0% found this document useful (0 votes)
751 views138 pages

PRESSURE AND (Repaired)

Uploaded by

Nitish
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
751 views138 pages

PRESSURE AND (Repaired)

Uploaded by

Nitish
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 138

PRESSURE AND

FLUID STATICS
3–1 PRESSURE
!

Pressure is defined as a normal force exerted by a fluid per unit area.


Since pressure is defined as force perunit area, it has the unit of newtons per square meter (N/m2),
which is called a pascal (Pa)

The actual pressure at a given position is called the absolute pressure,


and it is measured relative to absolute vacuum (i.e., absolute zero pressure).
Most pressure-measuring devices, are calibrated to read zero in the
atmosphere, and so they indicate the difference between the
absolute pressure and the local atmospheric pressure. This difference is
called the gage pressure. Pressures below atmospheric pressure are called
vacuum pressures and are measured by vacuum gages that indicate the difference
between the atmospheric pressure and the absolute pressure.

Pressure at a Point
Pressure is the compressive force per unit area, and it gives the impression
of being a vector. However, pressure at any point in a fluid is the same in all
directions. That is, it has magnitude but not a specific direction, and thus it
is a scalar quantity.
Variation of Pressure with Depth
Pressure in a fluid increases with depth because more fluid rests on deeper layers, and the effect of
this “extra weight” on a deeper layer is balanced
by an increase in pressure

If we take point 1 to be at the free surface of a liquid open to the atmosphere


(Fig. 3–8), where the pressure is the atmospheric pressure Patm, then
the pressure at a depth h from the free surface becomes

For fluids whose density changes significantly with elevation, a relation


for the variation of pressure with elevation can be obtained by dividing Eq.
3–6 by $x $z, and taking the limit as $z 0. It gives

The negative sign is due to our taking the positive z direction to be upward
so that dP is negative when dz is positive since pressure decreases in an
upward direction. When the variation of density with elevation is known, the pressure difference
between points 1 and 2 can be determined by integration
to be

Pressure in a fluid at rest is independent of the shape or cross section of


the container. It changes with the vertical distance, but remains constant in
other directions.
A consequence of the pressure in a fluid remaining constant in the horizontal
direction is that the pressure applied to a confined fluid increases the
pressure throughout by the same amount. This is called Pascal’s law.

“Pascal’s
machine” has been the source of many inventions that are a part of our
daily lives such as hydraulic brakes and lifts

Consider the manometer shown in Fig. 3–11 that is used to measure the
pressure in the tank. Since the gravitational effects of gases are negligible,
the pressure anywhere in the tank and at position 1 has the same value,
since pressure in a fluid does not vary in the horizontal direction
within a fluid, the pressure at point 2 is the same as the pressure at point 1,
P2 " P1.
where r is the density of the fluid in the tube. Note that the cross-sectional
area of the tube has no effect on the differential height h, and thus the pressure
exerted by the fluid. However, the diameter of the tube should be large
enough (more than a few millimeters) to ensure that the surface tension
effect and thus the capillary rise is negligible.
A wide variety of pressure transducers is available to measure gage,
absolute, and differential pressures in a wide range of applications. Gage
pressure transducers use the atmospheric pressure as a reference by venting
the back side of the pressure-sensing diaphragm to the atmosphere, and they
give a zero signal output at atmospheric pressure regardless of altitude. The
absolute pressure transducers are calibrated to have a zero signal output at
full vacuum. Differential pressure transducers measure the pressure difference

between two locations directly instead of using two pressure transducers


and taking their difference.
Strain-gage pressure transducers work by having a diaphragm deflect
between two chambers open to the pressure inputs. As the diaphragm
stretches in response to a change in pressure difference across it, the strain
gage stretches and a Wheatstone bridge circuit amplifies the output. A
capacitance transducer works similarly, but capacitance change is measured
instead of resistance change as the diaphragm stretches.
Piezoelectric transducers, also called solid-state pressure transducers,
work on the principle that an electric potential is generated in a crystalline
substance when it is subjected to mechanical pressure.

THE BAROMETER AND


ATMOSPHERIC PRESSURE
Atmospheric pressure is measured by a device called a barometer; thus, the
atmospheric pressure is often referred to as the barometric pressure.

the atmospheric pressure can be measured by inverting a


mercury-filled tube into a mercury container that is open to the atmosphere,
as shown in Fig. 3–17. The pressure at point B is equal to the atmospheric
pressure, and the pressure at C can be taken to be zero since there is only
mercury vapor above point C and the pressure is very low relative to Patm
and can be neglected to an excellent approximation.

where r is the density of mercury, g is the local gravitational acceleration,


and h is the height of the mercury column above the free surface. Note that
the length and the cross-sectional area of the tube have no effect on the
height of the fluid column of a barometer

The standard atmospheric pressure Patm changes from 101.325 kPa at sea
level to 89.88, 79.50, 54.05, 26.5, and 5.53 kPa at altitudes of 1000, 2000,
5000, 10,000, and 20,000 meters, respectively.

Skip to content

 Safari Home

 Resource Centers

 Playlists

 

 History

 Topics

 Learning Paths

 Offers & Deals

 Newsletters


 Highlights


o Settings
o Support
o Sign Out
o
o
o

Table of Contents for


Fluid Mechanics and Hydraulic Machines

 Search in book...

 Toggle Font Controls



o
o
o
o

Prev Previous Chapter


Symbols Used

Next Next Chapter


2. Mathematical Modelling of Flow

CHAPTER 1
Fluid Statics
OUTLINE

 Introduction
 Continuum Concept
 Pressure Variation and Hydrostablic Forces on Surfaces Submerged in Fluid at Rest
 Metacentric Height
 Pressure Variation in Liquids Confined in Moving Container

1.1 Introduction

Every substance in nature is made of matter, which may be organic or inorganic compounds and
may exist in different phases, viz. solid, liquid or gaseous or in two or even more phases in
equilibrium with each other. These materials, both organic or inorganic, are found in the natural
systems like the human or animal body, stars, rivers and so on that we see around us and are also
used to develop engineering systems. In this section we shall have a look at it.

1.1.1 About the Substance

Substance is the name used for the elements and their compounds. For example, carbon is an
element and water is a compound in which the hydrogen and oxygen are combined in the ratio
2:1. There are 112 elements and infinite number of compounds formed from the chemical
bondage of different elements. A substance may exist in a single phase, such as solid, liquid or
gas, or two phases in equilibrium with each other or all the three phases together that will depend
on pressure and temperature of the substance. We may even have mixture of substances that may
be homogeneous or heterogeneous. Anyway, the scope of this book is limited to pure substances
only and therefore we shall skip discussions about the mixture.

A substance exists in solid, liquid or gaseous phase depending on the gravitational force between
the molecules of the substance of which the substance is made of. We learn in kinetic theory that
a substance can be thought of made up of molecules that move randomly in all directions with all
possible velocities. That is, their motion is chaotic. This behaviour is confirmed by many
experimental results; Brownian movement in colloidal solutions is normally cited as an
experimentally observed phenomenon in its support. These molecules though extremely small,
visible only by a high magnifying power microscope, have mass and hence the molecules attract
each other as per the law of gravitation. The force of attraction varies as inverse of square of the
distance between the molecules. Thus, in solids the distance between the molecules is so small
that the gravitational force does not allow them to escape and the whole substance behaves as
solid. The solids therefore are difficult to be compressed by application of force and a large force
is needed to do so. When solid is heated, at some temperature, the distance between the
molecules increases because of the increase in internal energy and a solid starts behaving
differently. First it comes in semi-solid state and finally in the liquid state.

In liquids the distance between the molecules is more than that in solids but not so large that the
gravitational force is zero. Because of this, the substance can flow and acquire the shape of the
container in which it is stored. However, the molecules in the free liquid surface have enough
velocity to leave the free surface and therefore evaporation from the surface goes on though the
amount of molecules leaving the surface is small and depends on the temperature of the liquid.
At some temperature, called boiling point, more and more molecules acquire velocity high
enough to overcome gravitational force between the molecules and thus leave the surface. Thus
the liquid changes to gaseous phase. We may also explain the surface tension and capillary
effects in similar way. Liquids too, like solids, are difficult to be compressed and their density
can be assumed constant. These are often referred as incompressible fluids.
On the contrary, the distance between the molecules in gases is large enough for existence of any
gravitational force between them and the molecules fly in all directions. The volume occupied by
the gas is the volume of the container in which it is stored. Gases are compressible and their
density changes with the temperature and pressure. This compressibility introduces many
complications in the solution of differential equations to be studied later.

Continuum Concept

Two methods for analysis of system have been used.

In one approach, the system is considered at its molecular level as in kinetic theory of gases.
That is, the system is assumed to be made up of molecules in incessant random motion with wide
range of velocities. This kind of analytic approach is called microscopic approach and has been
useful to predict pressure, temperature and internal energy of gases contained in a container in
terms of the molecules density (number of molecules per unit volume) and their velocity
distribution, expressed by the average, most probable, and root mean square velocity. These
relations were obtained using the statistical tools. An important result is obtained for the mean
free path, which is the average of the free path defined as the distance travelled by a molecule
between its successive collision. This distance will vary and would depend on the proximity of
the other molecules around it. The mean free path λ is given by the relation, λ = 1/σn, where σ is
the collision cross-sectional area = πd2, d is the diameter of the molecule and n is the number of
molecules in a unit volume.

However, the microscopic analysis is very time consuming and in most of the applications it is
not required to go into details of the individual molecule and then taking out the average. When
the size of the system is much larger than the mean free path of the molecules, then what
happens to the individual molecule is of no consequence and the main thrust lies in the
determination of the interaction between the system and its surroundings and also the overall
effect left on the system (as well as the surroundings). Such approach is called the macroscopic
approach. Under such circumstances the matter may be considered as continuous. This is called
continuum concept, that is, the matter is uniformly and continuously distributed and not made up
of molecules with space in between. Thus, we shall follow macroscopic approach in this book.

Fig. 1.2.1 Variation of density with volume


We would like to study behaviour of a system as its size or volume is reduced. Let δV be the
volume occupied by some mass δm of the system at any time. Then their ratio, δm/δV, we know
is called the density, ρ. Figure. 1.2.1 shows variation of density ρ as volume increases. It is
observed that when volume is δV′ or less, the density changes sharply but the density approaches
asymptotically to a value as volume of the system increases; the limiting value is used in day-to-
day life as density. This happens because the mater can no longer be considered as continuous
when volume of the system has reached δV′ and the presence of molecules affect the result. This
means that we may consider the matter continuous when volume is more than a certain volume,
which is related to the mean free path. That is why the concept of macroscopic approach is
applicable only when the smallest size of system is more than the mean free path. This needs to
redefine the definition of pressure, density etc. as the value obtained by taking the limit when
area or volume of the element tends to zero. If δA′ is the smallest area up to which continuum
concept holds true then pressure is defined as

Where F is the force acting on surface area δA and δm is mass of the volume element δV. Note
the difference lies in taking the limit. Here it is not zero as then the continuum concept is not
applicable. Density is measured in kg/m3 in SI units whereas pressure in Pascal Pa, which is
N/m2. However, MPa, kPa, or bar is commonly used unit for pressure, where 1 Mpa ≡ 106 Pa, 1
kPa ≡ 103 Pa, and 1 bar ≡ 105 Pa. Specific gravity and specific weight are the other ways of
expressing density of any substance. Specific gravity is the ratio of the density of the substance
with that of water, which is taken as 1000 kg/m3 and specific weight is the product of density and
gravity, which is usually represented by the symbol w (or γ). Specific weight of water is 9810
N/m3.1 For air the density of air at atmospheric pressure at 20°C is 1.24 kg/m3.

