pressure forces in the y direction.

From our previous discussion, recall that two general types of forces may be
applied to a fluid: body forces and surface forces. The only body force that
must be considered in most engineering problems is due to gravity. In some
situations body forces due to electric or magnetic fields might be present;
they will not be considered in this text.
For a differential fluid element, the body force. dF̅B is
d𝐹̅ B = 𝑔̅ 𝑑𝑚 = 𝑔̅  𝑑

where g̅ is the local gravity vector,  is the density, and d is the volume of
the element. In Cartesian coordinates d = dx.dy.dz, so

24
d𝐹̅ B = 𝑔̅ 𝑑𝑥 𝑑𝑦 𝑑𝑧

In a static fluid no shear stresses can be present. Thus the only surface force
is the pressure force. Pressure is a field quantity, p = p(x, y, z): the pressure
varies with position within the fluid. The net pressure force that results from
this variation can be evaluated by summing the forces that act on the six
faces of the fluid element.
Let the pressure at the center, 0, of the element be p. To determine the
pressure at each of the six faces of the element, we use a Taylor series
expansion of the pressure about the point O. The pressure at the left face of
the differential element is
𝜕𝑝 𝜕𝑝 𝑑𝑦 𝜕𝑝 𝑑𝑦
𝑝𝐿 = 𝑝 + (𝑦𝐿 − 𝑦) = 𝑝 + (− )=𝑝−
𝜕𝑦 𝜕𝑦 2 𝜕𝑦 2

(Terms of higher order are omitted because they will vanish in the
subsequent limiting process.) The pressure on the right face of the
differential element is
𝜕𝑝 𝜕𝑝 𝑑𝑦
𝑝𝑅 = 𝑝 + (𝑦𝑅 − 𝑦) = 𝑝 +
𝜕𝑦 𝜕𝑦 2

The pressure forces acting on the two y surfaces of the differential element
are shown in Figure 3.1. Each pressure force is a product of three factors.
The first is the magnitude of the pressure. The magnitude is multiplied by
the area of the face to give the magnitude of the pressure force, and a unit
vector is introduced to indicate direction. Note also in Figure 3.1 that the
pressure force on each face acts against the face. A positive pressure
corresponds to a compressive normal stress.
Pressure forces on the other faces of the element are obtained in the same
way. Combining all such forces gives the net surface force acting on the
element. Thus
𝜕𝑝 𝑑𝑥 𝜕𝑝 𝑑𝑥
𝑑𝐹𝑆 = (𝑝 − ) (𝑑𝑦 𝑑𝑧)(𝑖̂) + (𝑝 + ) (𝑑𝑦 𝑑𝑧)(−𝑖̂)
𝜕𝑥 2 𝜕𝑥 2
𝜕𝑝 𝑑𝑦 𝜕𝑝 𝑑𝑦
(𝑝 − ) (𝑑𝑥 𝑑𝑧)(𝑗̂) + (𝑝 + ) (𝑑𝑥 𝑑𝑧)(−𝑗̂)
𝜕𝑦 2 𝜕𝑦 2
𝜕𝑝 𝑑𝑧 𝜕𝑝 𝑑𝑧
(𝑝 − ) (𝑑𝑥 𝑑𝑦)(𝑘̂) + (𝑝 + ) (𝑑𝑥 𝑑𝑦)(−𝑘̂)
𝜕𝑧 2 𝜕𝑧 2
Collecting and canceling terms, we obtain
𝜕𝑝 𝜕𝑝 𝜕𝑝
𝑑𝐹𝑆 = (− 𝑖̂ − 𝑗̂ − 𝑘̂) 𝑑𝑥 𝑑𝑦 𝑑𝑧
𝜕𝑥 𝜕𝑦 𝜕𝑧
𝜕𝑝 𝜕𝑝 𝜕𝑝
𝑑𝐹𝑆 = − ( 𝑖̂ + 𝑗̂ + 𝑘̂) 𝑑𝑥 𝑑𝑦 𝑑𝑧 (3.1a)
𝜕𝑥 𝜕𝑦 𝜕𝑧

The term in parentheses is called the gradient of the pressure or simply the
pressure gradient and may be written grad  or . In rectangular
coordinates

25
𝜕𝑝 𝜕𝑝 𝜕𝑝 𝜕 𝜕 𝜕
𝑔𝑟𝑎𝑑 𝑝 ≡ ∇ p ≡ − (𝑖̂ + 𝑗̂ + 𝑘̂ ) ≡ (𝑖̂ + 𝑗̂ + 𝑘̂ ) 𝑝
𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑧

The gradient can be viewed as a vector operator; taking the gradient of a


scalar field gives a vector field. Using the gradient designation, Eq. (3.1a)
can be written as

𝑑𝐹𝑠 = −𝑔𝑟𝑎𝑑 𝑝(𝑑𝑥 𝑑𝑦 𝑑𝑧) = ∇ 𝑝 𝑑𝑥 𝑑𝑦 𝑑𝑧 (3.1 b)

Physically the gradient of pressure is the negative of the surface force per
unit volume due to pressure. We note that the level of pressure is not
important in evaluating the net pressure force. Instead, what matters is the
rate at which pressure changes occur with distance, the pressure gradient.
We shall find this term very useful throughout our study of fluid mechanics.
We combine the formulations for surface and body forces that we have
developed to obtain the total force acting on a fluid element. Thus
𝑑𝐹 = 𝑑𝐹𝑠 + 𝑑𝐹𝐵 = (−𝑔𝑟𝑎𝑑 𝑝 + 𝜌𝑔 ) 𝑑𝑥 𝑑𝑦 𝑑𝑧
= (−𝑔𝑟𝑎𝑑 𝑝 + 𝜌𝑔 ) 𝑑∀
or on a per unit volume basis
𝑑𝐹
= −𝑔𝑟𝑎𝑑 𝑝 + 𝜌𝑔 (3.2)
𝑑∀

⃗ = a⃗ dm = a⃗ρ d∀. For a


For a fluid particle, Newton's second law gives dF
static fluid, a⃗ = 0. Thus
𝑑𝐹
= 𝑝𝑎 = 0
𝑑∀
Substituting for df/d∀ from Eq. (3.2), we obtain

−𝑔𝑟𝑎𝑑 𝑝 + 𝜌𝑔 = 0 (3.3)
Let us review this equation briefly. The physical significance of each term is
−𝑔𝑟𝑎𝑑 𝑝 + 𝜌𝑔 = 0

Net pressure force per Net pressure force per


unit volume at a point
+ unit volume at a point = 0
This is a vector equation, which means that it is equivalent to three
component equations that must be satisfied individually. The components
are:
𝜕𝑝
− + 𝜌𝑔𝑥 = 0 𝑥 𝑑𝑖𝑟𝑒𝑐𝑡𝑖𝑜𝑛
𝜕𝑥
𝜕𝑝
− 𝜕𝑦 + 𝜌𝑔𝑦 = 0 𝑦 𝑑𝑖𝑟𝑒𝑐𝑡𝑖𝑜𝑛
(3.4)
𝜕𝑝
− + 𝜌𝑔𝑧 = 0 𝑧 𝑑𝑖𝑟𝑒𝑐𝑡𝑖𝑜𝑛
𝜕𝑧

26
Equations (3.4) describe the pressure variation in each of the three
coordinate directions in a static fluid. To simplify further, it is logical to
choose a coordinate system such that the gravity vector is aligned with one
of the coordinate axes. If the coordinate system is chosen such that the z axis
is directed vertically upward, then gx = 0 and gy = 0 and gz = -g. Under these
conditions, the component equations become
𝜕𝑝 𝜕𝑝 𝜕𝑝
=0 =0 = −𝜌𝑔 (3.5)
𝜕𝑥 𝜕𝑦 𝜕𝑧

Equations (3.5) indicate that under the assumptions made, the pressure is
independent of coordinates x and y; it depends on r alone. Thus since p is a
function of a single variable, a total derivative may be used instead of a
partial derivative. With these simplifications, Eqs. (3.5) finally reduce to

𝑑𝑝
= −𝜌𝑔 ≡ −𝛾 (3.6)
𝑑𝑧

Restrictions: (1) Static fluid


(2) Gravity is the only body force
(3) The z axis is vertical and upward

This equation is the basic pressure-height relation of fluid statics. It is


subject to the restrictions noted. Therefore it must be applied only where
these restrictions are reasonable for the physical situation. To determine the
pressure distribution in a static fluid, Eq. (3.6) may be integrated and
appropriate boundary' conditions applied.

Before considering specific cases that are readily treated analytically, it is


important to note that pressure values must be stated with respect to a
reference level. If the reference level is a vacuum, pressures are termed
absolute, as shown in Figure 3.2.

Atmospheric pressure
101.3 kPa
at standard sea level
Conditions

Fig. 3.2 Absolute and gage pressures, showing reference levels.

Most pressure gages read a pressure difference - the difference between the
measured pressure and the ambient level (usually atmospheric pressure).
Pressure levels measured with respect to atmospheric pressure are
termed gage pressures. Thus
P absolute = P gage + P atmosphere

27
Absolute pressures must be used in all calculations with the ideal gas
equation or other equations of state.

3.2 Pressure Variation in a Static Fluid


We have seen that pressure variation in any static fluid is described by
the basic pressure-height relation
𝑑𝑝
= −𝜌𝑔 (3.6 a)
𝑑𝑧

Although pg may be defined as the specific weight, . it has been


written as g in Eq. (3.6) to emphasize that both  and g must be
considered variables. In order to integrate Eq. (3.6 a) to find the
pressure distribution, assumptions must be made about variations in
both  and g.
For most practical engineering situations, the variation in g is
negligible. Only for a situation such as computing very precisely the
pressure change over a large elevation difference would the variation in
g need to be included. Unless we state otherwise, we shall assume g to
be constant with elevation at any given location.
In many practical engineering problems the variation in p will be
appreciable, and accurate results will require that it be accounted for.
Several types of variation are easy to treat analytically.
Pressure variation in a compressible fluid can be evaluated by
integrating Eq. (3.6). Before this can be done, density must be
expressed as a function of one of the other variables in the equation.
Property information or an equation of state may be used to obtain the
required relation for density.
The density of gases generally depends on pressure and temperature.
The ideal gas equation of state
𝑝 = 𝑅𝑇 (3.6 b)

where R is the gas constant and T the absolute temperature, accurately


models the behavior of most gases under engineering conditions.
However, the use of Eq. (3.6 b) introduces the gas temperature as an
additional variable. Therefore, an additional assumption must be made
about temperature variation before Eq. (3.6 b) can be integrated.
For an incompressible fluid, = constant. Then for constant gravity,
𝑑𝑝
= −𝜌𝑔 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
𝑑𝑧
To determine the pressure variation, we must integrate and apply
appropriate boundary conditions. If the pressure at the reference level,
Zo, is designated as po, then the pressure, p, at location z is found by
integration
𝑝 𝑧
∫ 𝑑𝑝 = − ∫ 𝜌𝑔 𝑑𝑧
𝑝𝑜 𝑧𝑜

Or
𝑝 − 𝑝0 = −𝑔 (𝑧 − 𝑧𝑜) = 𝜌𝑔 (𝑧𝑜 − 𝑧)

28
For liquids, it is often convenient to take the origin of the coordinate
system at the free surface (reference level) and to measure distances as
positive downward from the free surface. With h measured positive
downward, we have
𝑧𝑜 − 𝑧 = ℎ
and
𝑝 − 𝑝0 = 𝑔ℎ (3.7)
Equation (3.7) indicates that the pressure difference between two points
in a static fluid can be determined by measuring the elevation
difference between the two points. Devices used for this purpose are
called manometers.
Atmospheric pressure may be obtained from a barometer, in which the
height of a mercury column is measured. The measured height may be
converted to pressure using Eq. (3.7). Although the vapor pressure of
mercury may be neglected, for precise work, temper ature and elevation
corrections must be applied to the measured level and the effects of
surface tension must be considered.

Figure 3.3 U-tube manometer for measuring gage pressure at A.

A simple U-tube manometer is shown in Fig. 3.3. Since the right leg is
open to the atmosphere, measurements of h1 and h2 will allow the
determination of the gage pressure at A. Using the notation of Fig. 3.3,
and applying Eq. 3.7 between A and B and between B and C, give
𝑝𝐴 − 𝑝𝐵 = 1 𝑔 (𝑧𝐵 − 𝑧𝐴 ) = −1 𝑔ℎ1

and

𝑝𝐵 − 𝑝𝐶 = 2 𝑔 (𝑧𝐶 − 𝑧𝐵 ) = −2 𝑔ℎ2

adding the two equations f

𝑝𝐴 − 𝑝𝐶 = 2 𝑔ℎ2 − 1 𝑔ℎ1

Since p c = P atm , then p A - p c = P Agage

29
If the density 1 is negligible compared with 2 , then pA gage = 2 gh2 .
Note that the pressures at B' and B are equal because they are at the
same elevation in a continuous volume of the same fluid.
Manometers are simple and inexpensive devices used frequently for
pressure measurements. Students sometimes have trouble analyzing
multiple-tube manometer situations. The following rules of thumb are
useful:
1. Any two points at the same elevation in a continuous volume of the
same liquid are at the same pressure.
2. Pressure increases as one goes down a liquid column (remember the
pressure change on diving into a swimming pool).