Pressure can be given as absolute pressure or how much it is above or below the atmospheric
pressure. The positive pressure is called the gauge pressure and the negative pressure the
vacuum. The atmospheric pressure acts on our whole body but the net force on the body is zero.
This atmospheric pressure is about 76 cm of mercury column, which is too large compared to
blood pressure of 12 cm of mercury in our body.

The other properties of fluids are dynamic viscosity and thermal conductivity. We learn from
Kinetic theory of substances that these properties are related to the movement of molecules from
one layer to another in laminar flow case in which there exists some gradient of velocity and
temperature across the layers. Molecules from lower velocity layer move with less momentum as
compared to those from the higher velocity layer; the effect is dragging of fast moving layer and
accelerating of slow speed moving layer. This phenomenon is expressed by the property of the
fluid called dynamic viscosity, which is given by the relation μ = 1/3 mn λ, where is average
velocity of the molecules. On macroscopic basis, dynamic viscosity is related to velocity
gradient and is given by the Stoke’s relation, in a three dimensional flow,

τ ∝ (∂u/∂y) + (∂v/∂x)
or,

where μ is coefficient of dynamic viscosity, τ is the shear stress on the surface along which flow
is taking place and u and v are the component of velocities along x and y direction of the
reference axes assumed to be fixed to the ground. Noting that dimensions for τ is M L–1T–2, for
velocity LT–1, and for x or y that of L, the dimension of dynamic viscosity is obtained from the
principle of dimensional homogeneity to give

[μ] = [τ]/[∂u/∂y]

But [∂u/∂y] = LT–1/L = T–1

Fig. 1.2.2 Variation of shear stress with velocity gradient

We have used notation [τ] in Eq. 1.2.3 to represent the dimensions of the quantity τ. Unit of
dynamic viscosity in SI system of units is kg/m-s.2 However, its unit were defined in CGS
system in which it is g/cm-s, called poise, which is 10–1 kg/m-s. We may realise that its value
will depend on temperature, since the molecular movement is related to temperature and the kind
of the fluid. Figure. 1.2.2 shows variation of shear stress at any point in the fluid with velocity
gradient for different kinds of the fluids. It is seen that for some fluids, this variation is linear.
Such fluids are called Newtonian fluids. For other fluids, the variation is non-linear which falls
under the category of non-Newtonian fluids. Therefore these also include the ones for whom the
curves do not start from the origin. This behaviour is observed when the substance is in the
plastic stage. Many mathematical models are available to describe the behaviour of the non-
linear shear stress with deformation rate (called the constitutive relations) for Newtonian fluids.
No such relation, however, can be prescribed for all kinds of non-Newtonian fluids. The power
law model, known as the Ostwald – de Waele model, named after the scientist who proposed it,
is used for the most common non-Newtonian fluids. According to this model

where m and n are the flow consistency and flow behaviour index, respectively. We observe that
the viscosity in such fluids is function of the deformation rate as well. When n = 1, the model is
same as for Newtonian fluid. Fluids such as gelatine, blood, milk, etc have n < 1 and are called
pseudoplastic fluids. Whereas the dilatant fluids such as sugar in water, aqueous suspension of
rice starch, etc have n > 1. Table 1.2.1 gives classification according to the relation between
shear stress and the rate of angular deformation:

TABLE 1.2.1 Classification of the fluids

Shear stress Kind of fluid


τ=o Ideal fluid
τ = μ (∂u/∂y) Newtonial fluid
τ = constat + μ (∂u/∂y) Ideal plastics or Bingham plastics
τ = constat + μ (∂u/∂y)n Thyxotropic fluids
τ = μ (∂u/∂y)n Non-Newtonial fluids

We shall consider only the Newtonian fluids for our discussion in this book. While dealing with
the equations, a group (μ/ρ) often appears and therefore this group of properties of the substance
has been given a separate name, the kinematic viscosity having unit m2/s in SI units, which is
named Stoke, and represented by the symbol ν. We may note that kinematic viscosity has the
dimensions L2T–1. Both the dynamic and kinematic viscosities are function of pressure and
temperature. But the variation of the viscosity with pressure for liquids is so small that it can be
neglected and dynamic viscosity at any temperature T can be obtained from the relation μT =
AeB/T, where A and B are the constants. On the other hand, the relation μT = AT1/2/(1 + B/T), in
which A and B are constants, gives variation of the viscosity of gases with temperature. We
notice that while viscosity of the liquids decreases with temperature, it increases for the gases.
Figure. 1.2.3 shows variation of dynamic viscosity for some liquids and gases with temperature.
Fig. 1.2.3 Viscosity of common fluids as a function of temperature

The following empirical relations have also been suggested for variation of viscosity with
temperature and pressure:

For liquids
and

μp = μ0 exp[K(p – p0)] = μ0 exp (Kpg)

where A, B, and μ0 are constants in the expression for the dynamic viscosity μt at temperature t
°C and depend on the type of the fluid. Here pg represents the gauge pressure measured above the
atmospheric pressure. For water μ0 = 0.0179 poise, A = 0.03368 and B = 0.000221. This relation
shows inverse parabolic variation of the viscosity with temperature that tends to zero at infinite
temperature. The other correlation in the Eq. shows variation of viscosity with pressure in which
the constant K depends on the nature of the fluid; p0 is the atmospheric pressure. That is we use
the gauge pressure in the expression. The increase in viscosity with pressure was expected
because the amount of energy required for the relative movement of the molecules is more. The
viscosity increases by 10 to 15 percent for a pressure increase of about 75 atmospheres and is
doubled for increase from 1 to 1000 atmospheres.

For gases

μt = μ0 [1 + At – Bt2]

For air, μ0 = 1.7 × 10–5 Ns/m2, A = 0.56 × 10–7, and B = 0.1189 × 10–9.

We note that dynamic viscosity is the result of velocity gradient in the flow region. If the fluid is
at rest or moves with uniform velocity then there exists no such gradient and there would not be
shear stresses on the face of the element considered in the fluid. However, the pressure that acts
normal to the surface would alone be acting. Without going into the proof, we make a statement
that the fluids can take only compressive forces and not the tensile forces, unlike a solid. That is
pressure is always negative, as per the convention adopted for the solids.

Example 1.2.1 Determine the kinematic viscosity of a fluid having dynamic viscosity 1.85 × 10–
4
poise and density 1.208 kg/m3.

Solution

Given μ = 1.85 × 10–4 poise


= 1.85 × 10–4 × 10–1
= 1.85 × 10–5 kg/m-s,
and ρ = 1.208 kg/m3.
∴ ν = μ/ρ = 1.531 × 10–5 m2/s.

Example 1.2.2 Determine the density of H2 at a temperature of 40°C and pressure of 2 atm.

Solution 1 atm = 1.01325 × 105 N/m2


→ p = 2 atm = 2.0265 × 105 N/m2. Also T = 313 K

For H2, R = 8.3143/2 kJ/kg K = 4157.15 J/kg.

Eq. of state gives, ρ = p/RT

= (2.0265 × 105)/(4157.15 × 313)


= 1.54 kg/m3
specific weight (W) = ρg
= 1.54 × 9.81
= 15.1074 N/m3.

Example 1.2.3 If the velocity distribution in a flow over a surface is given by the relation u = 2y
– 2y3 + y4, determine the shear stress on the surface at y = 0 and y = 0.1 m if dynamic viscosity of
the fluid is 1 × 10–3 kg/m-s.

Solution Differentiating the given velocity profile with reference to y, we get

du/dy = 2 – 6y2 + 4y3


∴ At y = 0, t = m(du/dy)y = 0
= 1 × 10–3 × 2
= 2 × 10–3 N/m2.
And at y = 0.1 m, t = m(du/dy)y = 0.1
= 1 × 10–3 × (2 – 0.06 + 0.0004)
= 1.994 × 10–3 N/m2.

Example 1.2.4 A plate of area 5.0 m2 is forming interface between two liquids of different
viscosities, as shown in Fig. E 1.2.4. The plate is moved with a force 150 N applied at its one
end. Determine the velocity with which it moves if the dynamic viscosity of the fluid above it is
0.1 Ns/m2 and the one below it is having three times the dynamic viscosity.

Fig. E 1.2.4 Sliding of a plate

Solution Given. Area of the plate (A) = 5.0 m2, Applied force (F) = 150 N, Distance of the plate
from either surface (h) = 5 mm = 0.005 m.
Let y be measured from the plate surface, as shown in the Fig., so that the plate surface is at y = 0
and U be the velocity of the plate under equilibrium. Then the total drag force due to the two
liquids on the two sides of the plate should be equal to the applied force. This generates the Eq.

[μ(dU/dy)y = 0 + 3μ (dU/dy)y = 0]A = 150

→ [0.1 × U/0.005 + 3 × 0.1 × U/0.005] × 5 = 150

Or, U = 0.375 m/s.

Example 1.2.5 Obtain an expression for the torque on the inner cylinder of the rotating cylinder
viscometer. In an experimental set up, R1 = 100 mm, R2 = 108 mm, h = 200 mm, a = 4.0 mm, N
= 100 rpm, and T = 0.393 Nm. Determine the dynamic viscosity of the liquid.

Fig. E 1.2.5 Rotating cylinder viscometer

Solution The viscometer shown in Fig. E1.2.5 consists of two cylinders placed one inside the
other with space between the cylinders filled with some fluid whose viscosity is to be
determined. The outer cylinder is rotated while the inner cylinder is kept fixed. We can measure
the torque applied on the inner cylinder using the transducers or some other method of
measurement. (It is also possible to keep the outer cylinder fixed and rotate the inner cylinder.)
The fluid layer in contact with the outer cylinder rotates with the cylinder having circumferential
velocity 2πR2N/60 and the inner layer in contact with the stationary cylinder has zero velocity.


where T0 is the torque exerted on the inner cylinder due to the liquid contained in the vertical
space between the cylinders and rotation of the outer cylinder. The liquid in the bottom of the
viscometer will also transfer some torque to the inner cylinder. To determine it, we consider a
circular ring of radial width dr at a distance r from the axis of rotation. The layer in contact with
outer cylinder has velocity 2πr N/60 and will induce shear stress = 2π r Nμ/(60 a), whereas a is
the depth of liquid between bottom surface of the two cylinders, as shown in the Fig..

and Total torque (T) = T0 + T1

However, (R2 – R1) is small and hence the term . Therefore the first term T0 is
often not considered in the calculations. Substituting the values in the above relation, we get μ =
0.885 kg/ms.

1.2.1 Introduction to Stress and Strain


From Eq. 1.2.2, we know that the stress at any point in the flow region is defined by the first
derivative of velocity, which is a vector quantity. From vectors we know that a derivative of a
vector makes it a tensor of order 2. Thus a stress has nine components being represented in the
form

where τ is the symbol for stress. Double subscript notation is used with first subscript for the face
on which the stress acts and the second subscript for the direction in which it acts. For example,
τxz is the stress on x-face in z-direction. Recall the direction normal to the face is used to
represent the face. Area is considered positive when the outward drawn normal to the face is
towards positive axis. Stress on three faces, normal to which is along the positive direction, is
shown in Fig. 1.2.4; the cuboid has been drawn in the fluid region around a point. When the
cuboid is shrunk letting its side tend to zero then it gives the stresses at the point.

Fig. 1.2.4 Stress representation

The diagonal elements represent the component acting normal to the surface and in many books,
σ has been used in place of τ. Whereas the off-diagonal components act in the surface and are
named shear stress. We may imagine a force acting on the face being resolved along normal to
the surface and in the surface; the component in the surface being further resolved along the two
mutually perpendicular directions, which normally are parallel to the direction of axes of the
reference system used. The diagonal elements, in general, are not equal in magnitude. We define
their average as the thermodynamic pressure at the point to which the volume element would
shrunk in the limit. Thus,

In mechanics of solids, we read about the stresses and how are these related to the corresponding
strains. We may similarly define strain components for fluids too, representing it by the symbol
ε. There is however one significant difference in the behaviour of solid and fluid. When a shear
force is applied to a solid, it deforms and stays in deformed position so long the force acts and
comes back to original position after removing the force. Fluids, on the contrary, deform
continuously even if an infinitesimal shear force acts on it. It is found that the time rate of
deformation becomes constant in fluids. Therefore, we use time strain instead of strain. That is,
the symbol e stands for time rate of strain. It will not be difficult to show that it is nothing but the
velocity gradient ∂u/∂y. The corresponding strain rate components are represented as below:

A fluid is said to be isotropic when these relations do not depend on the direction of axes and
homogeneous when it is independent of the position of the fluid particle in the fluid region. We
may also establish that the stress tensor is symmetric, which means the off-diagonal elements are
equal. That is, τxy = τyx. In such case, the tensor components are reduced from 9 to 4 — three
shear stress viz. τxy, τxz, τyz, and one pressure component.

The Newtonian model is the simplest one, which holds true for a parallel flow. Stokes, an
English mathematician, extended it to a general flow, which is described by Eq. 1.2.2 for the
shear stresses. For normal stresses, he suggested the following mathematical model

where μ′ is another constant, which is determined using Eq. 1.2.5. Substitution of normal stresses
from Eq. 1.27 in Eq. 1.2.5 and subsequent simplification gives the average of the normal stresses
as

–p + 1/3 (2μ + 3μ′)

This will be equal to thermodynamic pressure p (compressive) if

the negative sign shows that the pressure will always be compressive. We shall later see in
Chapter 2 that for incompressible fluids it is zero and hence Eq. 1.2.8 is satisfied automatically.
But for any fluid the equation is also satisfied if 2μ + 3μ′ = 0, or μ′ = –2 / 3 μ. This shows that μ′
is always negative but the cases have been reported for the fluids for which it is positive.
However, in some cases this condition though may not be satisfied but the value of the terms in
Eq. 1.2.8 is negligibly small and may safely be ignored. Therefore, μ′ can be taken as –2/3 μ and
then Eq. 1.2.7 takes the form

We shall use these constitutive relations later for derivation of Navier-Stokes Eq.; Navier was a
French engineer. The Eq. 1.2.9 gives only the normal stresses. The other components similarly
were obtained and can be written as (in tensor notation)

where δij, the Kronecker delta = 1 when i = j and 0 when i ≠ j. If we take i = j =1 then Eq. 1.2.10
takes the form of the first equation for τxx in Eq. 1.2.9. Similarly, we get the other equations of
Eq. 1.2.9 by taking i = j equal to 2 and 3, respectively. When i = 1 and j = 2, Eq. 1.2.10 gives Eq.
1.2.2; substitute x = 1 and y = 2 in Eq. 1.2.2 to make comparison.

1.2.2 Other Properties of Fluids

Another property of importance in fluids is the surface tension. We know a matter is made up
from the molecules that move randomly in different directions; the motion is characterised as
chaotic motion. Each molecule is surrounded around it by other molecules and is being pulled by
a force, called the cohesive force. The molecules, which are at a depth more than the molecular
diameter from the free surface of the fluid, experience zero force, since the forces from all the
sides balance out. However, for the molecules in a thin strip of the size of molecules’ diameter
near the free surface, the force balance is disturbed as number of molecules above it are less than
below it. Thus there is a net force acting on such molecules pulling these molecules to keep them
within the fluid surface. It looks like as if there is a membrane on the fluid surface that does not
allow the molecules to escape. This phenomenon is called surface tension. If the molecule is to
be pushed out of the fluid, as in evaporation process, the molecule is to be given enough kinetic
energy to overcome this force of attraction. The amount of energy required to push the molecule
just out per unit area of the surface is called surface tension, designated by the symbol σ.
However, the surface tension, as used in practice, is defined in terms of the force exerted by the
free liquid surface per unit of its length instead of energy term and both are equivalent to each
other. Thus surface tension is a property of the fluid that depends on the fluid itself and its
temperature. It is easy to recognise that the amount of energy required will depend on the
closeness of the molecules, which means more force of attraction and hence more the energy to
be supplied or surface tension. This implies that the surface tension would decrease with increase
in temperature. When the surface tension is zero, the temperature is called the critical
temperature.3

Fig. 1.2.5 Effect of adhesive and cohesive forces on shape of interface

When the two fluids or a fluid and a solid form the interface then molecules of one media pull
molecules in the near surface of another media that lie in the region of influence. This force of
attraction is called the adhesive force. The relative magnitude of cohesive and adhesive forces
decides the shape of the interface. Fig. 1.2.5 shows interface between a solid surface and a liquid.
In case of mercury, the cohesive force is so strong that the mercury drops are almost spherical, as
shown in Fig. 1.2.5 (a). When the size of the drop is large than it is flattened somewhat and does
not remain spherical. If we draw a tangent to the interface surface at the point of contact, the line
makes an angle θ with the surface occupied by the drop. The angle θ is called the angle of
contact and is around 130° – 150° in the case of glass-mercury interface, instead of 180°,
because of impurities. Such fluids are classified as non-wetting liquids. On the other hand
adhesive forces are so strong in water that it wets the surface almost. Contaminated water may
have a convex surface with θ upto 25°. Effect of the angle of contact is seen in the meniscus of
the free surface of liquid in a tube. We note that while the meniscus is convex in mercury, it is
concave for water.

Water in contact with air has surface tension of 0.073 N/m at 20°C; for mercury in contact with
air, it is 0.05 N/m at 20°C. Adding some other chemical compounds can modify the surface
tension. For example, surface tension of water can be reduced considerably by adding a small
quantity of an organic compound such as soap or detergent and can be increased by adding a
small quantity of salt such as sodium chloride, etc. Reduction in surface tension allows seaping
of water through the pores in th cloth and remove dust particles easily. This explains how do
detergent, soap, etc help in washing the clothes.

In normal situations, the surface tension force is significantly small compared to pressure and
other forces to be considered in the Eq. of motion of the fluid. But in some applications like
rising of liquid in a capillary, bubble mechanics in water boiling, liquid-jet studies and flow
through a porous media, the surface tension force is comparable and cannot be neglected. Fig.
1.2.6 shows the forces in some of these cases.
Fig. 1.2.6 Capillary action in water and mercury

Consider a capillary tube dipped in water (Fig. 1.2.6a) or in mercury (Fig. 1.2.6b). It is observed
that the water rises while the mercury level goes down in the capillary tube. This action can be
observed when the diameter of the tube is small, less than 1 cm. The water level will be concave,
because of acute angle of the contact as discussed above, and the surface tension force σ acts in
the direction as shown. This force will have a vertical component that forces the water to rise till
it is equal to the weight of the water raised in the capillary. For the same reason mercury level
goes down the free surface. Balancing the forces, we get

πdσ cos θ = (πd2/4)h.ρg

where h = height of the liquid column in the capillary tube of diameter d and θ is the angle of
contact.

Now consider a spherical bubble of diameter d enclosing a fluid at pressure p. Since the pressure
acts in all directions, the net force in any direction would be zero. However, if we take a rim on it
then the pressure force above and below it acts in opposite direction and hence a force acts on it,
which is the surface tension. The net pressure force responsible for bursting apart of the bubble
into two parts is = p × area perpendicular to the pressure direction = p × (πd2/4). Equating it to
the surface tension force, we get

Recall if the spherical shell is a metallic one then the stress developed is called hoop stress. If
this shell has some thickness such that the two radii are r1 and r2 then
Fig. 1.2.7 Forces acting on a surface separating two immiscible fluids from each other

To get this expression for the fluids, let us consider a curved liquid surface separating the fluid
on convex side from the rest of the fluid, as shown in Fig. 1.2.7; r1 and r2 be the radii of
curvature of the convex and concave surfaces. Let the angle subtended by these at the centre of
curvature O be dθ1 and dθ2, respectively. Forces are shown in two-dimensional frame also, which
is self-evident. Let p1 and p2 be the pressure on the convex and concave surfaces. For the
equilibrium, the vertical forces must balance. This gives the Eq.

2σr2 dθ2 sin(dθ1/2) + 2σr1 dθ1 sin (dθ 2/2)

= (p2 – p1) r1r2 dθ1dθ2

If we neglect p2 then Eq. 1.2.14 gives Eq. 1.2.13. There is also possibility that a thin liquid
surface may be trapped between the same fluids, as in bubble. In this air is present all around the
liquid film. In that case the lower surface will also exert the same amount of the force as the
upper surface and hence the net force will be doubled. Then Eq. 1.2.14 changes to

Example 1.2.6 Make an analysis of the shape of the water-air interface near a plane wall, as
shown in Fig. E 1.2.6. Assume that the slope is small, 1/R = d2η/dx2, (where R is the radius of
curvature of the interface), and the pressure difference across the interface is balanced by the
product of specific weight and interface height as Δp = ρgη. Boundary conditions: wetting
contact angle θ = θ0 at x = 0 and θ = 90° as x → ∞. What is the height η at the wall?