Example 3.1 illustrates the use of a multiple-liquid manometer for mea-


suring a pressure difference. Because the liquid level change is small at
low pressure differential, a U-tube manometer may be difficult to read
accurately. The level change can be increased by changing the
manometer design or by using two immiscible liquids of slightly
different density. Analysis of a typical reservoir manometer design i s
illustrated in Example 3.2.

30
EXAMPLE 3.1—Multiple-Liquid Manometer

Water flows through pipes A and B. Oil. With specific gravity 0.8, is in the
upper portion of the inverted U. Mercury (specific gravity 13.6) is in the
bottom of the manometer bends. Determine the pressure difference, pA -pB ,
in units of kilopascals.

SOLUTION OF EXAMPLE 3.1

GIVEN: Multiple-tube manometer as shown. Specific gravity of oil is 0.8;


specific gravity of mercury is 13.6.

FIND : The pressure difference, pA - pB, in kPa

31
SOLUTION

𝑑𝑝 𝑑𝑝 𝜌
Basic equations: = − = −𝜌𝑔 𝑆𝐺 =
𝑑𝑧 𝑑ℎ 𝜌 𝐻2 𝑂

Assumptions: (1) Static fluid


(2) Incompressible
Then
𝑝2 ℎ2
𝑑𝑝 = 𝑝𝑔 𝑑ℎ and ∫𝑝 𝑑𝑝 = ∫ℎ 𝑝𝑔 𝑑ℎ
1 1

For 𝜌 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
𝑝2 − 𝑝1 = 𝜌𝑔(ℎ2 − ℎ1 )
Beginning at point A and applying the equation between successive points around the
manometer gives
𝑝𝐶 − 𝑝𝐴 = +𝜌𝐻2𝑂 𝑔𝑑1
𝑝𝐷 − 𝑝𝐶 = −𝜌𝐻𝑔 𝑔𝑑2
𝑝𝐸 − 𝑝𝐷 = +𝜌𝑜𝑖𝑙 𝑔𝑑3
𝑝𝐹 − 𝑝𝐸 = −𝜌𝐻𝑔 𝑔𝑑4
𝑝𝐵 − 𝑝𝐹 = −𝜌𝐻2 𝑂 𝑔𝑑5
Adding, we obtain
𝑝𝐴 − 𝑝𝐵 = (𝑝𝐴 − 𝑝𝐶 ) + (𝑝𝐶 − 𝑝𝐷 ) + (𝑝𝐷 − 𝑝𝐸 ) + (𝑝𝐸 − 𝑝𝐹 ) + (𝑝𝐹 − 𝑝𝐵 )
= −𝜌𝐻2 𝑂 𝑔𝑑1 + 𝜌𝐻 𝑔 𝑔𝑑2 − 𝜌𝑜𝑖𝑙 𝑔𝑑3 + 𝜌𝐻 𝑔 𝑔𝑑4 + 𝜌𝐻2 𝑂 𝑔𝑑5
Substituting 𝜌 = 𝑆𝐺 𝜌𝐻 𝑦𝑖𝑒𝑙𝑑𝑠
2𝑂

𝑝𝐴 − 𝑝𝐵 = 𝑔(−𝜌𝐻2 𝑂 𝑑1 + 13.6 𝜌𝐻2 𝑂 𝑑2 − 0.8 𝜌𝐻2 𝑂 𝑑3 + 13.6 𝜌𝐻2 𝑂 𝑑4 + 𝜌𝐻2 𝑂 𝑑5 )


= 𝑔𝜌𝐻2 𝑂 (−𝑑1 + 13.6 𝑑2 − 0.8 𝑑3 + 13.6 𝑑4 + 𝑑5 )
= 𝑔𝜌𝐻2 𝑂 (−250 + 1020 − 80 + 1700 + 200)mm
= 𝑔𝜌𝐻2 𝑂 x 2590 mm
9.81 𝑚 1000 𝑘𝑔
= 𝑥 𝑥 2.59 𝑚
𝑠2 𝑚3
𝑝𝐴 − 𝑝𝐵 = 25.4 kPa

32
EXAMPLE 3.2 – Reservoir Manometer
A reservoir manometer is built with a tube
diameter of 10 mm and a reservoir diam-
eter of 30 mm. The manometer liquid is
Meriam red oil. Determine the manometer
sensitivity, i.e., the deflection in millime-
ters per millimeter of water applied pres-
sure differential.

EXAMPLE PROBLEM 3.2

GIVEN: Reservoir manometer as shown

d = 10 mm
D= 30 mm

FIND : Liquid deflection, h, in millimeters per millimeter of water applied


pressure differential.

SOLUTION OF EXAMPLE 3.2


𝑑𝑝 𝜌
Basic equations: = −𝜌𝑔 𝑆𝐺 =
𝑑𝑧 𝜌 𝐻2 𝑂

Assumptions: (1) Static fluid


(2) Incompressible

Then
𝑝2 𝑍2
𝑑𝑝 = −𝜌g dz and ∫𝑝 𝑑𝑝 = − ∫𝑍 𝜌g dz
1 1

For  = constant

𝑝2 − 𝑝1 = 𝜌𝑔(𝑧2 − 𝑧1 )
𝑝1 − 𝑝2 = 𝜌𝑔(𝑧2 − 𝑧1 ) = 𝜌𝑜𝑖𝑙 𝑔(ℎ + 𝐻)
To eliminate H. note that the volume of manometer liquid must remain constant. Thus
the volume displaced form the reservoir must be the same as that which rises into the
tube.

𝜋 𝜋 𝑑 2
𝐷2 𝐻 = 𝑑2 ℎ or 𝐻 = ( ) ℎ
4 4 𝐷

Substituting gives

𝑑 2
p1 - p2 = ρoil gh [1 + ( ) ]
𝐷

33
This equation can be simplified by expressing the applied pressure differential as an
equivalent water column of height he.

𝑝1 − 𝑝2 = 𝜌𝐻2 𝑜 𝑔∆ℎ𝑒

and noting that oil = SG oil𝜌𝐻2 𝑜 . Then

2
𝑑
𝜌𝐻2 𝑜 𝑔∆ℎ𝑒 = 𝑆𝐺𝑜𝑖𝑙 𝜌𝐻2𝑜 𝑔ℎ [1 + ( ) ]
𝐷

Or

ℎ 1
=
∆ℎ𝑒 𝑆𝐺𝑜𝑖𝑙 [1 + (𝑑⁄𝐷)2 ]

For Meriam red oil, SG = 0.827. Thus, the sensitivity is

ℎ 1 ℎ
= )2 ]
= 1.09
∆ℎ𝑒 0.827[1+ (10⁄30 ∆ℎ𝑒

This problem illustrates the effects of manometer design and choice of gage
liquid on sensitivity.

34
3.3 Forces on Plane Areas
In the design of devices and objects that are submerged, such as dams, flow
obstructions, surfaces on ships, and holding tanks, it is necessary to calculate
the magnitudes and locations of forces that act on both plane and curved
surfaces. In this section we consider only plane surfaces, such as the plane
surface of general shape shown in Figure 3.4. Note that a side view is given
as well as a view showing the shape of the plane. The total force of the liquid
on the plane surface is found by integrating the pressure over the area, that
is,
𝐹 = ∫𝐴 𝑝 𝑑𝐴 (3.8)

The x and y coordinates are in the plane of the plane surface, as shown.
Assuming that p = 0 at h = 0, we know that
𝑃 = 𝛾ℎ = 𝛾ℎ sin ∝ (3.9)
where h is measured vertically down from the free surface to the elemental
area dA and y is measured from point 0 on the free surface. The force may
then be expressed as
𝐹 = ∫𝐴 𝛾ℎ 𝑑𝐴
= ∫𝐴 𝛾 sin ∝ ∫𝐴 𝛾 𝑑𝐴 (3.10)

Figure 3.4 Force on an inclined plane area

The distance to a centroid is defined as


1
𝑦̅ = ∫ 𝑦 𝑑𝐴 (3.11)
𝐴 𝐴

The expression for the force then becomes


𝐹 = 𝛾𝑦̅𝐴 sin ∝

= 𝛾ℎ̅𝐴 = 𝑝𝑐 𝐴 (3.12)

35
where h is the vertical distance from the free surface to the centroid of the
area and pt is the pressure at the centroid. Thus we see that the magnitude
of the force on a plane surface is the pressure at the centroid multiplied by
the area. The force does not, in general, act at the centroid.
To find the location of the resultant force F, we note that the sum of the
moments of all the infinitesimal pressure forces acting on the area A must
equal the moment of the resultant force. Let the force F act at the point
(xp , yp ), the center of pressure (c.p.). The value of yp can be obtained by
equating moments about the x-axis:

𝑦𝑝 𝐹 = ∫ 𝑦𝑝 𝑑𝐴
𝐴

= 𝛾 sin ∝ ∫𝐴 𝑦 2 𝑑𝐴 = 𝛾𝐼𝑥 sin ∝ (3.13)

where the second moment of the area about the .v-axis is


𝐼𝑥 = ∫𝐴 𝑦 2 𝑑𝐴 (3.14)

The second moment of an area can be determined from the second moment
of an area / about the centroidal axis by the parallel-axis-transfer theorem,
𝐼𝑥 = 𝐼̅ + 𝐴𝑦̅ 2 (3.15)

Substitute Eqs. (3.12) and (3.15) into Eq. (3.13), and obtain
𝛾(𝐼̅ + 𝐴𝑦̅ 2 ) sin ∝
𝑦𝑝 =
𝛾 𝑦̅𝐴 sin ∝

𝐼̅
= 𝑦̅ + (3.16)
𝐴𝑦̅

Centroids and moments for several areas arc presented in the exercise
sheets. Using the expression above, we can show that the force on a
rectangular gate shown in Figure 3.5,

Figure 3.5 Force on a plane area with top edge in a free surface

with the top edge even with the liquid surface, acts two-thirds of the way
down. This is also obvious considering the triangular pressure distribution
acting on the gate. Note that Eq. (3.16) shows us that yp is always greater
than y̅; that is, the resultant force of the liquid on a plane surface always
acts below the centroid of the area, except on a horizontal area for which
y̅ = ∞; then the center of pressure and the centroid coincide.
Similarly, to locate the .x-coordinate xp of the c.p., we write
36
x p F   xp dA
A

  sin   xy dA   I xy sin  (3.17)


A

where the product of inertia of the area A is


I xy  xy dA (3.18)
A

Using the transfer theorem for the product of inertia,


I xy  I xy  Ax y (3.19)

Equation (3.17) becomes


I xy
xp  x  (3.20)
Ay

We now have expressions for the coordinates locating the center of pres-
sure.

37
EXAMPLE 3.3
Find the location of the resultant force F of the water on the triangular gate and the
force P necessary to hold the gate in the position shown.

SOLUTION OF EXAMPLE 3.3


First we draw a free-body diagram of the gate, including all the forces acting on the
gate. This is shown below. We have neglected the weight of the gate.

The y-coordinate of the location of the resultant F can be found using Eq. (3.16) as
follows:
y  257

I
yp  y 
Ay

2  33 / 36
7  7.071m
3 7
To find xp we could use Eq. (3.20). Rather than that, we recognize that the resultant
force must act on a line connecting the vertex and the midpoint of the opposite side
since each infinitesimal force acts on this line. Thus using similar triangles we have
xp 2.071

l 3
x p  0.690 m

The coordinates xp and yp locate where the force due to the water acts on the gate. If we
take moments about the hinge, we can determine the force P necessary to hold the gate
in the position shown:
38
M hinge 0
3 P  (3  2.071) F

 0.929 h A
 0.929  9810  (7 sin 53)  3
Hence
P = 50 900 N or 50.9 kN

39
3.4 Forces on Curved Surfaces
We do not use a direct method of integration to find the force due to the
hydrostatic pressure on a curved surface. Rather, a free-body diagram that
contains the curved surface and the liquids directly above or below the curved
surface is identified. Such a free-body diagram contains only plane surfaces
upon which unknown fluid forces act; these unknown forces can be found as
in the preceding section.
As an example, let us determine the force of the curved gate on the stop, shown
in Figure 3.6. The free-body diagram, which includes the gate and some of the
water contained F} and F2 are due to the directly above the gate, is shown in
Figure 3.6b; the forces Fx and Fy are the horizontal and vertical components,
respectively, of the force acting on the hinge; F} and F2 are due to the
surrounding water and are the resultant forces of the pressure distributions
shown; the body force FW is due to the weight of the water shown. By summing
moments about an axis passing through the hinge, we can determine the force P
acting on the stop.
If the curved surface is a quarter circle, the problem can be greatly simplified. This
is observed by considering a free-body diagram of the gate only (see Figure
3.6c). The horizontal force FH acting on the gate is equal to F1, of Figure 3.6b,
and the component Fv is equal to the combined force F2 + FW of Figure 3.6b.
Now. FH and FV are due to the differential pressure forces acting on the circular
arc; each differential pressure force acts through the center of the circular arc.
Hence the resultant force FH + Fv (this is a vector addition) must act through the
center. Consequently, we can locate the components FH and Fv at the
center of the quarter circle, resulting in a much simpler problem.