Solution One of the surface of the interface is plane, i.e. r2 = ∞ whereas for the other surface
radius of curvature is R. Therefore Eq. 1.2.15 gives

Fig. E 1.2.6 Water air interface near a wall

Also

Δp = ρgη (given)

Integrating the equation two times with respect to x and using the given boundary conditions, we
get

This equation defines the shape of the interface. Substituting x = 0, we get

1.3 Pressure Variation and Hydrostatic Forces on Surfaces Submerged in Fluid at Rest
Fig. 1.3.1 Pressure variation

The pressure force acts at any point of the fluid (which is at rest) that depends on the depth of the
point and the gravity field. This can be understood by considering equilibrium of the fluid
element, as shown in Fig. 1.3.1. A cylindrical element of cross-sectional area dA and height dz is
shown. Then the mass of this fluid element is m = ρ dA dz, where ρ is density of the fluid; the
cylinder height is so small that density of the fluid may be considered as constant over its height.
Let pressure p acts on its lower face and p + dp on the upper face. Let z axis4 be taken positive
along vertically up direction so that the gravity is along –ve z-direction and its weight mg acts
downwards. Balancing the forces in the vertical direction, we get

This equation means that the pressure at any point depends on the depth of the point from the
reference point. It implies that (i) pressure increases if we move down, i.e. pressure increases
with the depth, and (ii) the pressure in a horizontal plane would be same and independent of the
shape of the container. If we neglect effect of gravity then pressure would be same, a result we
read very early in science course and known as Pascal’s law. This property of the fluid has been
effectively utilised in hydraulic press and many other hydraulic machines. Equation 1.3.1 can be
integrated if relation between density and height is known. Four cases are possible:

Case 1: Density is constant. In this case, integration of Eq. 1.3.1 from ground to any point z
above it results in

Remember that pressure would increase if we go down, since z will be negative in Eq. 1.3.2. For
the matter of fact, pressure increases as we move down the water surface in sea, pond or river. Or
if a dam is constructed across a river then pressure on its walls increases with increase in height
of the water. A designer has to take care of such forces while designing the wall. We shall
discuss this aspect in more details in the next section.

Fig. 1.3.2 Different types of manometers

Here we shall consider application of Eq. 1.3.2 to manometers, which are used for measurement
of pressure. Figure 1.3.2 shows a few arrangements of manometers. Procedure to determine the
pressure is simple and is based on the property that pressure on all the points on a horizontal
plane is same provided all these points are connected through fluid. We start from the end at
which connection is taken for measurement of pressure and traverse through the arrangement to
reach another end. While moving across one fluid to another we determine pressure at the
interface, which then becomes starting point for another interface. This process continues till we
reach the free surface. We also know that the pressure increases while going down and decreases
while moving up in the same liquid by a value ρgh, where h is the vertical height between the
two points in the same fluid region having density ρ. This way an equation is obtained, which is
called manometric equation.

Let us illustrate the procedure for the arrangement shown in Fig. 1.3.2 (a) for a U-tube
manometer, which is often used in laboratories for measurement of pressure. It is connected to
the vessel, where pressure is to be measured, at a point A and is filled with some liquid, called
the manometric fluid having density ρm. Due to pressure in the vessel, the manometric liquid
takes free surfaces B and C, as shown, such that the pressure at horizontal line through B is same
in both the limbs. We start from point A and reach the point C following the principle mentioned
above to get the manometric equation.

pA + ρgz – ρmghm = pC

or,

where subscripts A and C refer to the points A and C, and ρ is the density of the fluid in the
vessel that occupies the space above the manometric fluid in the U-tube. Since the C end of the
manometer is open to atmosphere, pc = 1 atm. Therefore, knowing the heights appearing in the
equation, pA can be obtained from this equation. Normally ρ << ρm, ρgz term is dropped in
general. However, one should justify this.

In case the pressure to be measured is too small then an inclined manometer and a manometric
fluid of lower specific gravity may be used so that l, the length of the fluid in the inclined limb of
the manometer, may be sufficiently large to read accurately. We may also adjust inclination
angle of the tube; a decrease will increase the length l. You can write easily the manometric
equation remembering that while moving along the inclined limb the vertical distance moved is
not l but l sin θ. The inclined manometer is shown in Fig. 1.3.2(b).

On the other hand, we may put a number of U-tubes in series to measure a large pressure. Figure
1.3.2(c) shows two tubes put together. Three fluids of densities ρ1, ρ2, and ρ3 are used; the
difference in the meniscus in the two limbs respectively is h1, h2, and h3, respectively. If the
pressure at the open end is pa then the manometric equation is

ρA + ρgz – ρ1gh1 + ρ2gh2 – ρ3gh3 = pa

pA – pa ≈ g (ρ1h1 – ρ2h2 + ρ3h3)

Case 2: Isothermal atmosphere. In atmosphere, density of air decreases as we go up. That is,
there exists some relation between density and height of the air column above the sea level. It is
known that temperature decreases uniformly in stratosphere upto height of 11 km after which it
is constant at that temperature up to 20.1 km; Figure 1.3.3 shows variation of temperature with
altitude and Appendix B1 gives details of the standard atmosphere, which of course varies from
location to location. For stratosphere, we may assume isothermal condition so that ρ = kp, where
k = 1/RT, a constant that can be determined knowing temperature of air in that zone. Substituting
for ρ in Eq. 1.3.1 and integrating from a point at hight z1 to any point z m further above it, we get

ln (p/p1) = –kg (z– z1)

or,

p1 is obtained by consedering pressure, density and temperature variation in the troposhere,


which is discussed next.
Case 3: Temperature decreases linearly with altitude z, as in troposphere. We know that the
temperature decreases uniformly with altitude near the ground surface upto 11 km. Let the
temperature at any height z from the ground be T = T0 – m z, where m is slope of the temperature
curve and represents the change in temperature per metre of the height from the ground, it is
called temperature lapse rate. Temperature lapse rate upto 11 km is 0.0065°C/m. Assuming the
air to obey the perfect gas equation, the density at any height z is given by ρ = p/{R (T0 – mz)}.
Substituting it in Eq. 1.3.1 and integrating the Eq. from a point on the ground surface to some
point at height z, we get

or,
Fig. 1.3.3 Variation of air temperature with altitude

The air in contact with the earth surface becomes hot because of heat convection and radiation
from the Sun and rises above where the temperature and pressure decrease. Air being bad
conductor of heat gets cooled adiabatically at a cooling rate of about 0.01 °C. This rate of
cooling is more than the natural temperature lapse rate in the troposphere. Therefore its density
increases and air falls back to the earth thus bringing stability in atmosphere. If however, it is not
so then the atmosphere is unstable that results in thunderstorm.

Case 4: Isentropic conditions. In many processes the time required for a process to occur, such
as propagation of wave in air, is so small that the process may be assumed to be isentropic for
which p = kργ, where k is a constant = p0/ρ0γ. Substituting it in Eq. 1.3.1, using the equation of
state and integrating we get after simplification


We may also find out temperature at any point using the equation of state.

p/ργ = p0/ρ0γ

and p = ρRT (Equation of state)

Differentiating Eq. 1.3.7 w.r.t. z, we get the temperature lapse rate

Example 1.3.1 An inverted vertical manometer is connected across two pipes carrying water as
shown in Fig. E 1.3.1. The manometer fluid is gasoline of sp. gr. = 0.796. If the pressure of water
in pipe A is 8.35 kPa (gauge) what is the pressure of water in pipe B?
Fig. E 1.3.1 Inverted manometer

Solution Inverted U-tube manometer is used frequently in laboratory when the pressure
difference between the two points is small — a manometer fluid of specific gravity less than that
of the fluid in pipes is to be used; one such arrangement is shown in Fig. E 1.3.1. We start from
one end, say A and reach the other end B to write the manometric Eq.. The equation with pA and
pB in kPa is

and

pB = 0.75 × 9.81 = 7.379 kPa.

In this Eq. w is specific weight of water in kN/m3.

Example 1.3.2 A glass tube of uniform bore is bent into the form of a square of sides a, and
filled with equal amounts (volume) of three immiscible liquids of densities ρ1, ρ2, and ρ3. It is
known that ρ1 < ρ2 < ρ3. If the tube arrangement is filled with the liquids, as shown in the Fig.,
show that (2ρ3 + ρ1)/3>ρ2 > (ρ3 + 2ρ1)/3. If the relative densities of the first and third liquids are
1.0 and 1.2 respectively, find the range of the relative density of the second liquid which makes
the above arrangement possible.

Fig. E 1.3.2 Three fluids in equilibrium


Solution The three liquids in equilibrium with each other are shown in the figure. Let AE = x.
Since the tube is of same area of cross section and liquids are having same volume, their lengths
will also be equal. This gives the following dimensions of the liquids in the tube:

Length of each liquid = 4/3 a, since the total length is 4a.

Liquid 2: AB = a, BF = 4/3 a – a – x = a/3 – x.

Liquid 1: ED = a – x and DG = 4/3 a – (a – x) = a/3 + x.

Liquid 3: CG = a – DG = 2a/3 – x and CF = 4/3 a – CG = 2a/3 + x.

For equilibrium, the pressure at the interface F from both the sides must be equal. This condition
demands:

Pressure of the column GD + pressure of the column CG = pressure of the column AB.

ρ1g(a/3 + x) + ρ3g(2a/3 – x) = ρ2ga

But from the given condition of equal lengths of the three liquids, we infer that 0 < x < (4a/3 – a)
= a/3. It implies that

From these two inequalities, we infer (2ρ3 + ρ1)/3> ρ2 > (ρ3 + 2ρ1)/3.

Substitution of ρ1 and ρ3 in the inequalities gives after simplification: 1.0667 < ρ2 < 1.1333.

1.3.1 Pressure Force on a Plane Surface


Fig. 1.3.4 Plane surface immersed in liquid

Let us consider a plane surface AB in liquid, as shown in Fig. 1.3.4, immersed such that it makes
an angle θ with the free surface. The surface, seen along normal to AB, is also shown just below
it. Arrows show pressure acting on the surface, which is perpendicular to the surface because the
fluid is at rest. Magnitude of pressure would change along its length, since depth of the surface
from the free surface is increasing; a free surface is represented by the inverted triangle symbol
Δ. Pressure at any depth z is given by ρgz. Force acting on area dA of the strip considered at
depth z is δF = ρgz dA = ρgy sinθ. dA, where y and θ are shown in the Fig.. Summation of these
forces gives the total force acting on the surface. Taking moment about the point O, we can
locate position of the force acting on the surface. This procedure yields:

where is depth of centre of gravity from the free surface. Taking moments about O, we get

But z = y sin θ and Io = Ic + A 2


and

Here I0 is the second moment of plane surface about an axis perpendicular to the plane of paper
and passing through the point O, the point where the plane surface would meet the free surface
when extended. Ic is the second moment of area about the axis passing through the centroid of
the area and parallel to the axis thorough O and the two moments are related to each other by the
theorem of parallel axes. Table 1.3.1 gives the second moment of area about the axis passing
through the centroid.

TABLE 1.3.1 Geometric properties of some lamina


Equation 1.3.10 tells that the centre of pressure, the point on the surface where the total force is
assumed to act, lies below the centre of gravity, since all the quantities are positive. This was
expected since the pressure force on downside is more and hence its moment will be more.

We get some useful information from these two equations that are summarised below.

1. The component of the force on the plane surface can be resolved along and perpendicular
to the free surface. These components are F sin θ and F cos θ, respectively. Therefore, Fx
= ρg (A sin θ). But A sin θ, as seen in the figure, is the projection of area A of the
surface along the free surface, that is, it gives projected area of A on the vertical plane.
We may similarly determine the point of application of the force. Thus to find out the
horizontal force along the x-direction, we resolve the inclined surface along the vertical
direction and then determine the magnitude and the point of action of the horizontal force
on this vertical surface.
2. Similarly, the force in the vertical direction can be determined by taking the horizontal
projection passing through the centroid. This gives the vertical force = ρg × (area of the
horizontal plane). You may notice that this is the weight of the fluid contained in the
cylinder over the horizontal projected area, passing through the centroid of the surface,
upto the free surface.

Fig. 1.3.5 Force at the base of different shaped beakers

3. An issue related to the weight supported over a horizontal surface is about the force on
the base of beakers having different shapes, as shown in Fig. 1.3.5. These beakers have
the same base area and are filled by the same liquid upto the same height. We know the
hydrostatic force on the base in all these cases will be same though the actual weight of
the fluid will be different. The weight of the liquid above the base contained in a cylinder
above it is called the virtual weight and this weight is the vertical force on the beaker’s
base.

We now know that the force on any surface can be determined by determining the force on its
horizontal and vertical projection; the horizontal projection is taken to pass through the centroid.
The pressure on the vertical projection increases with depth as shown in Fig. 1.3.6 (a). When the
surface is fully dipped in a single fluid, the pressure variation is along the slant line. In this
figure, AD represents the projection with ends A and D at a depth of z1 and z2, respectively. The
pressure at these edges would be AB = ρgz1 and CD = ρgz2. The pressure distribution is a
trapezoid whose area is ρg (z1 + z2) (z2 – z1)/2 = ρgh + , where h = z2 – z1 = height of projected
area and = (z1 + z2)/2 is depth of CG from free surface. This is the magnitude of the pressure
force per unit width of the vertical surface AD acting on it. This is an important observation in
the sense that the computation time can be saved when there are layers of the fluid one over the
other. We may also show, in the same way that the centre of pressure lies at the centroid of the
trapezoid.

Fig. 1.3.6 Vertical surface in fluid

However, when the surface lies in two or more fluids then slope of the pressure curve would
change at the interface. Figure 1.3.6 (b) shows pressure variation over the vertical surface dipped
in two fluids having densities ρ1 and ρ2, ρ2 > ρ1. Pressure varies linearly with the depth from the
free surface of the top liquid till the liquid interface is reached and there after it remains constant.
However, the slope of the pressure curve is changed at the interface and the second fluid adds its
contribution to the pressure that varies linearly with the depth from the interface. The pressure
variation is as shown in the figure. We may divide this area into three parts A, B and C whose
area and centroid can easily be determined. The sum of the three areas gives the total force. To
determine the centre of pressure, we take moment about the point D and equate it to the moment
of the total force. While taking moment we may take the distance of the centre of gravity from
the vertical wall or the horizontal line passing through D thus giving exact location of the centre
of pressure. This problem can also be handled by considering the surface submerged totally in
liquid at the bottom and converting the height of the liquids above into the equivalent head of
this liquid. This procedure is explained in Example 1.3.6. This is one of the many methods that
have been used. One can also find out the pressure force by assuming a fluid of density ρ1 over
the entire surface and another fluid of density (ρ2 – ρ1) over the portion dipped in fluid of density
ρ2. Pressure force due to these fluids is determined as if other fluid is not present and then it is
summed up.

Example 1.3.3 Figure E1.3.3 shows the vertical trapezoidal surface that stores water up to height
h m. Determine the centre of pressure and the total force on the surface.
Fig. E 1.3.3 Vertical trapezoidal surface

Solution The surface in Fig. E1.3.3 shows the cross-section of the wall which stores water. The
total force and its point of application is given by Eqs. 1.3.9 and 1.3.10, which need the cross-
sectional area and the depth of the centroid of the area. This can be calculated by dividing the
area into two triangles ADE and BCF and a rectangle DEFC. Their properties are obtained from
Table 1.3.1. Area of the triangle is ½ × b/2 × h and the centroid is at h/3 from AB.; G1 and G2 are
the centroid. Area of the rectangle is a × h and the depth of the centroid G3 is h/2 from AB.
Therefore, the total area of the trapezoid is

To find the depth of G from the free surface, take moments of areas about AB. This gives


Taking moments of the forces about CD, we get

Example 1.3.4 The automatic tipper shown in Fig. E 1.3.4 operates when the water level reaches
a certain height. Calculate the water level.

Fig. E 1.3.4 Tipper gate

Solution Tipper gate operates automatically as soon as a particular water level reaches a certain
height upstream. Thus it is possible to store water in the canal upto a certain height. If the water
level increases, the gate opens and water flows down the canal. Let h is the height of the water
level at that moment. We can see that the centre of pressure must lie above the hinge O at such a
distance that its moment is more than the moment of the weight of the gate. Since the weight is
not given, the centre of pressure must lie at or above O. Let us consider a unit span5 of the tipper.
Then

= h/2, A = h/sin 60 × 1

= 1.16h

Ic = × 1 × (1.16h)3

zp = h/2 + 0.13h3 × sin2 60/(1.16h × h/2)

= 0.67h.

But the hinge at which C.P. would lie is at a depth of (h – 1)m. Equating it to zp obtained above,
we get

h = 3.03 m.

Example 1.3.5 A pipeline 4 m in diameter contains a gate valve. The pressure at the centre of the
pipe is 200 kPa. If the pipe is full of oil of specific gravity 0.87, find the force exerted by oil on
the gate and the position of the centre of pressure.

Solution A gate valve may be considered as plane circular plate at whose centre the oil pressure
is 200 kPa. This pressure at the centre is equivalent to a liquid head z = 200 × 1000/(870 × 9.81)
= 23.43 m. That is, the centre of gravity of the surface is at a depth of 23.43 m, which is z in Eqs.
1.3.9 and 1.3.10. Further area A = π/4 × 42 and Ic = π/4 × 44.

F = 870 × 9.81 × 23.43 = 2512.87 kN

and
= 23.43 + 0.043 m.

That is, the centre of pressure lies below the gate valve centre at a depth of 0.043 m.

Example 1.3.6 Find the total pressure and the depth of the centre of pressure for the vertical
lamina shown in Fig. E1.3.6. The upper triangle is dipped in a liquid of specific weight 500 N/m3
and the lower one in a liquid of specific weight 1000 N/m3. Both the triangles are equilateral and
equal in size.

Fig. E 1.3.6 Centre of pressure

Solution We shall attempt the solution using the concept of equivalent head. Further, the area
and the second moment of area about the axis through the C.G are given by 1/2 b × h, bh3/36,
respectively, where b and h are the side and the height of the triangle. The height of the triangle
is given to be 3 m from which we get the side of the triangle b = 3/sin 60 = 2√3 m. For the upper
triangle

1 = 2/3 × 3 = 2 m

A= ×2 ×3=3

and

Ic = 2 × 33/36 = 3 /2

F1 = 500 × 3 × 2 = 3000 N

and

zp,1 = 2 + 3 /(2 × 3 × 2) = 2.25 m


We may replace the upper liquid by the heavy liquid having its free surface at 3 × 500/1000 =
1.5 m. That is, we assume the lower triangle is at a depth of 1.5 m from the free surface of the
liquid having specific weight 1000 N/m3. We use Eqs. 1.3.9 and 1.3.10 to give for the lower
triangle

2 = 1.5 + 1/3 × 3 = 2.5 m, A = × 2 ×3=3

and

Ic = 2 × 33/36 = 3 /2

F2 = 1000 × 3 × 2.5 = 7500 N

and

zp, 2 = 2.5 + 3 /(2 × 3 × 2.5) = 2.70 m.

We note that the equivalent free surface is at a depth of 3 – 1.5 = 1.5 m of the original given
condition. Therefore the centre of pressure of the lower triangle is at a depth of 1.5 + 2.7 = 4.2 m
from the original free surface. We may now proceed to find out the total force and the point of its
application by taking moments about the free surface. This gives

F = F1 + F2 = 18186.533 N

and

F × zp = F1 × zp, 1 + F2 × zp, 2

or,

18186.5 × zp = 3000 × 2.25 + 7500 × 4.2

zp = 3.642 m.

Example 1.3.7 Determine the pressure force exerted by a liquid of specific weight w and the
point of its application on a vertical parabolic lamina shown in Fig. E 1.3.7.
Fig. E 1.3.7 Force on a parabolic lamina

Solution Area of a parabola = 2/3 × area of the corresponding rectangle = 2/3 × l × h = 2lh/3.
Depth of the centroid from the free surface is = 3/5 h. With this information we can find out the
hydrostatic force on the lamina F = w × 2lh/3 × 3/5 × h = 2 w l h2/5. To find out the point of
application of the force, consider the origin of the coordinate axes at O with y along the down
side and a strip of width dy at a depth y from the free surface. The parabola is represented by the
Eq. y = 4h (x/l)2; you may note that when y = h, x = ± l/2. we shall find out the force on this strip,
then take the moment about the free surface and finally sum up these moments to get the total
moment of the force. This moment should be equal to the moment of the total hydrostatic force
on the lamina. This procedure gives

dA = 2x dy, dF = w.2x dy.y

But

M = F × zp → zp = 5h /7.

Example 1.3.8 A hemispherical depression of diameter 1.2 m exists on one of the vertical sides
of a tank. If the tank contains water to an elevation of 3.0 m above the centre of the hemisphere,
calculate the forces acting on it.
Fig. E 1.3.8 Forces on a hemispherical projection in a tank

Solution

Vertical force on the projection = weight of water NSTQ – weight of water MSTQ
= weight of water contained in the hemisphere

or.

Horizontal force = force on the vertical circular section MN

FH = w × πR2 × 3 = 132.9 kN.

Example 1.3.9 The gates of a lock gate 5 m wide and 5 m high include an angle of 120° when
closed. Each gate is held on two hinges, one placed at the top and the other at the bottom of the
gate. If the water levels are 4.5 m and 3 m on the upstream and downstream sides respectively,
determine the magnitude of the forces on the hinges due to the water pressure.
Fig. E 1.3.9 Lock gate

Solution A lock gate is used in sea to maintain different water levels in a dockyard on its two
sides. The ship is brought to the dock area from the sea or vice versa by operating one or the
other gate to maintain water level as per the situation. The level of the ship can be lowered or
raised accordingly. This can also be used to store water during the tides; while tide is over then
the gates can be closed not to allow the water to go back to the sea. Lock gate has been used to
connect two seas, which are having water at different levels, by a canal carrying the lock gate at
both the ends. A canal is constructed across the land separating two seas and lock gate is placed
at each end of the canal. When the ship is to be moved from sea at higher level then the gate
close to this sea is opened and other gate is closed. This raises water level in the canal to bring it
to the level of the sea. Then the ship is moved. Now this gate is closed and when the ship is to
move out, the other gate is opened bringing its level to that of this sea so that the ship can easily
move.