Figure 3.6 Forces acting on a curved surface: (a) curved surface; (b) free-body
diagram of water and gate; (c) free-body diagram of gate only.

If the pressure on the free surface is p0, we can simply add a depth of liquid
necessary to provide p0 at the location of the free surface, and then work the
resulting problem, with a fictitious free surface located the appropriate distance
above the original free surface.

40
3.4.1 Hydrostatic Force on a Curved Submerged Surface
The pressure force acting on the element of area dA , is

dF   pdA
by integration
F    pdA
A

F  iˆ FX  ˆj Fy  Kˆ FZ

Where.
Fx , Fy and Fz are the components of F in the positive x, y and z respectively.

FX   d FX  iˆ.F   iˆ. d F    iˆ. pd A


A

FX    iˆ. pdAX


AX

where

dAx is the projection of d A on a plane perpendicular to the x axis


i.e. Fx is a force on a flat surface  to the x axis. Similarly, Fy and Fz.
But: FZ    pdA
AZ
Z “ vertical component”

with the free surface at atmospheric pressure, i.e. p =  g h

FZ    gh dAZ     gd     g 

1) The horizontal component is equal (in magnitude and point of application)


to the force exerted on the projection of the curved surface on a vertical
plane.

2) The vertical component is equal to the weight of a volume of fluid


occupying the space directly above the surface up to the level of a free
surface and is applied at the centroid of this volume.

41
EXAMPLE 3.3
Determine and locate the components of the force due to the water acting on
the curved area AB in Figure below, per meter of its length.

Hinge

SOLUTION OF EXAMPLE 3.3


PH = force on vertical projection CB = ghcgACB
2 4
= 9810 (1) (2×1) = 19620 N [𝑎𝑐𝑡𝑖𝑛𝑔 (2) = 𝑚 from C]
3 3

PV = Weight of water above area AB = 9810 ( ( 22/ 4) × 1)


= 30820 N
acting through the center of gravity of the volume of liquid. The center of gravity of a
quadrant of a circle is located at a distance 4R/3  from either mutually perpendicular
radius. Thus
4
Xcp   2 /   0.85 m to the left of line BC
3
Note:
Each force dP acts normal to the curve AB and would therefore pass through hinge
Coupon being extended). The total force should also pass through C. In order to
confirm this statement, take moments of the components about C, as follows.
4
M C  19620   30820  0.85  0 ( Satisfied )
3

42
EXAMPLE 3.4
A block of concrete weighs 100 kg in air and weighs only 60 kg when immersed
in fresh water. What is the average specific weight of the block?
From free body diagram
SOLUTION OF EXAMPLE 3.4
From free body diagram

W = 100 Kg

F Z  0  60  FB 100

FB  40 kg  ( f g) volume of block ()


40
∀= = 0.0041 m 3
1000× 9.81

 specific weight of block = block g

100 𝑁
𝑃𝑏𝑙𝑜𝑐𝑘 g = = 24525 3
0.0041 𝑚

3.4.2 Buoyancy Force


FB =  (p2 – p1) dAH = g (h2 – h1) dAH

= g (body volume)

Immersed body in static liquid

3.4.3 Archimedes Law


A body immersed in a fluid experiences a vertical buoyant force equal
to the weight of the fluid it displaces.

43
EXAMPLE 3.5
A spherical buoy has a diameter of 1.5 m, weighs 8.5 kN, and is anchored to the sea
floor with a cable as in the shown figure. Although the buoy normally floats on the
surface, at certain times the water depth increases so that the buoy is completely
immersed as illustrated. For this condition what is the tension of the cable?

SOLUTION OF EXAMPLE 3.5

T = FB – W
 FB = w g  =  
 3
for water  = 10.10 kN/m 3 , = d
6
N  
FB   10.110 3 3   1.5 m 3  1.785 10 4 N
 
 m  6 
Tension = T = 1.785 x 10 4 - 0.85 x 104 N = 9.35 kN

44
EXAMPLE 3.6
Referring to Figure below determine the horizontal and vertical forces due to the water
acting on the 2 m diameter cylinder per meter of its length

SOLUTION OF EXAMPLE 3.6


(a) Net PH = force on CDA - force on AB
Using the vertical projection of CDA and of AB,
P H (CDA) = 9 8 1 0 ( 1 . 2 + 0.85) (1.7 x 1)
= 34200 N
P H (AB) = 9 8 1 0 ( 1 . 2 + 1.55) (0.3 x 1 )
= 8100 N
Net PH = 34200 - 8100 = 26100 N

(b) Net Pv = upward force on DAB - downward force on DC


= weight of ( volume DABFED - volume DCGED ).
The crosshatched area (volume) is contained in each of the above volumes, one force
being upward and the other downward. Thus they cancel and
Net Net P v = weight of volume DABFGC
Dividing this volume into 3 convenient geometric shapes,
net Pv = weight of ( rectangle GFJC + triangle CJB + semicircle CDAB)

 9810 ( 1.2 x 1.4  1  1.4  1.4  1  12 )(1)


2 2
= 9810 (1.86 + 0.98 + 1.57) = 41500N

45
Chapter 4:
THE INTEGRAL FORMS OF THE FUNDAMENTAL LAWS
4.1 INTRODUCTION
Quantities of interest to engineers can often be expressed in terms of integrals. For example,
volume flow rate is the integral of the velocity over an area; heat transfer is the integral of the
heat flux over an area; force is the integral of a stress over an area; mass is the integral of the
density over a volume; and kinetic energy is the integral of V2/2 over each mass element in a
volume. There are, of course, many other integral quantities. To determine an integral quantity
the integrand must be known, or information must be available so that a good approximation to
the integrand can be made. If the integrand is not known or cannot be approximated with any
degree of certainty, appropriate differential equations must be solved yielding the needed
integrand; the integration is then performed giving the engineer the desired integral quantity.
In this chapter we present the integral quantities of interest, develop equations that relate the
integral quantities, and work a number of problems for which the integrands are given or can be
approximated. This includes a surprisingly large variety of problems. There are, however, many
integral quantities that cannot be determined since the integrands are unknown. These would
include the lift and drag on an airfoil, the torque on the blades of a wind machine, and the
kinetic energy in the wake of a submarine. To determine such integrands, it would be necessary
to solve the appropriate differential equations, a task that is often quite difficult.
Also, there are many quantities of interest that are not integral in nature. Included would be the
point of separation of the flow around a body, the concentration of a pollutant in a stream at a
certain location, the pressure distribution on the side of a building, and the wave-shore inter-
action along a lake. To study subjects such as these, it is necessary to consider the differential
equations that describe the flow situation. Most of the topics mentioned are relegated to
specialized graduate courses.
The four basic laws of fluid dynamics are:

 Conservation of Mass “Continuity Equation”.


 Conservation of Momentum.
 Conservation of Energy “First Law of Thermodynamics”.
 Second Law of Thermodynamics.

46
Approaches:

There are three common approaches to study fluid dynamics, these are:

 Experimental (Reynolds)
 Differential (small) Analysis (Euler & Lagrange)
 Integral (Control Volume) Analysis 1944

Definitions :

System : It is an identified quantity of matter surrounded by


identified boundary [M = Const.]

Control Volume : It is a definite volume defined in space with identified


boundary (known as control surface)
[Volume =  = Constant]
The integral quantities of primary interest in fluid mechanics are contained in three basic laws:
conservation of mass, first law of thermodynamics, and Newton's second law. These basic laws
are expressed in terms of a system, a fixed collection of material particles. For example, if we
consider flow through a pipe, we could identify a fixed quantity of fluid at time t as the system
(Fig. 4.1); this system would then move due to velocity to a downstream location at time t +  t.
Any of the three basic laws could be applied to this system. This is, however, not an easy task.
First let us state the basic laws in their general form.

Figure 4.1 Example of a system in fluid mechanics.

4.1.1 Conservation of Mass


The law stating that mass must be conserved is:
The mass of a system remains constant.
The mass of a fluid particle is dV, where dV is the volume occupied by the particle and p is
its density. Knowing that the density can change from point to point in the system, the
conservation of mass can be expressed in integral form as

D
Dt 
sys
d  0 (4.1.1)

Where D/Dt is used since we are following a specified group of material particles, a system.

47
4.1.2 Newton's Second Law
Newton's second law also called the momentum equation, states:
The resultant force acting on a system equals the rate at which the momentum of the system
is changing.
The momentum of a fluid particle of mass d is a vector quantity given by V d;
consequently, Newton's second law may be expressed as
D
 F  Dt  sys
V d  (4.1.2)

recognizing that both density and velocity may change from point to point in the system.
This equation reduces to  F = ma if V and  are constant throughout the entire system;  is
often a constant, but in fluid mechanics the velocity vector invariably changes from point to
point. Again, D/Dt is used to provide the rate-of-change, since Newton's second law is
applied to a system.

4.1.3 First Law of Thermodynamics


The law that relates heat transfer, work, and energy change is the first law of
thermodynamics or, more simply, the energy equation; it states:
The rate of heat transfer to a system minus the rate at which the system does work equals the
rate at which the energy of the system is changing.
Recognizing that both density and specific energy may change from point to point in the
system, it may be expressed as
. . D
Q W 
Dt 
sys
e d  (4.1.3)

where the specific energy e accounts for kinetic energy, potential energy, and internal
energy. Other forms of energy (chemical, electrical, nuclear) are not included in an
elementary course in fluid mechanics. In its basic form stated here, the first law of
thermodynamics applies only to a system, a collection of fluid particles; therefore, D/Dt is
used.

4.1.4 Second Law of Thermodynamics


If an amount of heat, Q, is transferred to a system at temperature T, the second law of
thermodynamics states that the change in entropy, dS, of the system satisfies:
dQ
dS  (4.1.4a)
T
On a rate basis we can write:
dS 1
)  Q (4.1.4b)
dt system T
Where the total entropy of the system is given by

S)   sdm  s d (4.1.4)
system mass( system) volume( system)

48
4.2 SYSTEM-TO-CONTROL-VOLUME TRANSFORMATION:
The system-to-control-volume transformation, or equivalently, the Reynolds transport theorem
is:
DN sys d
dt c.v
   d      nˆ.V dA (4.2.5)
Dt c.s

The first integral represents the rate of change of the extensive property in the control volume. The
second integral represents the flux of the extensive property across the control surface; it may be
nonzero only where fluid crosses the control surface. We study this flux term in considerable detail
in the following sections. Thus we can now express the basic laws in terms of a fixed volume in
space. We will do this in subsequent sections for each of the basic laws.
We can move the time derivative of the control volume term inside the integral since the limits
on the volume integral are independent of time (it is a fixed control volume) and write

DN sys 
   d∀    nˆ.V dA (4.2.6)
Dt c .v
t c.s

In this form we have used /t since  and  are, in general, dependent on the position
variables.

Figure 4.2 Illustration used to show of an extensive property.

4.2.1 Simplifications of the System-to-Control-Volume Transformation


Many flows of interest are steady flows, so that  () / t = 0. Our system-to-control-
volume transformation then takes the form
DN sys
 nˆ.V dA (4.2.7)
Dt C .S

Furthermore, there is often only one area A1 across which fluid


enters the control volume and one area A2 across which fluid
leaves the control volume; assuming that the velocity vector is
normal to the area (see Figure 4.3), we can write 𝑛̂• V 1 = -V 1
over area A1 and 𝑛̂ • V2 = V 2 over area A 2 . Then Eq. 4.2.7
becomes

Figure 4.3 Flow into and from


49 a device
DN sys
   2  2 V2 dA -   11V1 dA (4.2.8)
Dt A2 A1

Finally, there are many situations that are modeled acceptably by assuming uniform
properties over each plane area (see Fig. 3.9); then the equation simplifies to

DN sys
 2  2 V2 A 2 - 11V1 A1 (4.2.9)
Dt

We will find that the system-to-control-volume transformation in this simplified form is


most often used in the application of the basic laws to problems of interest in the
introductory course in fluid mechanics. Some applications will, however, be included
that will illustrate nonuniform distributions and unsteady flows.

If we generalize Eq. (4.2.9) to include several areas across which the fluid flows, we
could write

DN sys N
  i  i Vi .nˆ i A i (4.2.10)
Dt i 1

where N is the number of areas. The dot product n̂ .V would provide us with the
appropriate sign at each area; for an inlet area, n̂. V introduces a negative sign, and for an
exit area, n̂.V introduces a positive sign.

For an unsteady flow in which flow properties are assumed to be uniform throughout the
control volume, the system-to-control-volume equation takes the form

DN sys d  
 C .V .  2  2 V2 A 2 - 11V1 A1 (4.2.11)
Dt dt

for one inlet and one outlet with uniform properties.