Figure E1.3.9 shows two views of the lock gate. Water is at a level of 4.5 m on the upstream side
and 3 m on the downstream side. Thus there is a net force F on each gate whose close position is
shown in the plan. Because of this force, the edges of both the gates are pressed against each
other resulting in a reaction force of magnitude Ro acting on each gate in the direction shown. If
we consider one gate then there are two forces namely F and R0. For equilibrium of the gate, a
third force Rh must act on it which is provided by the hinges. We may determine the force F by
determining the force and its position that acts on each side of the gate due to hydrostatic
pressure. The gate may be considered as rectangular plate having width 5 m. Calculations can be
made in the following tabular way using Eqs. 1.3.9 and 1.3.10 and Table 1.3.1, y is measured
from the bottom of the gate.

Upstream Downstream
h 4.5 m 3m
F 1000 × 9.81 × (4.5/2) × (5 × 4.5) = 497 kN 1000 × 9.81 × (3/2) × (5 × 3) = 221 kN
Upstream Downstream
zp (2/3) × 4.5 = 3 m (2/3) × 3 = 2 m
yp 4.5 – 3 = 1.5 m 3–2=1m

Note that when the edge of the rectangle of width b and depth l is in free surface then A = bl, =

l/2 and . Therefore, which has been taken in above


calculations. The net force on the gate is F = 497 – 221 = 276 kN. To locate the height of this
force from the base, take moments about the bottom hinge to give

276 y = 497 × 1.5 – 221 × 1

y = 1.90 m from the base.

The force Ro will act normal to the centre line and therefore makes an angle 30° with the gate.
Thus the forces F, Ro and Rh should meet at a point, B. The triangle of forces can be drawn,
which is shown enlarged. We note that the triangles BCD and ABD are congruent because the
side AD = DC, the side BD is common to the two triangles and ∠ADB = 90°. Therefore ∠BCD =
30° and Ro = Rh. Resolving the forces in the direction of F, we get

F = 2Rh sin 30 → Rh = 276 kN.

Not only the three forces meet at a point, they should also lie in the same plane. Therefore we
conclude that the net force exerted by the hinges is 276 kN which acts in a plane at a distance of
1.90 m from the base. Let the force in the top and bottom hinges be RT and RB, respectively.
Taking moment of the hinge forces about the bottom hinge, we get

RT × 5 = 276 × 1.9

RT = 105 kN and RB = 276 – 105 = 171 kN.

Example 1.3.10 The hydrostatic water pressure acts only on one side and to a depth of 10 m
from the top of a dock gate that is reinforced with three horizontal beams. Locate the position of
these beams so that each of these is equally loaded.
Fig. E 1.3.10 Dock gate

Solution Water is upto a height of h3 on one side of the dock gate, which is shown in Fig. E
1.3.10. We divide the vertical height in three sections such that the hydrostatic force is equal on
each of these. We have learnt how to determine the forces and the location of these forces. We
may start ab-initio for which we consider an area element of height dh at a depth h. We can
easily get the forces on these sections; these are

The total force on the dock gate can also be determined similarly to give .

We know h3 = 10 m (given). Therefore the above equations can be solved to give h1 = 5.77 m
and h2 = 8.16 m. Substituting these values, the force on any section is .

Let us now determine depth of centre of pressure for the second section. The depth of the centre
of pressure for the first section is 2/3 h1 = 3.846 m and that of both the first and second section is
2/3 × h2 whereas the total force on the combined section is . We know the moment of the
total force about the free surface is equal to the sum of the force in the individual sections. This
gives the depth h2,p of the centre of pressure of the second section from the free surface

h2,p = 7.4 m.
Similarly we may find out the depth of the centre of pressure of the third section, which comes
out 9.1 m. At these depths we have to provide horizontal beams that can withstand a load of
163.3 kN.

Example 1.3.11 A 4 m square gate provided in an oil tank is hinged at its top edge, as shown in
Fig. E1.3.11. The tank contains gasoline of specific gravity 0.7 upto a height of 2 m above the
top edge of the gate. The space above the oil is subjected to a negative pressure of 6865 N/m2.
Determine the necessary vertical pull to be applied at the lower edge to open the gate.

Fig. E 1.3.11 Negative pressure at the free surface

Solution In this problem the pressure at the free surface is negative, which can be handled by
decreasing the height of the gasoline in the tank by some height of the gasoline equivalent to the
negative pressure. The equivalent height h is given by the relation h = p/w = 6865/(700 × 9.81) =
1 m of gasoline. Now, rest of the process of calculating the hydrostatic force F and its point of
application on a plane surface is straightforward.

= (2 – 1) + (4/2) sin 45 = 2.414 m, A = 4 × 4 = 16 m2,

F = wA = 700 × 9.81 × 16 × 2.414

= 265150 N and

Depth of the centre of pressure from the hinge is 2.804 – 1 = 1.804 m. Now taking moments
about the hinge (this will eliminate the moment of the force acting on the hinge),
P × 4 sin 45 = F × 1.804/sin 45 → P = 239.3 kN.

Example 1.3.12 (a) A masonry dam of trapezoidal section is 9 m high with a top width of 1 m
and a bottom width of 6 m, the water face of the dam being vertical. If the depth of water
impounded is 8 m, find the extreme pressure intensities at the base.

(b) If the height of the dam is increased to 11 m by adding an additional 1 m × 2 m deep


rectangular section and the depth of water impounded is raised to 9 m, what would be the
extreme pressure intensities at the base? Masonry and water weigh 22000 N/m3 and 9810 N/m3,
respectively. What should be minimum coefficient of friction between the masonry and the
ground not to allow slip of the dam?

Fig. E 1.3.12 Pressure intensity at the dam base

Solution Section of the dam is shown in Fig. E1.3.12. The various forces that act on the structure
are: (i) the water pressure force on the water side of the dam that acts at the centre of pressure
and (ii) the weight of the dam. The water force generates a overturning couple resulting in high
pressure at the dam base on the far-side corner A and also tries to move the dam in the direction
of the force. Therefore the dam is likely to fail (i) by sliding on the soil on which the dam is
made, (ii) by overturning, (iii) tensile stress developed at the nearby end B of the dam and (iv)
due to excessive compressive stress at the point A. We have to ensure that the dam is safe under
all the conditions.

We shall make calculations for 1 m span of the dam. We can easily get the following values:6

Weight of dam, is the distance


of C.G. from the vertical face.
Hydraulic force, from the bottom

The centre of gravity and centre of pressure will be at different points but in the same plane.
Their resultant R, which meets the base of the dam at point C at a distance x from the base corner
B, can be determined by taking moments about the point B. For equilibrium of the dam, an equal
and opposite force will be provided by the base of the dam having the vertical component equal
to W and the horizontal component equal to P that passes through B and hence its moment being
zero. Writing the moment Eq., we get

W + P × h/3 = W.x

→ x = 3.256 m and e, eccentricity = x – b/2 = 3.256 – 6/2 = 0.256 m,

where b is the base width. The vertical component W of R acts at a distance e from the centre of
the base and causes bending stress. The compressive stress due to weight of the masonry dam
and the bending stresses are

Compressive stress = W/(b × 1) and

maximum bending stress = M/Z = ,

where Z is the section modulus of the base of a rectangle, along the axis perpendicular to the
plane of the section passing through O, the mid-point of the base of the dam. We note that the
minimum stress is at B and maximum at A. The stress at point A should not become negative and

at B not to exceed the safe compressive limit. This gives the condition or, e ≤ b/6

In the question, however, we have to determine the stress intensity, which comes out to be

= 145 kN/m2 (max) and 85.93 kN/m2 (min).

We note that the minimum stress is more than zero and hence the dam is safe at both the ends A
and B, since the compressive stress, RCC can take is much higher. We may also check for
slipping of the dam because of the hydrostatic force. This requires P ≤ μW, where μ is the
coefficient of friction at the surface of the base. For the given problem, μ must be ≥ P/W =
313920/693000 = 0.453 for no-slippage of the dam.

(b) When the height of the dam is increased to 11m then the above procedure can be repeated.
The difference comes only in determining the weight of the masonry.

Weight of dam, W = weight of trapeziod + weight of rectangle


= 693 + 44 = 737 kN;
1 = 2.048 m and 2 = 0.5 m

= 1.956 m from vertical face

Hydraulic force,
= 397.305 kN zp = 9/3 = 3 m from the bottom
z = 3.752 m and e = z – b/2 = 0.752 m.

Extremne pressure intensity = 215.204 kN/m2(max) and 30.463kN /m2 (min)

1.3.2 Pressure Force on a Curved Surface

Fig. 1.3.7 Hydrostatic force on curved surface

Let us now consider a curved surface immersed in a liquid of density ρ, as shown in Fig. 1.3.7. In
this case the pressure does not only change in the magnitude but also in the direction along the
surface. Consider an elemental area δA at a depth z from the free surface where p = ρgz and it
acts in a direction making an angle θ with the y-direction. The force in y-direction is Fy = p × (δA
cos θ). But δA cos θ is the projected area of δA on the vertical plane. Similarly, δA sin θ is its
projected area on the horizontal plane. Therefore we may take the projected area of the curved
surface on the horizontal and vertical planes and determine the force on these plane surfaces, as
we did in the case of plane surfaces.

It is interesting to note that the area of projection of the closed curved surfaces on such planes
would be zero, since the projection of surface on one side of the vertical plane passing through
the centroid points in one direction and that on the other side in other direction. These projected
areas are equal in magnitude but opposite in direction. As far as horizontal force is concerned,
the product of pressure and the elemental area at the same depth on both sides of the vertical
plane would also be same and the y-component of the force would be zero. Hence, the sum
would be zero. However, while concluding like this for the vertical force, product of the pressure
and area element on both sides of the horizontal plane passing through the centroid will not
cancel out as the pressure at these two area elements would be different due to their different
depths from the free surface. The force on the lower-half portion will be greater than that on the
upper part.

Consider the vertical component of the pressure force on element δA; it is w z × δA sin θ equal to
weight of the liquid over the projected area of δA on the x-y face passing through the element.
Now summing this force all over the surface, the sum will be equal to the weight of the volume
of the liquid displaced by the closed surface. This can be easily understood by dividing the whole
surface by a horizontal plane passing through the centre of gravity. The force on the upper
portion acts downwards while that on the bottom side it is upwards and these forces will equal to
the weight supported by these curved surfaces upto the free surface of the liquid.

This is the Archimedes principle. As a result of this vertical force, acting in the upward direction,
the weight of the body is reduced by the amount equal to this force, which is called buoyant
force or the force of buoyancy. The force acts at the centroid of the displaced liquid. The body
will float and dip partially in a liquid if the buoyancy force is more than the weight of the body
otherwise it will sink. The maximum buoyancy force will act when the body is fully immersed in
the liquid. The vertical component of the force on the body requires determination of the area
above the surface upto the free surface for which we have to use the knowledge from the
geometry and divide the whole area in such a way that these relations may be used effectively.
This is explained with the help of solved problems.

Example 1.3.13 Obtain an expression for change in height of immersion of a hydrometer and
relative density of the liquid.