4.3 CONSERVATION OF MASS


A system is a given collection of fluid particles; hence its mass remains fixed:

Dm sys D
Dt

Dt 
sys
 0 (4.3.1)

In Eq. 4.1.6, Nsys represents the mass of the system, so we simply let  = 1. Thus the
conservation of mass, referring to Eq. (4.2.5), becomes
d
dt c.v
0  d   nˆ.V dA (4.3.2)
C .S

Or, if we prefer,


0 c . v t
d   nˆ.V dA
C .S
(4.3.3)

Either of the equations above is called the continuity equation.


50
If the flow is steady, there results

 C .S
nˆ.V dA  0 (4.3.4)

which, for a uniform flow with one entrance and one exit, takes the form
 2 A 2 V2  1 A1V1 (4.3.5)
where we have used 𝑛̂. V = - V 1 and 𝑛̂.V 2 = V 2 .
If the density is constant in the control volume, the derivative /t = 0 even if the flow is
unsteady. The continuity equation (4.3.3) then reduces to
A1V1  A2V2 (4.3.6)
This form of the continuity equation is used quite often, particularly with liquids and low-speed
gas flows.
Before presenting some examples applying the continuity equation, two fluxes are defined that
will be useful in specifying the quantity of flow.

Figure 4.4 Nonuniform velocity profile.


.
The mass flux m , the mass rate of flow, is
.
m   Vn d A (4.3.7)
A

and has units of kg/s (slug/sec); Vn is the normal component of velocity. The flow rate Q, the
volume rate of flow2, is
Q   Vn d A
A (4.3.8)
3 3
and has units of m /s (ft /sec). The mass flux is usually used in specifying the quantity of
flow for a compressible flow and the flow rate for an incompressible flow.
In terms of average velocity, we have

Q  AV (4.3.9)
.
m  AV (4.3.10)

where for the mass flux we assume a uniform density profile; we also assume that the
velocity is normal to the area.

51
EXAMPLE 4.1

Water flows in and out of a device as shown. Calculate the rate of change of the mass of water
(dm/dt) in the device if V1=12 m/s.
v

SOLUTION OF EXAMPLE 4.1

The control volume selected is shown. For the control surface surrounding the device, the
continuity equation, with three surfaces across which water flows, ca be written as (Eq. 4.3.2)
d
dt c.v.
0  d     nˆ.V d A
c .s.

or
dm
0 - 1 A1 V1   2 A 2 V2   3 A 3 V3
dt
where we have used 𝑉1 . 𝑛̂1 = −𝑉1 since 𝑛̂1 points out of the volume, opposite to the direction
of V 1 . In terms of the quantities given, the above can be ex pressed as
dm .
0 - 1 A1 V1  m2   3Q3
dt
Or

-1000    12  0.02 2  20  1000  0.01


dm

dt
This is solved to yield
dm
  14.9 kg / s
dt

52
EXAMPLE 4.2
A 1.4-m3 container is filled with pressurized air at 20°C. At time t = 0 air escape out of a small 0.1-
cm2 tube in the side of the tank. The velocity out of the tube may be approximated by
V  2 p  patm  /  , where  is the absolute tank pressure. Determine the time at which the
pressure (gage) in the tank is 20 kPa if the initial pressure is 200 kPa (gage). Assume that the air
temperature remains at 20°C.

SOLUTION OF EXAMPLE 4.2

The control volume is shown with the control surface coinciding with the intern wall. The air
density is uniform; therefore, the continuity equation (4.3.2) applied to this control volume is
d 
  V1 A1  0
dt

or, since  is constant,


d
   V1 A1  0
dt
The velocity is given as a function of pressure, while the density in the continued equation may be
expressed in terms of pressure using the ideal gas law p = RT.
Thus, substituting for ,
 dp pA1 2  p  p atm RT
 0
RT dt RT p

This is a first-order differential equation with  as the unknown dependent variable. It can be
written
dp A 2 RT
 1 dt
p p  patm  

Using the initial condition p = 200 000 Pa (gage) or p = 301 000 Pa absolute, we integrate until p =
20 000 Pa (gage) or 121 000 Pa absolute. Using integral tables, we find that

2 In  p   p  p atm 301000
121000

A1 2 RT

t

Using  = 1.4 m 3 , A1 = 10 -5 m2 , T =273 +20 = 293 K, R = 287 J/kg-K, this takes the form

2 In  121000  20000  2 In  301000  


200000   0.00293 t

Solving for t, we find that t = 485 s


53
EXAMPLE 4.3

This example shows that there may be more than one good choice for a control volume. We want to
determine the rate at which the water level rises in an open container if the water coming in through
a 0.10-m2 pipe has a velocity of 0.5 m/s and the flow rate going out is 0.2 m3/s. The container has a
circular cross section with a diameter of 0.5 m.

SOLUTION OF EXAMPLE 4.3

First we select a control volume that extends above the water surface as shown. Apply the
continuity equation (Eq. 4.3.2)

 d     V1  A1   V2 A2  0
d
dt 
C .V

In which the first term describes the rate of change of water mass in the control volume. Hence
𝑑 (𝜌 𝜋𝐷2 ℎ/4)
− 𝜌 𝑉1 𝐴1 + 𝜌𝑄2
𝑑𝑡

Or, dividing by ,
D 2 dh
V1 A1  Q2  0
4 dt

The rate at which the water level is rising is then


dh V1 A1  Q2

dt D 2 / 4

Thus
dh 0.5  0.1  0.2
   0.764 m/s
dt   0.5 2 / 4

54
4.4 CONSERVATION OF MOMENTUM (MOMENTUM EQUATION)
4.4.1 General Equation
Newton's second law, often called the momentum equation, states that the resultant force
acting on a system equals the rate of change of momentum of the system when measured in an
inertial reference frame; that is
D
 F  Dt  sys
V d  (4.4.1)

Using Eq. 4.2.9, with  replaced by V, this is written for a control volume as

 V V .nˆ  d A
d
 F  dt  C .V
V d   
C .S
(4.4.2)

Where, the quantity in parentheses is simply a scalar for each differential area dA.
When applying Newton's second law, the quantity  F represents all forces acting on the
control volume. The forces include the surface forces resulting from the surroundings
acting on the control surface and body forces that result from gravity and magnetic fields.
The momentum equation is often used to determine the forces induced by the flow. For
example, the equation allows us to calculate the force on the support of an elbow in a
pipeline or the force on a submerged body in a free-surface flow.

4.4.2 Notes on Momentum Equation


Force = Surface Force + Body Force = 𝐹⃑𝑠 + 𝐹⃑𝐵
Surface Force = Forces From Normal stress (Pressure)
+ " " tangential " (shear)
+
External Farces on the control volume
  
  pdA   dAt  R
CS CS

Body Force = Forces due to gravity and electromagnetic fields. In the absence of
electromagnetic field, this becomes:


  gd
CV

All velocities and time derivatives are measured relative to control volume, [i.e. have
your axes fixed to control volume.]

55
EXAMPLE 4.4 - Flow through Elbow: Use of Gage Pressures

Water flows steadily through the 90° reducing elbow shown in the
diagram. At the inlet to the elbow, the absolute pressure is 221
kPa and the cross-sectional area is 0.01 m2. At the outlet, the
cross-sectional area is 0.0025 m2 and the velocity is 16 m/sec.
The pressure at the outlet is atmospheric. Determine the force
required to hold the elbow in place.

SOLUTION OF EXAMPLE 4.4

GIVEN: Steady flow of water through 90° reducing elbow.

p 1 =221 kPa (abs) A 1 =0.01 m 2



V   16 ˆj m / sec A 2 =0.0025 m 2

FIND: Force required to held elbow in place.

SOLUTION:
Choose control volume as shown by the dashed line.
Rx and Ry are components of force required to hold the elbow in place. (They are assumed positive
and are forces acting on CV.)
A3 is the area of vertical sides of CV excluding A1; Avertical sides =A1 + A3.
A4 is the area of horizontal sides of CV excluding A2; Ahorizontal sides = A2 +A4.
Basic equations:

Momentum:
= 0 (4)
       
F  FS  FB 
t CV
V d    VV .dA
CS

Continuity:
= 0 (4)
  
0
t 
CV
 d    V .dA
CS

Assumptions:

(1) Uniform flow at each section


(2) Atmospheric pressure, pa = 101 kPa
(3) Incompressible flow
(4) Steady flow

56
Writing the x component of the momentum equation results in
   
FS X   uV .d A   uV .d A {FS X  0 and u 2  0}
CS A1

Pressure over right side Pa.


 
p1 A1  pa A3  pa ( A1  A3 )  R   uV .d A Pressure over left side is P1
A1
on A1 and Pa on A3

(𝑝1 − 𝑝𝑎 ) 𝐴1 + 𝑅𝑥 = ∫𝐴 𝑢 {−|𝜌 𝑉1 𝑑𝐴|} ⃗⃑ . 𝑑𝐴⃑ 𝑖𝑠 𝑛𝑒𝑔𝑎𝑡𝑖𝑣𝑒 𝑎𝑡 𝐴1 }


{𝑉
1

𝑅𝑥 = − 𝑝1 𝐴1 − 𝑢1 |𝜌 𝑉1 𝐴1 | { p1 - pa = p1 gage

To find V1, use the continuity equation:


     
 
cs
V .dA  0   V .dA  .dA
A1
V
A2

0    V .dA   V .dA   V1 A1  V2 A2


A1 A2

and

A2 m 0.0025 
V1  V2 16   4 m sec V1  4iˆ m sec
A1 sec 0.01
𝑅𝑥 = − 𝑝1𝑔 𝐴1 − 𝑢1 |𝜌 𝑉1 𝐴1 |

2
N m kg m 2 N . sec
 1.20 10 5  0.01 m 2
 4 999 4  0.01m
m2 sec m 3 sec kg.m

{R to hold elbow acts to left} R x


R x   1.36 kN x 

Writing the y component of the momentum equation gives.


   
FS y  FBy   V .dA   V .dA {1  0}
cs A2

Pr essure is p a over top and 


p a A4  p a A2  p a A4  p a A2  FB y  R y    V dA     
bottom of CV .V .dA is positive at 2
A2

RB y  Ry  2 V2 A2

Since we do not knowthe volume or mass of 


 
R  F   2 V A the elbow, we cannot evaluate F 
y B 2 2 B
y 
 y 

Substituting numbers, recognizing 𝑉̅ = −16 𝑗 ́ 𝑚/𝑠𝑒𝑐, so 𝑉2 = −16 𝑚/𝑠𝑒𝑐

 m  kg  m  N .sec 2
Ry   FB y    16  999   16  0.0025m 2

 sec  m3  sec  kg.m

57
Ry   FBy  639 N

Neglecting 𝐹𝐵𝑦

{R to hold elbow acts to down} R


y y
R   639 N             
y

This problem illustrates the application of the momentum equation to an inertial control volume in which
the pressure is not atmospheric across the entire control surface.
Since the pressure forces over the entire control surface must be included in the analysis, use of gage
pressure on all surfaces gives correct (and often more direct) results.

58
EXAMPLE 4.5: Conveyor Belt Filling: Rate of Change of Momentum in Control Volume

A horizontal conveyor belt moving at 0.9 m/sec receives sand from a hopper. The sand falls
vertically from the hopper to the belt at a speed of 1.5 m/sec and a flow rate of 250 Kg/sec (the
density of sand is approximately 1600 Kg/m3). The conveyor belt is initially empty but begins to fill
with sand. If friction in the drive system and rollers is negligible, find the tension required to pull
the belt while the conveyor is filling.

SOLUTION OF EXAMPLE 4.5

GIVEN: Conveyor and hopper shown in sketch

FIND: Tbelt at the instant shown.

SOLUTION:

Use the control volume and coordinates shown. Apply the x component of the momentum equation.
Basic equations:
= 0 (2)
𝜕 𝜕
𝐹𝑆𝑥 + 𝐹𝐵𝑥 = ̅ 𝑑𝐴̅
∫ 𝑢𝜌𝑑∀ + ∫ 𝑢𝜌𝑉. 0= ̅ 𝑑𝐴̅
∫ 𝜌𝑑∀ + ∫ 𝜌𝑉.
𝜕𝑡 𝐶𝑉 𝐶𝑆 𝜕𝑡 𝐶𝑉 𝐶𝑆

Assumptions: (1) 𝐹𝑆𝑥 = 𝑇𝑏𝑒𝑙𝑡 = 𝑇


(2) 𝐹𝐵𝑥 = 0
(3) Uniform flow at section (1)
(4) All sand on belt moves with Vbelt = Vb
Then

𝜕
𝑇= ∫ 𝑢𝜌𝑑∀ + 𝑢1 {−|𝜌 𝑉1 𝐴1 |} + 𝑢2 {−|𝜌 𝑉2 𝐴2 |}
𝜕𝑡 𝐶𝑉
𝜕
Since u1 = 0, and there is no flow at section (2), then 𝑇 = 𝜕𝑡 ∫𝐶𝑉 𝑢𝜌𝑑∀ From assumption (4),
inside the CV, u = Vb = constant, and hence
𝜕 𝜕𝑀𝑠
𝑇 = 𝑉𝑏 ∫ 𝜌𝑑∀ = 𝑉𝑏
𝜕𝑡 𝐶𝑉 𝜕𝑡

where Ms is the mass of sand on the belt (inside the control volume). From the continuity equation,

59
𝜕 𝜕
∫ 𝜌𝑑∀ = ̅ 𝑑𝐴̅ = 𝑚̇ 𝑠 = 250 𝐾𝑔/𝑠𝑒𝑐
𝑀 = − ∫ 𝜌𝑉.
𝜕𝑡 𝐶𝑉 𝜕𝑡 𝑠 𝐶𝑆

Then

𝑚 𝐾𝑔 𝑁. 𝑠𝑒𝑐2
𝑇 = 𝑉𝑏 𝑚𝑠 = 0.9 × 250 ×
𝑠𝑒𝑐 𝑠𝑒𝑐 𝐾𝑔. 𝑚
𝑇
𝑇 = 225 𝑁 ←

{This problem illustrates an application of the momentum equation to a problem in which the rate of
change of momentum within the control volume is not equal to zero}

60
4.4.3 Control Volume Moving with Constant Velocity

EXAMPLE 4.6: Vane Moving with Constant Velocity


The sketch shows a vane with a turning angle
of 60°. The vane moves at constant speed,
U = 10 m/sec, and receives a jet of water that
leaves a stationary nozzle with speed V = 30
m/sec. The nozzle has an exit area of 0.003
m2. Determine the force that must be applied
to maintain the vane speed constant.

SOLUTION OF EXAMPLE 4.6

GIVEN: Vane, with turning angle  = 60°, moves with constant velocity, 𝑈⃗⃑ = 10 𝑖̂ m/sec.
2
Water from a constant-area nozzle, A= 0.003 m , with velocity 𝑉 ⃗⃑ = 30 𝑖̂ m/sec,
flows over the vane as shown.
Find : The force that must be applied to
maintain the vane speed constant
= 60

SOLUTION:
⃗⃑, as shown by the
Select a control volume moving with the vane at constant velocity, 𝑈
dashed lines. Rx and Ry are the components of force required to maintain the velocity of
the control volume at 10 𝑖̂ m/sec.
The control volume is inertial, since it is not accelerating (U = constant). Remember that
all velocities must be measured relative to the control volume in applying the basic
equations.
Basic equations:
𝜕
𝐹⃑𝑠 + 𝐹⃑𝐵 = ⃗⃑𝑥𝑦𝑧 𝜌 𝑑 ∀ + ∫ 𝑉
∫𝐶𝑉 𝑉 ⃗⃑ 𝜌𝑉⃗⃑ . 𝑑𝐴⃑
𝜕𝑡 𝐶𝑆 𝑥𝑦𝑧 𝑥𝑦𝑧

𝜕
0= ⃗⃑𝑥𝑦𝑧 . 𝑑𝐴⃑
∫ 𝜌 𝑑 ∀ + ∫ 𝜌𝑉
𝜕𝑡 𝐶𝑉
𝐶𝑆

Assumptions: (1) Flow is steady relative to the vane


(2) Magnitude of relative velocity along the vane is constant:
⃗⃑1 | = |𝑉
|𝑉 ⃗⃑2 | = 𝑉 − 𝑈

61
(3) Properties are uniform at sections (1) and (2)
(4) 𝐹𝐵𝑥 =0
(5) Incompressible flow
The x component of the momentum equation is
= 0(4) = 0(1)
𝜕
𝐹𝑆𝑥 + 𝐹𝐵𝑥 = ⃗⃑𝑥𝑦𝑧 . 𝑑𝐴⃑
∫ 𝑢 𝜌𝑑∀ + ∫ 𝑢𝑥𝑦𝑧 𝜌𝑉
𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧 𝐶𝑆

There is no net pressure force, since p atm acts on all sides of the CV. Thus

𝑅𝑥 = ∫ 𝑢 { − |𝜌𝑉𝑑𝐴|} + ∫ 𝑢 { |𝜌𝑉𝑑𝐴|} = −𝑢1 |𝜌𝑉1 𝐴1 | + 𝑢2 |𝜌𝑉2 𝐴2 |


𝐴1 𝐴2

(All velocities are measured relative to xyz.) From the continuity equation

0 = ∫ { − |𝜌𝑉𝑑𝐴|} + ∫ { |𝜌𝑉𝑑𝐴|} = −|𝜌𝑉1 𝐴1 | + |𝜌𝑉2 𝐴2 |


𝐴1 𝐴2

Or

|𝜌𝑉1 𝐴1 | + |𝜌𝑉2 𝐴2 |

Therefore,

𝑅𝑥 = (𝑢2 − 𝑢1 ) |𝜌𝑉1 𝐴1 |

All velocities must be measured relative to the CV, so we note that


V1 = V – U V2 = V - U
u1 = V – U u2 = (V – U) cos
Substituting yields
𝑅𝑥 = [(𝑉 − 𝑈) cos 𝜃 − (𝑉 − 𝑈)]|𝜌 (𝑉 − 𝑈)𝐴1 | = (𝑉 − 𝑈) (cos 𝜃 − 1) |𝜌 (𝑉 − 𝑈)𝐴1 |
𝑚 𝑘𝑔 𝑚 𝑁. 𝑠𝑒𝑐 2
= (30 − 10) (0.50 − 1) |999 3 (30 − 10) × 0.003𝑚2 |
𝑠𝑒𝑐 𝑚 𝑠𝑒𝑐 𝑘𝑔. 𝑚
Rx = - 599 N {to the left}

Writing the y component of the momentum equation, we obtain


= 0(1)
𝜕
𝐹𝑆𝑦 + 𝐹𝐵𝑦 = ⃗⃑𝑥𝑦𝑧 . 𝑑𝐴⃑
∫ 𝜐 𝜌𝑑∀ + ∫ 𝑣𝑥𝑦𝑧 𝜌𝑉
𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧 𝐶𝑆
Denoting the mass of the CV as M, then
All velocities are
⃗⃑ . 𝑑𝐴⃑ = ∫ 𝜐𝜌𝑉
𝑅𝑦 − 𝑀𝑔 = ∫𝐶𝑆 𝜐𝜌𝑉 ⃗⃑ . 𝑑𝐴⃑ {1 = 0} measured relative to
𝐴 2
xyz

= ∫𝐴 𝜈 |𝜌𝑉𝑑𝐴| = 𝜈2 |𝜌𝑉2 𝐴2 | = 𝜈2 |𝜌𝑉1 𝐴1 | {Recall |𝜌𝑉2 𝐴2 | = |𝜌𝑉1 𝐴1 |}


2

= (V – U) sin  |(V – U)A1|

62
𝑚 𝑘𝑔 𝑚 𝑁. 𝑠𝑒𝑐 2
= (30 − 10) (0.866) |999 3 (30 − 10) × 0.003𝑚2 |
𝑠𝑒𝑐 𝑚 𝑠𝑒𝑐 𝑘𝑔. 𝑚

Thus the vertical force (in addition to the weight of the vane and water within the CV)
required to maintain the motion is
Ry – Mg = 1.04 kN {upward}

Then the net force required to maintain the constant vane speed is
⃗⃗⃗⃑ = −0.599 𝑖̂ + 1.04 𝑗̂ 𝑘𝑁
𝑅 ⃗⃗⃗⃑
𝑅
{This problem illustrates the fact that in applying the momentum equation to an inertial
control volume all velocities must be measured relative to the control volume}

4.4.4 Momentum Equation For Control Volume Moving with a Rectilinear Acceleration

∴ ∑ 𝐹̅ = 𝑚𝑎̅𝑖

with
𝑎̅𝑖 : acceleration measured from (inertial) fixed frame of reference XYZ.

𝑎̅𝑖 = 𝑎̅𝑥𝑦𝑧 + 𝑎̅𝑟𝑓


Where

𝑎̅𝑟𝑓 : rectilinear acceleration of non inertial frame of reference xyz moving relative
to inertial frame of reference XYZ.

∴ ∑ 𝐹̅ − 𝑚𝑎̅𝑟𝑓 = 𝑚𝑎̅𝑥𝑦𝑧

therefore, Reynolds Transport theorem gives

𝜕
𝐹̅𝑠 + 𝐹̅𝐵 − ∫𝐶𝑉 𝑎̅𝑟𝑓 𝜌𝑑∀ = ∫𝐶𝑉 𝑉̅ 𝜌𝑑∀ + ∫𝐶𝑆 𝑉̅ 𝜌𝑛̂𝑉̅ 𝑑𝐴 (4.4.3)
𝜕𝑡

where all velocities and time derivatives are measured from a frame of reference [axes]
moving with the control volume.
The scalar components of the above vector equation are:
𝜕
𝐹𝑠𝑥 + 𝐹𝐵𝑥 − ∫ 𝑎𝑟𝑓𝑥 𝜌𝑑∀ = ∫ 𝑢 𝜌𝑑∀ + ∫ 𝑢𝑥𝑧𝑦 𝜌𝑉̅𝑥𝑧𝑦 . 𝑑𝐴̅
𝐶𝑉 𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧 𝐶𝑆
𝜕
𝐹𝑠𝑦 + 𝐹𝐵𝑦 − ∫ 𝑎𝑟𝑓𝑦 𝜌𝑑∀ = ∫ 𝜐 𝜌𝑑∀ + ∫ 𝜐𝑥𝑧𝑦 𝜌𝑉̅𝑥𝑧𝑦 . 𝑑𝐴̅
𝐶𝑉 𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧 𝐶𝑆
𝜕
𝐹𝑠𝑧 + 𝐹𝐵𝑧 − ∫ 𝑎𝑟𝑓𝑧 𝜌𝑑∀ = ∫ 𝑤 𝜌𝑑∀ + ∫ 𝑤𝑥𝑧𝑦 𝜌𝑉̅𝑥𝑧𝑦 . 𝑑𝐴̅
𝐶𝑉 𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧 𝐶𝑆

63
EXAMPLE 4.7: Vane Moving with Rectilinear Acceleration
A vane, with turning angle  = 60°, is attached to a cart. The cart and vane, of mass M = 75 kg,
roll on a level track. Friction and air resistance may be neglected. The vane receives a jet of
water, which leaves a stationary nozzle horizontally at V = 35 m/sec. The nozzle exit area is A =
0.003 m2 . Determine the velocity of the cart as a function of time and plot the results.

SOLUTION OF EXAMPLE 4.7


GIVEN: Vane and cart as sketched, with M = 75 kg.

FIND; (a) U (t)

(b) plot results.

SOLUTION:
Choose the control volume and coordinate systems shown for the analysis. Note that XY is a
fixed frame, while frame xy moves with the cart. Apply the x component of the momentum
equation.
Basic equation:
= 0(1) = 0(2)  0(3)
𝜕
𝐹𝑠𝑥 + 𝐹𝐵𝑥 − ∫ 𝑎𝑟𝑓𝑥 𝜌𝑑∀ = ∫ 𝑢 𝜌𝑑∀ + ∫ 𝑢𝑥𝑧𝑦 𝜌𝑉̅𝑥𝑧𝑦 . 𝑑𝐴̅
𝐶𝑉 𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧 𝐶𝑆

Assumptions: (1) 𝐹𝑠𝑥 =0, since no resistance is present


(2) FBx =0

(3) Neglect the mass and rate of change of uxyz for water in contact with
the vane compared to the cart mass
𝜕
∫ 𝑢 𝜌𝑑∀ ≅ 0
𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧
(4) Uniform flow at sections (1) and (2)
(5) Water stream is not slowed by friction on the vane, so |𝑉̅𝑥𝑦𝑧1 | = |𝑉̅𝑥𝑦𝑧2 |
(6) A2 = A1 = A

Then

− ∫ 𝑎𝑟𝑓𝑥 𝜌𝑑∀ = 𝑢𝑥𝑦𝑧1 {− |𝜌𝑉𝑥𝑦𝑧1 𝐴1 |} + 𝑢𝑥𝑦𝑧2 {− |𝜌𝑉𝑥𝑦𝑧2 𝐴2 |}


𝐶𝑉

where all velocities must be measured relative to the xyz frame. Dropping subscripts rf and
xyz, we obtain
− ∫𝐶𝑉 𝑎𝑥 𝜌𝑑∀ = 𝑢1 {−|𝜌𝑉1 𝐴1 |} + 𝑢2 {−|𝜌𝑉2 𝐴2 |}

64
Evaluating these terms separately gives
𝑑𝑈
− ∫ 𝑎𝑥 𝜌𝑑∀ = − 𝑎𝑥 𝑀𝐶𝑣 = −𝑎𝑥 𝑀 = − 𝑀
𝐶𝑉 𝑑𝑡

𝑢1 {−|𝜌𝑉1 𝐴1 |} = (𝑉 − 𝑈) {– |𝜌 (𝑉 − 𝑈)𝐴|} = − 𝜌 (𝑉 − 𝑈)2 𝐴

𝑢2 {|𝜌𝑉2 𝐴2 |} = (𝑉 − 𝑈) cos 𝜃 {|𝜌 (𝑉 − 𝑈)𝐴|} = 𝜌 (𝑉 − 𝑈)2 𝐴 cos 𝜃

Absolute value signs have been dropped from the flux terms, since V≥U. Substitution into Eq.
(4.7.1) gives