Fig. E 1.3.13 Hydrometer

Solution Hydrometer is used for measurement of specific gravity of a liquid. It consists of a bulb
with a long neck fitted with scale. Initially it is immersed in water and it floats; reading at the
free surface of the water is taken. Now it is dipped in the liquid whose density is to be obtained
experimentally. If the liquid is heavier then the neck of the hydrometer would go up. Let Δh be
the difference in reading in the two cases, S the specific gravity of the liquid, V the volume of the
water displaced by the hydrometer, W the weight of the hydrometer, and A the cross sectional
area of the neck. We assume that the bulb of the hydrometer is always under the free surface of
the liquid and the neck is peeping out. Since in both the cases, the hydrometer is floating, we
may write that the buoyant force is equal to the weight of the hydrometer. This gives

W = wV, and W = wS (V – AΔh)

or,

For a given hydrometer A and V are fixed, knowing Δh, S can be calculated. This relation is not
linear and hence it is difficult to put a scale on the neck of the hydrometer.

Example 1.3.14 Fig. E 1.3.14 shows the section of a tank which is 2 m long and is full of water
under pressure at 19.62 kN/m2. A cylinder of radius 1 m is resting on the tank at one of its
corners as shown. Find the horizontal and vertical components of the force acting on the curved
surface ABC of the cylinder. Neglect the weight of the cylinder and assume it is to be empty.

Fig. E 1.3.14 Pressure on curve ABC

Solution Given OA = OC = 1 m and cylinder rests at one corner ⇒ CD = 1.5 m. Therefore, OD =


1.5 – 1 = 0.5 m.

→ sin θ = ½, or θ = 30° and AD = OA cos 30 = 0.886.


We may add a water column of height h corresponding to pressure 19.62 kN/m2 in the tank,
which is wh = 19.62 × 1000 → h = 2 m. this is shown in Fig. E 1.3.14 (b). To find horizontal
component of the force, we take projection of the surface ABC on the vertical plane, which is a
rectangle of height CD = 1.5 m and width 2 m, the length of the cylinder, with its top surface 2 m
below the free surface.

Px = wA = 9.81 × (1.5 × 2) × (2 + 1.5/2)


= 80.9325 kN.

Py = Upward force on BC – downward force on AB


= Weight of water above BC – Weight of water above AB
= Weight of water in the block A′ABCO′
= w × area of block A′ABCO′ × length of cylinder.
= 9.81 × {Rectangle A′ADO′ + triangle OAD + sector OABC} × length of cylinder
= 9.81 × {1 cos 30 × 2 + 1/2 × 0.5 × AD + (120/360) × π × 12} × 2
= 58.776 kN.

Example 1.3.15 Figure E 1.3.15 shows the section of a dam. The water face of the dam is
parabolic following the law y = x2/4. Find the horizontal and vertical component of the force
exerted by water on the dam per metre length. Find also the resultant force on the dam per metre
length.

Fig. E 1.3.15 Force on parabolic dam

Solution
Px = wA = 9.81 × (9 × 1) × 9/2
= 397.305 kN.

Py = weight of water over the parabolic profile upto the free surface

= 9.81 = 353.16kNs
Resultant force = = 531.576 kN and
tanθ = 353.16/397.305

θ = 41°387 with the horizontal.

Example 1.3.16 A masonry weir is of trapezoidal section with a top width of 2.0 m and of height
5 m with upstream slope of 1 vertical in 0.1 horizontal and a down stream slope of 1 vertical in
0.75 horizontal. If the weir has stored water up to its crest on the upstream side and has a tail race
of 2 m depth on the downstream side, calculate, per unit length of the weir, (i) the resultant force
on the base of the weir, and (ii) the minimum and the maximum stresses on the base of the weir.
Assume specific weight of masonry as 22 kN/m3, that of water 9.79 kN/m3 and neglect uplift
forces.

Fig. E 1.3.16 Maximum and minimum

Solution For convenience, we divide the whole weir region into three geometrical sections,
namely triangular section ACF (I), rectangular section (II) and another triangular section (III).
The horizontal and vertical forces acting on these sections can be determined; these are tabulated
in the following table. Note we are given slope of the faces AC and BD from which we determine
their horizontal projected length, which are 5 × 0.1 = 0.5 and 5 × .75 = 3.75m, respectvely.
Table Example 1.3.16. Calculation of forces and their point of application

Note: The distances in the above table are those which would be needed to determine the
moment of the forces about the axis through the point B. You should be able to get the data as
given in the table. From the table we get

Z, sum of vertical forces = 480.7 kN

and X, sum of horizontal forces = 102.8 kN

R, Resultant =

= 491.6 kN

and

tan θ = Z/X

θ = 77.93°

θ is the angle made by the resultant force on the weir with the horizontal direction and is acing
towards the base of the weir, as the vertical force is downwards. Now we may locate the point of
the resultant force from point B by taking moments of the vertical forces about B; (the
anticlockwise moment is taken positive); the moments are given in the table. This gives

x = ∑ M/Z = 1614.2/480 = 3.358 m.


b = distance of midpoint of weir base from B = 3.75 + 2 + 0.5 = 6.25 m

e, eccentricity = x – b/2 = 3.358 – 6.25/2


= 0.233 m.

The maximum and the minimum stresses are:

σmax = 94.1 kPa and σmin = 59.7 kPa

Example 1.3.17 A sector gate, as shown in Fig. E 1.3.17 under the equilibrium conditions, is of
radius 4 m and length 5 m. It controls the flow of water in a horizontal channel. Determine the
total force on the gate.

Solution From the figure, CD = OC sin 30 = 2 m and length of the gate is 5 m (given). Also OD
= 4 cos 30 = 4√3/2 and ED = 4 – OD = 4(1 – √3/2)

Therefore, the horizontal component of the force is the force on vertical rectangular plate CD of
width 5 m.

∴ Px = 9.81 × (2 × 5) × (1 + 2/2) = 196.2 kN

and zp = 2 + = 2.167 m.

The vertical component can be determined by determining the weight of the water in the block
ABDCEA. But
Fig. E 1.3.17 Forces on radial gate

Volume ABDCEA = volume ABDEA + volume of sector OCE – volume of triangle OCD

= {(1 × ED) + (30/360) × (π/4) × 42 – ½ × OD × CD} × 5

Py = w × Volume ABDCEA = 61.8 kN.

The resultant force = √(Px2 + Py2) = 205.7 kN.

We obtain its point of action from the vertical wall AE by taking moment of areas about AE.

Alternative procedure. We may determine the forces by considering an element δA at an angle θ


subtending angle dθ, as shown in Fig.. Depth of this element from the free surface is h = 1 + 4
sin θ and δA = 4 dθ × 5 = 20 dθ. The horizontal and the vertical components of the force acting
on this element are δF cos θ and θF sin θ, respectively. integration w.r.t. θ from θ = 0 to 30° =
π/6 radians gives

Px = 9.81 × (20dθ) × h × cos θ = 196.2 kN

and

Py = 9.81 × (20dθ) × h × sin θ = 196.2 kN

Example 1.3.18 A steel rectangular block of specific gravity 7.85 floats at mercury water
interface such that its side in water and mercury is a and b respectively. Determine the ratio a/b
if specific gravity of mercury is 13.57.

Solution For floatation of the block at the interface, the weight of the block should be equal to
the sum of the weight of the displaced mercury and water. Therefore, if A is the cross-sectional
area of the block,

A × (a + b) × 7.85 × g = A × [b × 13.57 + a] × 1000 × g


a/b = 0.835.

1.4 Metacentric Height

Stability study of a floating body is of importance because of the vessels floating in the sea,
aeroplanes flying in the air and many other such applications. By stability of a body we mean
that if any temporary extraneous displacement acts on the body then it should return back to the
original position from its disturbed position. If so then the bodyis said to be in stable equilibrium.
There are also possibilities that the body maynot return to the same position and takes some other
position; it is said to be in unstable equilibrium. Or the body may stay in any position and is in
neutral or metastable equilibrium.

Fig. 1.4.1 Different positions of a ball

To understand equilibrium, let us imagine a ball placed in position A at the top of a hill, as shown
in Fig. 1.4.1. An infinitesimal disturbance would displace it from this position on either side. Let
us suppose it moves to its right then it will finally stay at any point on the horizontal portion BC
of the hill if it does not cross over the point D to move further down. Before coming to rest it
oscillates a few number of times in the valley. We conclude that the position A is unstable, since
an infinitesimal disturbance will move the position of the ball to some new position. The position
of the ball anywhere on BC represents neutral equilibrium situation. However, if instead of a
horizontal portion there is a curved surface at BC then the ball will always stay at the bottom of
the curve and it corresponds to the stable equilibrium position. We also observe that position A
has higher energy compared to a point on BC. In other words the ball will take the position of
stable equilibrium where it has minimum energy. This can be generalised for any system and for
linear or angular motion.

Fig. 1.4.2 States of euilibrium of a floating body


Now let us consider equilibrium of a floating body having weight W that acts at its centre of
gravity C. The portion of the body dipped in the liquid will displace it. As a result, buoyancy
force FB acts on the body at the centre of gravity of the displaced liquid. The point B at which the
buoyancy force acts is called the centre of buoyancy. Initially, in the equilibrium position of the
body, the point B will lie vertically below C, say at B; draw a line BC joining B with C. Let the
body be displaced clockwise by some angle so that the line BC takes the inclined position, as
shown in the Fig.. In this position, the buoyancy force acts at the point B1; the vertical line
through B1 meets the line BC at point M. The point M is called metacentre and the distance CM
metacentric height. Three possibilities arise, which are:

1. When M lies above C, i.e. CM is positive. In this case, see Fig. 1.4.2 (a), the weight and
the buoyancy force will generate anticlockwise couple, called restoring couple, and the
body would return back to the original position.
2. When M lies below C, i.e. CM is negative. From Fig. 1.4.2 (b), we observe that the
couple produced by the weight and the force of buoyancy tilts the body further in the
clockwise direction making it unstable.
3. When M lies on C, i.e. CM is zero. In this case, as shown in Fig. 1.4.2(c), there is no
restoring couple and the body will stay in the displaced position.

We may also verify the above conclusions if the initial disturbance rotates the body about any
axis in the anticlockwise direction. So we conclude that for the stability of a floating body, the
metacentre should lie above the centroid of the body. The metacentric height depends on the
shape and the size of the floating body.

Fig. 1.4.3 Oscillation of a floating body


We know that when a body is displaced from its stable equilibrium position, it returns back to its
initial position after oscillating for a couple of times about its mean position. We would wish to
bring the body back to its equilibrium position with minimum oscillations. Normally the floating
body of interest is a ship or a boat and therefore we will explain the concept with reference to it
though it is applicable to any floating body. Figure 1.4.3 (a) shows the plan and elevation, taken
at the section AA, of a ship and the name of its different portions. The vessel can oscillate about
the fore-aft axis, called rolling, or about the horizontal axis HH, called pitching. The vessel can
also oscillate about the third axis, perpendicular to the plane of the board, which is called
yawing. Moreover, the vessel contains the rotating parts whose axis turns with such motion and
in turn Coriolis force acts which changes one kind of oscillation to another kind of oscillation.
Pitching occurs when the wave front hits the fore or aft of the ship. The Coriolis component will
change it to rolling or yawing. However, the rolling motion is generally considered in the study
of oscillations of the vessel because moment of inertia about the fore-aft axis is minimum due to
large length and height as compared to the width of the ship.