𝑑𝑈
−𝑀 = − 𝜌 (𝑉 − 𝑈)2 𝐴 + 𝜌 (𝑉 − 𝑈)2 𝐴 cos 𝜃
𝑑𝑡
Or
𝑑𝑈
−𝑀 = (cos 𝜃 − 1) 𝜌 (𝑉 − 𝑈)2 𝐴
𝑑𝑡
Separating variables, we obtain

𝑑𝑈 (1 − 𝑐𝑜𝑠𝜃)𝜌𝐴 (1 − 𝑐𝑜𝑠𝜃)𝜌𝐴
= 𝑑𝑡 = 𝑏 𝑑𝑡 𝑤ℎ𝑒𝑟𝑒 𝑏 =
(𝑉 − 𝑈 )2 𝑀 𝑀

Note that since V = constant, d U = – d (V – U). Integrating between limits U = 0 at t= 0, and U = U


at t = t,
𝑈 𝑈 𝑡
𝑑𝑈 − 𝑑 (𝑉 − 𝑈) 1 𝑈
∫ = ∫ = ] = ∫ 𝑏 𝑑𝑡 = 𝑏𝑡
0 ( 𝑉 − 𝑈)
2
0 (𝑉 − 𝑈)2 ( 𝑉 − 𝑈) 0 0

Or
1 1 𝑈
− = 𝑏𝑡
(𝑉−𝑈) 𝑉 𝑉 (𝑉 − 𝑈)
Solving for U, we obtain
𝑈 𝑉 𝑏𝑡
=
𝑉 1+𝑉𝑏𝑡
Evaluating the term Vb gives
( 1 − cos 𝜃) 𝜌𝐴
𝑉𝑏 = 𝑉
𝑀

𝑚 (1 − 0.5) 𝑘𝑔
𝑉𝑏 = 35 × × 999 3 × 0.003 𝑚2 = 0.699 𝑠𝑒𝑐 −1
𝑠𝑒𝑐 75 𝑘𝑔 𝑚

65
Thus
𝑈 0.699𝑡
= (t in sec) U (t)
𝑉 1+0.699𝑡

{As the vane speed, U, nears the jet speed, V, the mass flow rate crossing the control surface
decreases toward zero. The plot shows the corresponding decrease in vane acceleration.}

66
EXAMPLE 4.8: Rocket Directed Vertically
A small rocket, with an initial mass of 400 kg, is to be launched vertically. Upon ignition the rocket
consumes fuel at the rate of 5 kg/sec and ejects gas at atmospheric pressure with a speed of 3500
m/sec relative to the rocket. Determine the initial acceleration of the rocket and the rocket speed
after 10 sec, if air resistance is neglected.

SOLUTION OF EXAMPLE 4.8


GIVEN: Small rocket accelerates vertically from rest. Air resistance may be neglected. Rate of
fuel consumption, thmt = 5 kg/sec. Exhaust velocity, Ve =3500 m/sec, relative to rocket,
leaving at atmospheric pressure.

FIND: (a) Initial acceleration of the rocket,

(b) Rocket velocity after 10 sec.

SOLUTION:
Choose a control volume as shown by dashed lines. Because the control volume is accelerating,
define inertial coordinate system XY and coordinate system xy attached to the CV. Apply the y
component of the momentum equation.

Basic equation:
𝜕
𝐹𝑠𝑦 + 𝐹𝐵𝑦 − ∫ 𝑎𝑟𝑓𝑦 𝜌𝑑∀ = ∫ 𝑣 𝜌𝑑∀ + ∫ 𝑣𝑥𝑧𝑦 𝜌𝑉̅𝑥𝑧𝑦 . 𝑑𝐴̅
𝐶𝑉 𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧 𝐶𝑆

Assumptions: (1) Atmospheric pressure acts on all surfaces of the CV; since air
resistance is neglected, then 𝐹𝑠𝑦 = 0
(2) Gravity is the only body force; g is constant
(3) Flow leaving the rocket is uniform and Ve is constant
Under these assumptions the momentum equation reduces to
𝜕
𝐹𝐵𝑦 − ∫𝐶𝑉 𝑎𝑟𝑓𝑦 𝜌𝑑∀ = 𝜕𝑡
∫𝐶𝑉 𝜐𝑥𝑦𝑧 𝜌𝑑∀ + ∫𝐶𝑆 𝜐𝑥𝑧𝑦 𝜌𝑉̅𝑥𝑧𝑦 . 𝑑𝐴̅ (4.8-1)

A B C D
Let us look at the equation term by term:

A 𝐹𝐵𝑦 = − ∫𝐶𝑉 g𝜌𝑑∀ = −g ∫𝐶𝑉 𝜌𝑑∀ = −g𝑀CV {Since g is constant}

The mass of the CV will be a function of time because mass is leaving the CV at rate 𝑚̇𝑒 . To
determine MCv as a function of time, we use the conservation of mass equation

67
𝜕
∫ 𝜌𝑑∀ + ∫ 𝜌𝑉̅ . 𝑑𝐴̅ = 0
𝜕𝑡 𝐶𝑉 𝐶𝑆

Then
𝜕
∫ 𝜌𝑑∀ = − ∫ 𝜌𝑉̅ . 𝑑𝐴̅ = − ∫ 𝜌𝑉̅ . 𝑑𝐴̅ = − ∫ {|𝜌𝑉 𝑑𝐴|} = −|𝑚̇𝑒 |
𝜕𝑡 𝐶𝑉 𝐶𝑆 𝐴𝑒 𝐴𝑒

The minus sign indicates that the mass of the CV is decreasing with time. Since the mass of the CV
is only a function of time, we can write
𝑑 𝑀𝐶𝑉
− |𝑚̇𝑒 |
𝑑𝑡
To find the mass of the CV at any time, t, we integrate
𝑀 𝑡
∫𝑀 𝑑𝑀𝐶𝑉 = − ∫0 |𝑚̇𝑒 |𝑑𝑡 where at t = 0, MCV = M0, and at t = t, MCV = M
0

Then, M-M 0 =−|𝑚̇𝑒 |𝑡, or M = M 0 - 𝑚̇𝑒 t.

{Since 𝑚̇𝑒 is positive, we have dropped the absolute value sign.}

Substituting the expression for M into term A , we obtain

𝐹𝐵𝑦 = − ∫ g𝜌𝑑∀ = −g𝑀CV = −g (M0 − ṁe 𝑡)


𝐶𝑉

B − ∫𝐶𝑉 𝑎𝑟𝑓𝑦 𝜌𝑑∀


The acceleration, arfy, of the CV is that seen by an observer in the XY coordinate system. Thus arfy is
not a function of the coordinates xyz, and

− ∫ 𝑎𝑟𝑓𝑦 𝜌𝑑∀ = − 𝑎𝑟𝑓𝑦 ∫ 𝜌𝑑∀ = − 𝑎𝑟𝑓𝑦 𝑀CV = = − 𝑎𝑟𝑓𝑦 (𝑀0− 𝑚̇𝑒 𝑡)


𝐶𝑉 𝐶𝑉

𝜕
C 𝜕𝑡
∫𝐶𝑉 𝜐𝑥𝑦𝑧 𝜌𝑑∀
Is the time rate of change of the y momentum of the fluid in the control volume measured relative
to the control volume.

Even though the y momentum of the fluid inside the CV, measured relative to the CV, is a large
number, it does not change appreciably with time. To see this, we must recognize that:

(1) The unburned fuel and the rocket structure have zero momentum relative to the
rocket.
(2) The velocity of the gas at the nozzle exit remains constant with time as does the
velocity at various points in the nozzle.

68
Consequently, it is reasonable to assume that

𝜕
∫ 𝜐 𝜌𝑑∀ ≈ 0
𝜕𝑡 𝐶𝑉 𝑥𝑦𝑧

D ∫𝐶𝑆 𝜐𝑥𝑧𝑦 𝜌𝑉̅𝑥𝑧𝑦 . 𝑑𝐴̅ = ∫𝐴𝑒 𝜐𝑥𝑧𝑦 |𝜌𝑉𝑥𝑧𝑦 𝑑𝐴| = 𝜐𝑥𝑧𝑦 |𝑚̇𝑒 |

Since 𝑉̅𝑒 = 𝑉𝑒 𝑗̂, then


𝜐𝑥𝑧𝑦 |𝑚̇𝑒 | = −𝑉𝑒 |𝑚̇𝑒 | = − 𝑉𝑒 𝑚̇𝑒

Substituting terms A through D into Eq. (4.8-1), we obtain

−g (M0 − ṁe 𝑡) − 𝑎𝑟𝑓𝑦 (M0 − ṁe 𝑡) = − 𝑉𝑒 𝑚̇𝑒


Or
𝑉𝑒 𝑚̇𝑒
𝑎𝑟𝑓𝑦 = − g (4.8-2)
M0 − ṁe 𝑡

At time t = 0,
𝑉𝑒 𝑚̇𝑒 m kg 1 m
𝑎𝑟𝑓𝑦 )𝑡=0 = − g = 3500 ×5 × − 9.81
M0 sec sec 400 kg s𝑒𝑐 2
𝑎𝑟𝑓𝑦 )𝑡=0 = 33.9 𝑚/s𝑒𝑐 2
𝑎𝑟𝑓𝑦 )𝑡=0

The acceleration of the CV is by definition


𝑑𝑉𝐶𝑉
𝑎𝑟𝑓𝑦 =
𝑑𝑡
Substituting from Eq. (4.8-2),
𝑑𝑉𝐶𝑉 𝑉𝑒 𝑚̇𝑒
= − g
𝑑𝑡 M0 − ṁe 𝑡
Separating variables and integrating gives
𝑉𝐶𝑉 𝑡 𝑡 M0 𝑚̇𝑒 𝑡
𝑉𝑒 𝑚̇𝑒 𝑑𝑡
𝑉𝐶𝑉 = ∫ 𝑑𝑉𝐶𝑉 = ∫ − ∫ g𝑑𝑡 = −𝑉𝑒 In [ 𝑒 ] − g𝑡
0 0 M0 − ṁe 𝑡 0 M0
At t = 10 sec,
m 350𝑘𝑔 m
𝑉𝐶𝑉 = − 3500 × ln [ ] − 9.81 10 sec
sec 350𝑘𝑔 sec2
𝑉𝐶𝑉 )𝑡 =10𝑠𝑒𝑐
𝑉𝐶𝑉 = 369 𝑚/ sec←

{This problem illustrates the application of the momentum equation to a linearly accelerating control
volume}

69
4.5 ENERGY EQUATION
Many problems involving fluid motion demand that the first law of thermodynamics, often
referred to as the energy equation, be used to relate quantities of interest. If the heat transferred
to a device, or the work done by a device, is desired, the energy equation is obviously needed. It
is also used to relate pressures and velocities when Bernoulli's equation is not applicable; this is
the case whenever viscous effects cannot be neglected. Let us express the energy equation in
control volume form. For a system itis
𝐷
Q̇ − 𝑊̇ = ∫ 𝑒𝜌𝑑∀ (4.5.1)
𝐷𝑡 𝑠𝑦𝑠

where the specific energy e includes specific kinetic energy V2/2, specific potential energy gz,
and specific internal energy ũ; that is,
𝑉2
𝑒= + g 𝑧 + 𝑢̃ (4.5.2)
2
We will not include other forms of energy, such as energy due to magnetic or electric field-flow
field interactions or those due to chemical reactions. In terms of a control volume, Eq. (4.5.1)
becomes
𝑑
Q̇ − 𝑊̇ = ∫ 𝑒𝜌𝑑∀
𝑑𝑡 𝐶.𝑉.
+ ∫𝐶.𝑆. 𝜌𝑒𝑉. 𝑛̂𝑑𝐴 (4.5.3)
This can be put in simplified forms for certain restricted flows, but first let us discuss the rate-
of-heat transfer term Q̇ and the work-rate term W.
The term Q̇ represents the rate-of-energy transfer across the control surface due to a temperature
difference. The rate-of-heat transfer term is either given or results from using Eq. (4.5.3). The
calculation of Q̇ is the objective of a course in heat transfer. The work-rate term is discussed in
detail in the following section.

4.5.1 Work-Rate Term

The work-rate term results from work being done by the system.

The work-rate term becomes

𝑊̇ = ∫𝑐.𝑠. 𝑝𝑛̂. 𝑉𝑑𝐴 + 𝑊̇𝑠 + 𝑊̇𝑠ℎ𝑒𝑎𝑟 + 𝑊̇𝐼 (4.5.4)

The terms are summarized as follows:

∫ 𝑝𝑛̂. 𝑉𝑑𝐴 work rate resulting from the force due to pressure moving at the control surface.
It is often referred to as flow work.
𝑊̇𝑠 work rate resulting from rotating shafts such as that of a pump or turbine, or the
equivalent electric power.
𝑊̇𝑠ℎ𝑒𝑎𝑟 work rate due to a moving boundary such as a moving belt.
𝑊̇𝐼 work rate that occurs when the control volume move relative to a fixed reference
frame.

70
4.5.2 General Energy Equation

𝑑 𝑃
Q̇ − 𝑊̇𝑠 − 𝑊̇𝑠ℎ𝑒𝑎𝑟 − 𝑊̇𝐼 = ∫ 𝑒𝜌𝑑∀ + ∫𝐶.𝑆. (𝑒 + 𝜌) 𝜌𝑛̂. 𝑉𝑑𝐴
𝑑𝑡 𝐶.𝑉.
(4.4.5)

𝑑 𝑉2 𝑉2
Q̇ − 𝑊̇𝑠 − 𝑊̇𝑠ℎ𝑒𝑎𝑟 − 𝑊̇𝐼 = ∫ ( 𝐼 + g𝑧 + 𝑢̃ )𝜌𝑑∀ + ∫𝐶.𝑆. ( 2𝐼 + g𝑧 + 𝑢̃ +
𝑑𝑡 𝐶.𝑉. 2
𝑃
) 𝜌𝑉. 𝑛̂𝑑𝐴 (4.4.6)
𝜌

This general form of the energy equation is useful in analyzing fluid flow problems that may
include time-dependent effects and non uniform profiles. Before we simplify the equation for
steady flow and uniform profiles let us introduce the notion of “Losses”.
We define losses as the sum of all the terms representing unusable forms of energy:
𝑑
Losses = −Q̇ + ∫ 𝑢̃𝜌𝑑𝑉
𝑑𝑡 𝐶.𝑉.
+ ∫𝐶.𝑆. 𝑢̃𝜌𝑉. 𝑛̂𝑑𝐴 (4.4.7)

We can now rewrite the energy equation as

𝑑 𝑉2 𝑉2
−𝑊̇𝑠 − 𝑊̇𝑠ℎ𝑒𝑎𝑟 − 𝑊̇𝐼 = ∫ ( 𝐼 + g𝑧)𝜌𝑑∀ + ∫𝐶.𝑆. ( 2𝐼 + g𝑧 +
𝑑𝑡 𝐶.𝑉. 2
𝑃
) 𝜌𝑉. 𝑛̂𝑑𝐴 + 𝑙𝑜𝑠𝑠𝑒𝑠 (4.4.8)
𝜌

Losses are due to two primary effects:


1. Viscosity causes internal friction that result in increased internal energy (temperature
increase) or heat transfer.
2. Changes in geometry result in separated flows that require useful energy to maintain the
resulting secondary motions that are generated.

In a conduit, the losses due to viscous Effects are distributed over the entire length, whereas
the loss due to a geometry change (a valve, an elbow, an enlargement) is concentrated in the
vicinity of the geometry change.

4.5.3 Steady, Uniform Flow


Consider a steady-flow situation in which there is one entrance and one exit across which
uniform profiles can be assumed. Also, assume that Ẇshear = ẆI = 0. For such a flow the
term (V2/2 + gz + p/) in Eq. (4.4.8) is constant across the cross section, because V is
constant (we assume a uniform velocity profile) and the sum of p/+ gz is constant if the
streamlines at each section are parallel. The energy equation (Eq. (4.4.8) then simplifies to
𝑉 2 𝑃2 𝑉12 𝑃1
−𝑊̇𝑠 = 𝜌2 𝑉2 𝐴2 ( 2 + + g𝑧2 ) − 𝜌1 𝑉1 𝐴1 ( + + g𝑧1 ) + Losses (4.4.9)
2 𝜌2 2 𝜌1

Where the subscripts 1 and 2 refer to the entrance and exit, respectively. The mass flux is
given by 𝑚̇ = 𝜌1 𝑉1 𝐴1 = 𝜌2 𝑉2 𝐴2 . After dividing by 𝑚̇g, we have

71
𝑊̇𝑠 𝑉22 −𝑉12 𝑃2 𝑃1
= + − + 𝑍2 − 𝑍1 + ℎ𝐿 (4.4.10)
𝑚̇g 2g 𝛾2 𝛾1

where we have introduced the head loss hL, defined to be

𝑄̇ ̃2 −𝑢
𝑢 ̃1
ℎ𝐿 = − + (4.4.11)
𝑚̇g g
It is often written in terms of a loss coefficient K as
𝑉2
ℎ𝐿 = 𝐾 (4.4.12)
2g
A final note to this section regards nomenclature for pumps and turbines in a flow system. It
is often conventional to call the energy term (𝑤̇𝑆 /𝑚̇g) associated with a pump “the pump
headHP”, and the term (𝑤̇𝑆 /𝑚̇g) associated with a turbine “the turbine headHT”. Then the
energy equation takes the form
V12 p1 V2 p
HP    z1  H T  2  2  z 2  hL (4.4.13)
2g  2g 

In this form we have equated the energy at the inlet plus added energy to the energy at the
exit plus extracted energy (energy per unit weight, of course). If any of the quantities is zero
(e.g., there is no pump), the appropriate term is simply omitted.
In some situations the turbine head or the pump head is known. 1 this is the case, the power
generated by the turbine with an efficiency of T simply

𝑊̇ 𝑇 = 𝑚̇g𝐻𝑇 η 𝑇 (4.4.14)

The power requirement by a pump with an efficiency of p would be


𝑚̇ g𝐻𝑝 𝑄𝛾𝐻𝑝
𝑊̇ = = (4.4.15)
𝜂𝑝 𝜂𝑝

We will calculate power in watts, ft-lb/sec, or horsepower. Recall that one horsepower is
equivalent to 746 W or 550 ft-lb/sec.

4.5.4 The Bernoulli Equation


The head loss is referred to as a "head" since it has dimensions of length. We could also refer
to V2/2g as the velocity head and p/ as the pressure head since those terms also have
dimensions of length. Also the term p/ + z is called the piezometric head. Further the sum
of the piezometric head and the velocity head is called the total head.
The energy equation, in the form of Eq. (4.4.10), is useful in man; applications and is,
perhaps, the most often used form of the energy equation. If the losses are negligible, if there
is no shaft work and if the flow is incompressible, we note that the energy equation takes the
form

72
V22 p V2 p
 2  Z 2  1  1  Z1 (4.4.16)
2g  2g 

Observe that the energy equation has been reduced to a form identical with Bernoulli's
equation. We must remember, however, that Bernoulli' equation is a momentum equation
applicable along a streamline and the equation above is an energy equation applied between
two sections of a flow. It is, not surprising that both should predict identical results from the
conditions stated because the velocity head is constant over a cross section and the sum of
pressure head and elevation remains constant over a cross section.

Figure 4.5 Pressure probes: (a) piezometer; (b) pitot probe; (c) pitot-static probe.

This is the well-known Bernoulli equation. Note the assumptions:


Inviscid flow (no shear stresses).
Steady flow (V/t = 0).
Along a streamline
Constant density (/s = 0).
Inertial reference frame
The pressure p is often referred to as static pressure and the sum of the two terms

V2
P  PT (4.4.17)
2

is called the total pressure PT or stagnation pressure.

73
EXAMPLE 4.9: Nozzle Flow: Application of Bernoulli Equation
Water flows steadily through a horizontal nozzle, discharging to the atmosphere. At the nozzle inlet
the diameter is D1; at the nozzle outlet the diameter is D2. Derive an expression for the minimum
gage pressure required at the nozzle inlet to produce a given volume flow rate, Q. Evaluate the inlet
gage pressure if D1 = 75.0 mm, D2 = 25.0 mm, and the desired flow rate is 0.02m3/sec.

SOLUTION OF EXAMPLE 4.9

GIVEN: Steady flow of water through a horizontal nozzle, discharging to the atmosphere.
D1=75 mm D2 = 25mm p2 = patm

FIND: (a) p1g as a function of volume flow rate, Q.


(b) Evaluate for Q = 0.02m3/sec.

SOLUTION:
Basic equations:
p1 V12 p V2
  gZ1  2  2  gZ 2
 2  2

  
0    d    Vxyz . dA
t cv cs

Assumptions: (1) Steady flow


(2) Incompressible flow
(3) Frictionless flow
(4) Flow along a streamline
(5) Z1= Z2
(6) Uniform flow at sections (1) and (2)
Apply the Bernoulli equation along a streamline between points (1) and (2). Then to evaluate p1.
Then
𝜌 2 2
𝜌 2 𝑉2 2
𝑝1g = 𝑝1 − 𝑝𝑎𝑡𝑚 = 𝑝1 − 𝑝2 = (𝑉2 − 𝑉1 ) = 𝑉1 [( ) − 1]
2 2 𝑉1

Apply the continuity equation

74
0   V1 A1  V2 A2  or V1 A1 V2 A2  Q

so that
V2 A1 Q
 and V1 
V1 A2 A1

Then

Q 2  A1  
2

p1g    1
2 A12 
 A2 
 
 

Since A = D2/4, then

8Q 2  D 
4

p1g  2 4 
D
1

 1
 D1 
 2  

With Dl =75 mm, D2 = 25 mm, and  = 999Kg/m3,

p1g 
8
2
 999
Kg

1
m3 (0.075) 4 m.4
 Q 2
(3.0) 4
1
Kg.m

N . sec 2

N . sec 2
p1g  2.05  109 Q 2 p1g
m8

With Q =0.02m3/sec, then p1g = 820Kpa.


{This problem illustrates the application of the Bernoulli equation to a flow where the restrictions of
steady, incompressible, frictionless flow along a streamline are a reasonable flow model.}

75
EXAMPLE 4.10
The wind reaches a speed of 100 km/h in a storm. Calculate the force acting on a 1 m  2 m window
facing the storm. The window is in a high-rise building, so the wind speed is not reduced due to
ground effects. Use  = 1.2 kg/m3.

SOLUTION OF EXAMPLE 4.10

The window facing the storm will be in a stagnation region where the wind speed is brought to zero.
Working with gage pressures, the pressure p upstream in the wind is zero. The velocity V must have
units of m/s. It is

km 1h 1000 m
V  100    27.8 m / s
h 3600 s 1 km

Bernoulli's equation then allows us to calculate the pressure on the window us follows:

V22 p V2 p
 2  z 2  1  0 1  z1
0
2g  2g 

V12
 p2 
2
1.2  (27.8) 2
  464 N m 2
2
where we have used z2 = z1, p1= 0, V2 = 0, and = g. Multiply by the area and find the force to be
F = pA
= 464  1  2 = 928 N

76
EXAMPLE 4.11
The static pressure head in an air pipe is measured with a piezometer as 16 mm of water. A pitot
probe at the same location indicates 24 mm of water. Calculate the velocity of the 20°C air. Also,
calculate the Mach number and comment as to the compressibility of the flow.

SOLUTION OF EXAMPLE 4.11

Bernoulli's equation is applied between two points on the streamline that terminates at the stagnation
point of the pitot probe. Point 1 is upstream and p2 is the total pressure; then
V12 p p
 1  T
2g  
We use the ideal gas law to calculate the density:
p

RT
9810  0.016  101000
  1.20 kg m 3
287  ( 273  20)
where the pressure has been found using p = h and atmospheric pressure, which is 101 000 Pa, is
added since absolute pressure is needed in the preceding equation. The velocity is then
2
V1  ( pT  p1 )

2 (0.024  0.016)  9810


  11.44 m s
1.20
To find the Mach number, we must calculate the speed of sound. It is

C  kRT

 1.4  287  293  343 m s

The Mach number is then


V 11.44
M   0.0334
C 343
Obviously, the How can be assumed to be incompressible since M < 0.3. Thevelocity would have to
be much higher before compressibility would be important.

77
EXAMPLE 4.12
A venturi meter reduces the pipe diameter from 10 cm to 5 cm. Calculate the flow rate and the mass
flux assuming ideal conditions.

SOLUTION OF EXAMPLE 4.12

The control volume is selected such that the entrance and exit correspond to the sections where the
pressure information of the manometer can be applied. The manometer's reading is interpreted as
follows:
p1 +  (z + 1.2) = p2 + z + 13.6 1.2

where z is the distance from the pipe centerline to the top of the mercury column. The manometer
then gives
p1  p 2
 15.12m

Continuity allows us to relate V2 to V1 as
V1A1 = V2A2
 V2 = 4V1
The energy equation with no losses takes the form
0
V 2  V12 P  P1
0 2  2  z2  z1 
2g 

16V12 V12
 15.12
2g

 V1 = 4.45 m/s
78
The flow rate is
Q = V1A1,
   0.052  4.45  0.0350 m3 s

The mass flux is


𝑚̇ = 𝜌 𝒬
= 1000 x 0.035 = 35.0 kg/s

79
EXAMPLE 4.13
Water flows from a reservoir through a 0.8-m-diameter pipeline to a turbine-generator unit and exits
to a river that is 30 m below the reservoir surface. If the flow rate is 3m3/s and the turbine-generator
efficiency is 80%, calculate the power output. Assume the loss coefficient in the pipeline (including
the exit) to be K = 2.

SOLUTION OF EXAMPLE 4.13

The control volume to be used extends from section 1 to section 2; we consider the water surface of
the left reservoir to be the entrance and the water surface of the river to be the exit. Because we
assume the water surfaces to be large, the velocities at the surfaces are negligible. The velocity in the
pipe is
Q
V
A
3
  5.968 m s
  0. 8 2 4

Now, consider the energy equation. We will use gage pressures so that p1 = p2 = 0 the datum is
placed through the lower section 2 so that z2 = 0; the velocities V1 and V2 are negligibly small; K is
assumed to be based on the 0.8-m-diameter pipe velocity. The energy equation (4.4.13) then
becomes

0 0 0 0 0
0
2
V p V2 p V2
HP   1 z1  HT  2  2  z2  K
1
2g  2g  2g
5.968 2
30  H T  2
2  9.81

 HT  26.4m

From this the power output is found, using Eq. (4.4.14), to be



WT  QH TT

= 3 x 9810 x 26.4 x 0.8 = 622000 W or 622 kW

80
In this example we have used gage pressure; the potential-energy datum was assumed to be placed
through section 2, V1 and V2 were assumed to be insignificantly small, and K was assumed to be
based on the 0.8 m-diameter pipe velocity.

81
EXAMPLE 4.14 —Compressor: First Law Analysis

Air at 101 KPa (abs), 21C, enters a compressor with negligible velocity and is discharged at 350
KPa (abs), 38 C through a pipe with 0.9 m2 area. The flow rate is 10 Kg/sec. The power input to the
compressor is 450 kW. Determine the rate of heat transfer.
GIVEN: Air enters a compressor at  and leaves at  with conditions as shown. The air flow rate
is 10 Kg/sec and the power input to the compressor is 450 kW.

FIND: The rate of heat transfer.

SOLUTION OF EXAMPLE 4.14

Basic equations:

 0(1)

t cv 
0  d   V .d A
cs

 0(4)  0(1)
. . .   V2 
t cv cs 
Q  W s  W shear  e  d   u  p   gz  V . d A
2 

Assumptions: (1) Steady flow


(2) Properties uniform over inlet and outlet sections
(3) Treat air as an ideal gas, p = RT
(4) Area of CV at (1) and (2) perpendicular to velocity, thus 𝑤̇𝑠ℎ𝑒𝑎𝑟 = 0
(5) z1=z2
(6) Inlet kinetic energy is negligible
Under the assumptions listed, the first law becomes
. .  V2 
Q  Ws   
 u  p   V . d A
 gz 
cs
 2 

 V2 
h  u  p
. .
Q  Ws    h   gz  V . d A where
cs
 2 

or
. .  V2 
Q  Ws    h   gz  V . d A
cs
 2 

For uniform properties (assumption 2) we can write


82
 0(6)
   
Q Ws   h1  1  gz1   1V1 A1   h2  2  gz 2   2V2 A2 
. . V2 V2
 2   2 

For steady flow, from conservation of mass,


cs
V . d A  0

Therefore, − |𝜌1 𝑉1 𝐴1 | + |𝜌2 𝑉2 𝐴2 | = 0,or |𝜌1 𝑉1 𝐴1 | = |𝜌2 𝑉2 𝐴2 | = 𝑚̇. Hence we can write

 0(5)
. . .  V2 
Q Ws  m (h2  h1 )  2  g ( z 2  z1 )
 2 

Assume that air behaves as an ideal gas with constant cp. Then h2 – h1 =cp (T2 — T1), and
. . .  V2
Q Ws  m c p (T2  T1 )  2 
 2 

From continuity, |𝑉2 | = 𝑚̇/𝜌2 𝐴2 . Since p2 = 𝜌2 RT2, then


.
m RT2 Kg 1 J m2
V2   10   287  311K 
A2 p2 sec 0.09m 2 Kg.K 350 kN

|V2| = 28.3m/sec
V22
Q Ws  m c p T2  T1   m
. . . .

Note that power input is to the CV, so W, =-450 kW, and


. Kg J
Q   450kW  20 100 0  17 K
sec Kg.K

Kg 28.3 m 2
2
J . sec 2
 10  
sec 2 sec 2 Kg.m 2

heat rejectionQ
. .
Q   276 kJ sec

{In addition to demonstrating a straightforward application of the first law, this problem illustrates
the need for keeping units straight.}

83
EXAMPLE 4.15: Tank Filling: First Law Analysis
A tank of 0.1 m3 volume is connected to a high-pressure air line; both line and tank are initially at a
uniform temperature of 20 C. The initial tank gage pressure is 100 kPa. The absolute line pressure is
2.0 MPa; the line is large enough so that its temperature and pressure may be assumed constant. The
tank temperature is monitored by a fast-response thermocouple. At the instant after the valve is
opened, the tank temperature rises at the rate of 0.05 C/sec. Determine the instantaneous flow rate
of air into the tank if heat transfer is neglected.

SOLUTION OF EXAMPLE 4.15

GIVEN: Air supply pipe and tank as shown. At t = 0+, T/t = 0.05 C/sec.
FIND: 𝑚̇at t = 0+
Choose CV shown, apply energy equation.
 0(1)  0(2)  0(3)  0(4)

e d   e  p V . d A
. . . .

t cv
Q  Ws  W shear  W other 
cs

 0(5)  0(6)
V2
eu  gz
2
Assumptions: (1) Q̇ =0 (given)
(2) 𝑊̇𝑆 = 0
(3) 𝑊̇𝑠ℎ𝑒𝑎𝑟 = 0
(4) 𝑊̇𝑂𝑡ℎ𝑒𝑟 = 0
(5) Velocities in line and tank are small
(6) Neglect potential energy
(7) Uniform flow at tank inlet
(8) Properties uniform in tank
(9) Ideal gas, p = RT, du =cdT

84
Then


0  utan k  d  uline  p  VA 
t cv

But initially.T is uniform, so utank = uline = u and



0  u d  u  p  VA 
t cv

Since tank properties are uniform, /t may be replaced by d/dt, and
d
uM   u  p  m
.
0
dt

or
dM du . .
0u M  u m  p m
dt dt (4.15-1)

The term dM/dt may be evaluated from continuity:



Basic equation: 0
t cv
 d   V . d A
cs

  VA 
dM dm .
0 or m
dt dt

Substituting into Eq. (4.15-1) gives


. du . . dT .
0u m  M  u m  p m  Mc  p m
dt dt

Or
. Mc dT dt  c dT dt  c dT dt 
m  
p p RT (4.15-2)

But at t = 0, ptank = 100 kPa (gage), and

 1.00 1.01105 2 
ptan k N kg.K 1
   tan k    2.39 kg m3
RT m 287 N .m 293 K

Substituting into Eq. (4.15-2), we obtain


. kg N .m K kg.K 1 g
m  2.39 3
 0.1 m 3  717  0.05   1000
m kg.K sec 287 N .m 293 K kg

. .
m  0.102 g sec m

{This problem illustrates the application of the energy equation to an unsteady flow situation.}

85
EXAMPLE 4.16
Consider the symmetrical flow of air around the cylinder. The control volume, excluding the
cylinder, is shown. The velocity distribution downstream of the cylinder is measured to be as shown.
Determine the drag force per meter of length acting on the cylinder. Use  = 1.23 kg/m3.

SOLUTION OF EXAMPLE 4.16


First, we must recognize that not all of the mass flux entering through AB exits through CD;
consequently, some mass flux must exit AD and BC, as shown. The momentum equation for the
steady flow, applied to the control volume ABCD. takes the form
F   VxV . nˆ dA
c. s .

. .
 u 2 dA  u m AD  u m BC   u 2 dA
ACD AAB

2
10  y2  .
2 1.23  29   dy  2  30 m AD 1.23 30 2  20
0
 100 

where𝑚̇𝐵𝐶 = 𝑚̇𝐴𝐷 is the mass flux crossing BC and AD with the x-component velocity equal to
30m/s. Use continuity to find 𝑚̇𝐴𝐷 :
. 10
0  m AD   u ( y ) dy   10  30
0

. 10  y2 
 m AD  
1.23  29   dy 1.2310  30
0
 100 
.
 m AD  8.2 kg s per meter

Evaluating the terms in the momentum equation above gives us


F = – 21170 – 492 + 22140
= 480 N per meter

86
EXAMPLE 4.17
Find an expression for the head loss in a sudden expansion in a pipe in terms of V1 and the area
ratio. Assume uniform velocity profiles and assume that the pressure at the sudden enlargement is p1.

SOLUTION OF EXAMPLE 4.17


A sketch is shown of a sudden expansion, with the diameter changing from d1 to d2. The pressure at
the sudden enlargement is p1, so that the force acting on the left end shown is p1A2. Newton's second
law applied to the control volume yields, assuming uniform profiles,
.
F x  m (V2 V1 )

( p1  p2 ) A2  A2V2 (V2 V1 )

p1  p2
  V2 (V2  V1 )

The energy equation provides


V22 V12 p2  p1
0   z2  z1  hL
2g 
With z1 = z2 , then

p1  p2 V22 V12
 hL  
 2g

2V2 (V2 V1 ) (V2  V1 ) (V2 V1 ) (V1 V2 ) 2


hL   
2g 2g 2g

To express this in terms of only V1, we can use continuity and relate
A1
V2  V1
A2

Then expression above for the head loss becomes


2
 A  V2
hL  1  1  1
 A2  2 g

87
TABLE 4.1 INTEGRAL FORMS OF THE FUNDAMENTAL LAWS
Continuity Energy Momentum
General Form
𝑑 𝑉2 𝑑
𝑑 − ∑ 𝑤̇ = ∫ ( + g𝑧) 𝜌𝑑∀ ∑𝐹 = ∫ 𝜌𝑉𝑑∀ + ∫ 𝜌𝑉(𝑉. 𝑛̂ )𝑑𝐴
0= ∫ 𝜌𝑑∀ + ∫ 𝜌𝑉. 𝑛̂𝑑𝐴 𝑑𝑡 𝑐.𝑣 2 𝑑𝑡 𝑐.𝑣 𝑐.𝑠.
𝑑𝑡 𝑐.𝑣 𝑐.𝑠.
𝑉2 𝑃
+ ∫ ( + + g𝑧) 𝜌𝑉. 𝑛̂𝑑𝐴 + 𝑙𝑜𝑠𝑠𝑒𝑠
𝑐.𝑠. 2 𝜌

Steady flow

𝑉2 𝑃
− ∑ 𝑤̇ = ∫ ( + + g𝑧) 𝜌𝑉. 𝑛̂𝑑𝐴 + 𝑙𝑜𝑠𝑠𝑒𝑠 ∑ 𝐹 = ∫ 𝜌𝑉(𝑉. 𝑛̂ )𝑑𝐴
0 = ∫ 𝜌𝑉. 𝑛̂𝑑𝐴 𝑐.𝑠. 2 𝜌 𝑐.𝑠.
𝑐.𝑠.

Steady, Non Uniform

− ∑ 𝑊̇ 𝑉̅22 𝑃2 𝑉̅12 𝑃1 ∑ 𝐹𝑥 = 𝑚̇(𝑉̅2𝑥 − 𝑉̅1𝑥 )


𝑚̇ = 𝜌1 𝐴1 𝑉̅1 = 𝜌2 𝐴2 𝑉̅2 = 𝛼2 + + 𝑍2 − 𝛼1 − − 𝑍1 + ℎ𝐿
𝑚̇g 2g 𝛾2 2g 𝛾1 ∑ 𝐹𝑦 = 𝑚̇(𝑉̅2𝑦 − 𝑉̅1𝑦 )
Steady, Uniform flow

− ∑ 𝑊̇ 𝑉22 𝑃2 𝑉12 𝑃1
𝑚̇ = 𝜌1 𝐴1 𝑉1 = 𝜌2 𝐴2 𝑉2 = + + 𝑍2 − − − 𝑍1 + ℎ𝐿 ∑ 𝐹 = 𝑚̇(𝑉2 − 𝑉1 )
𝑚̇g 2g 𝛾2 2g 𝛾1

Steady, Uniform, Incompressible flow

𝑄 = 𝐴1 𝑉1 = 𝐴2 𝑉2 − ∑ 𝑊̇ 𝑉22 𝑃2 𝑉12 𝑃1
= + + 𝑍2 − − − 𝑍1 + ℎ𝐿 ∑ 𝐹 = 𝑚̇(𝑉2 − 𝑉1 )
𝑚̇g 2g 𝛾2 2g 𝛾1
Or
𝑉12 𝑃1 𝑉22 𝑃2
𝐻𝑝 + + + 𝑍1 = 𝐻𝑇 + + + 𝑍2 + ℎ𝐿
2g 𝛾1 2g 𝛾2

88
5. REFERENCES

1)Introduction to Fluid Mechanics


By R. W. Fox
Alan T. Me Donald
John Wiley & Sons. Fourth, 1994

2)Fundamentals of Fluid Mechanic


By Bruce R. Munson, Donald F. Young
and Theodore H. Okiishi
John Wiley & Sons 1994

3)Fluid Mechanics
By Frank M. White
McGraw Hill, Inc., Second Edition, 1986

4)Mechanics of Fluids
By M. C. Potter
D. C. Wiggert
Prentice Hall, Inc., 1991

89

You might also like