Figure 1.4.3 (b) shows the rolled position of the vessel at any time t. In this position, portion
OP′Q′ of the ship goes into and OPQ comes out of water. As a result, the new centre of
buoyancy is at B1 where force of buoyancy F = W acts, where W is the weight of the ship. The
line BG meets the vertical line through B1 at M at an angle θ. The restoring couple due to the
weight W of the body is W × GD = W.GM sin θ ≈ W.GM θ, since θ is small. Therefore the Eq. of
motion by the de Alembert’s principle is

or,

θ = A sin αt + B cos α t, where

Here k is the radius of gyration of the water line section7 about the fore-aft axis. If T is the
periodic time of the oscillations of the body then we can use the conditions θ = 0 at t = 0 and t =
T/2 to determine the constants A and B. This gives

B = 0 and A sin (αT/2) = 0

But
A ≠ 0, ∴ αT/2 = π

The periodic time can be reduced by increasing the metacentric height or reducing the radius of
gyration. We would like that vibration of the ship is dampened quickly and periodic time of
vibration be small so that passengers are not disturbed.

1.4.1 Determination of Metacentric Height

Metacentric height plays a vital role in stability of the floating bodies and its oscillation
frequency; its value depends on the distribution of the material, shape and the size of the floating
body. Therefore we would like to determine it in terms of geometrical shape and dimensions of
the floating body so that it may be used as useful information while designing such bodies, like
ship, boat, etc. Figure 1.4.3 (a) shows two views of the body. The top view is the plan of the deck
and the bottom one shows the elevation with the water surface at PP, which is the plane of
floatation. Let the body rolls clockwise by an angle θ, as shown enlarged in Fig. (b). As a result,
the portion OPQ comes out of the liquid whereas the portion OP′Q′ goes in and displaces more
liquid. As such the point of application of the buoyant force is displaced to the point B1, to the
right. The original water surface line PP takes the position QQ′ intersecting the new water line
PP′ at the point O. we take O as the origin of the coordinates with OY and OZ as the axes, as
shown in the Fig.; OZ is vertically down. Further the total volume of the displaced liquid will not
change during the rolling of the body and therefore the two areas OPQ and OP′Q′ will be equal.

To determine the position of the centre of buoyancy, consider an area element δA, as shown in
the plan, in the plane of the floatation at a distance y from the origin where the height of the body
is z. This displaces volume of liquid equal to (δAz). Taking moments about the X-Z plane
passing through O gives the y-position of the centre of gravity of the water displaced. In the
original undisturbed position of the body, we get

In the displaced position, z has increased by (y tanθ) as a result of the portion of the body going
into the liquid for the element shown in the Fig.. In this case, the new position of centre of
buoyancy is

Shift BB1 is then obtained by taking difference of the two position of the centre of buoyancy to
give
But BB1 = BM tan θ and ∫ y2δA = Ixx, the second moment of area about the fore-aft axis in the
water line surface.

Thus, as expected, the metacentric height depends on the second moment of area, i.e. the
distribution and the size of the floating body. We have so far not considered the loading of the
ship that will affect the second moment of area and therefore plays a vital role in the stability of
the vessel. The problem is more aggravated when it carries liquids, as in tankers, in which case
the free surface of the liquid in its container changes depending on the acceleration of the ship.
Thus, all these factors contribute towards the stability and proper care needs to be taken to ensure
stability of the ship in all the conditions not only for the rolling but for the other types of ship
motion such as pitching or yawing too.

Experimental Determination of Metacentric Height

Fig. 1.4.4 Experimental determination of metacentric height

To determine the metacentric height, a simple experiment is performed in which a weight P is


placed on the floating body of weight W such that forces P, W and buoyancy force F are in
vertical line. Then the weight P is moved to one side by a distance X that heels the floating body.
The body acquires the equilibrium position since the centre of gravity of P and W also shifts to
the new point G′ such that once again the point of buoyancy and the gravity lie vertically in line.
In this new position, the vertical line will intersect the original line BG at the point M in the
water line. This point is the metacentre. Let the angle of heel of the body is θ, which can be
measured through experiment. The shift in G is due to movement of P, we can write
P X = W × GG′ = W GM tan θ

The metacentric height can thus be determined.

Example 1.4.1 A cube of side a and relative density S (S < 1) floats in water. Determine the
conditions for its stability if it is given an angular tilt.

Fig. E 1.4.1 Stability of a cube

Solution Let x be the height of the cube immersed in water. By Archimedes principle,

a3 S w = a2 × w × x

Let G and B represent the centre of gravity of the cube and the displaced water; their height from
the bottom of the cube is a/2 and x/2, respectively. Therefore, BG = (a – x)/2. Using Eq. 1.4.5 to
get BM, we get

BM = I/V = (a4/12)/(a2 x) = a2/(12x) and

GM = BM – BG = a2/(12 x) – ½ (a – x) > 0 for stability

or,
or,

or,

S2 – S + 1/6 > 0

or,

(S – 0.789) (S – 0.211) > 0.

We get the factors 0.789 and 0.211 taking equality sign. This condition is satisfied if S > 0.789 or
< 0.211; i.e. the specific gravity of the cube should be less than 0.211 or greater than 0.789. If the
specific gravity is within these limits then the cube will be unstable. In case the cube has any of
these two values then it is in neutral equilibrium.

Example 1.4.2 A cylinder of 1.5 m diameter and 2 m long floats in seawater with its axis
vertical. The base of the cylinder is 1.5 m below the water surface. Find the weight of the
cylinder and the position of gravity if it is 0.3 m below the metacentre. Seawater weighs 10055
N/m3.

Solution Weight of the cylinder = weight of the seawater displaced

= (π/4) × 1.52 × 1.5 × 10055 = 26653 N

Fig. E 1.4.2 Weight of cylinder

Figure E 1.4.2 shows the position of the cylinder with O, G, B, M representing the centre of the
base, centre of gravity, centre of buoyancy and the metacentre, respectively.
OB = 1.5/2 = 0.75 m and BM = 0.09 m.

But OM = OB + BM = 0.84 m and also = OG + MG

→ OG = 0.54 m, since MG = 3 m (given).

Example 1.4.3 A barge in the shape of a rectangular block which is 8 m wide, 12.8 m long and 3
m deep floats in water with a draft of 1.8 m. The centre of gravity of the barge is 0.3 m above the
water surface. State whether the barge is in stable equilibrium. Calculate the righting moment
when the barge heels by 10°.

Fig. E 1.4.3 Righting moment and stability of the range

Solution Centre of buoyancy will lie below the free surface at a distance = 1.8/2 = 0.9 m.

Therefore, BG = 0.9 + 0.3 = 1.2 m. Further, BM = = 2.963 m and

GM = 1.763 m. We observe that BM > BG and hence the barge is in stable equilibrium. The
righting moment for 10° heel is given by

Fb × GM sin 10, where Fb = 9.81 × 8 × 12.8 × 1.8 = 1808 kN

righting moment = 553.56 kNm.

Example 1.4.4 A cylindrical buoy 2 m in diameter and 1.5 m in height weighs 12 kN and floats
in saltwater of density 1020 kg/m3. Centre of gravity of buoy is 0.65 m above O. If a load of 2
kN is placed on its top symmetrically, find the maximum height of the centre of gravity of this
load above the bottom if the buoy is to remain in stable equilibrium.

Fig. E 1.4.4 Weight placed over a buoy

Solution Let G1, G2 and G be the centre of gravity of the buoy, the additional load and of the
combination, respectively and h be the maximum depth of immersion of the buoy in water under
the extreme limiting condition of stable equilibrium. This occurs when the metacentre coincides
with the point G when the combination is in neutral equilibrium. By Archimedes principle

× h × 1020 × 9.806 = (12 + 2) × 1000

h = 0.445 m above the bottom

Buoyancy force acts at h/2 = 0.223 m above the bottom of the buoy. Also

and

OG = OB + BG = 0.785 m.

Taking moments about the base, we get

2 × h2 + 12 × 0.65 = 14 × OG


h2 = 1.595 m.

Example 1.4.5 A ship displaces 49250 kN of seawater of density 1025 kg/m3. The second
moment of area of the water line section about the axis through the centre is 120000 m4 and the
centre of buoyancy is 2 m below the centre of gravity. If the radius of gyration is 3.7m, calculate
the period of oscillation.

Solution The metacentric height is obtained from Eq. 1.4.5 and the time of oscillation from Eq.
1.4.1.

Example 1.4.6 An 80 mm diameter composite cylinder consists of an 80 mm diameter 20 mm


thick metallic plate having specific gravity 4.0 attached at the bottom of an 80 mm diameter
wooden cylinder of specific gravity 0.8. Find the limits of the length of the wooden portion so
that the composite cylinder floats in stable equilibrium in water with its axis vertical.

Fig. E 1.4.6 Stability of cylinder

Solution The minimum height of the wooden portion is decided so that the composite cylinder
should float for which the weight of the cylinder must be less than or equal to the weight of the
liquid displaced. Let L is the height of the wooden portion. Then

L ≥ 0.3 m.
Thus the wooden portion should be more than 0.3 m so that the cylinder floats in water. This will
remain in stable equilibrium condition till the metacentric height is positive, which condition
determines the maximum length of the wooden portion. We determine position of G by taking
moments of weights about the point O to give

Since the cylinder is floating, its weight should be equal to the weight of the displaced liquid.
This gives another equation

BM = after substituting for h from Equation B.

∴ GM = BM – BG =

or,

or,

L2 – 0.6L – 0.035 < 0

or,

(L – 0.653) (L + 0.053) < 0.

This condition is satisfied only when L < 0.653 m. Thus the limits of the wooden portion are 0.35
m and 0.653 m.

1.5 Pressure Variation in Liquids Confined in Moving Container


Tankers are used for shipment of liquid in sea or road transports where the carrier is sometimes
accelerated, decelerated or moves with constant acceleration. In such case the free surface of the
liquid in the tanker changes and acquires equilibrium state. As discussed in last section, this
affects the stability of the vessel. Therefore, we would now discuss how to determine the free
surface and its effects on the pressure variation. The other kind of container motion of
importance is its rotation about any vertical axis passing through the liquid. The common feature
in these kind of motions P is that laws of fluid statics are applicable.

1.5.1 Container Moving with Constant Acceleration

Figure 1.5.1 shows a vessel containing liquid moving with constant acceleration α having
components ay and az along the y and z axes with the origin in the free surface. Because of the
inertia of the liquid and its property to move relative to the surface, the free surface will shift and
would take inclined position at an angle θ with the horizontal direction. This slope will depend
on the magnitude and direction of the acceleration. We know the effect of inertia is to oppose the
main motion as a result when the vessel accelerates, the liquid will go back resulting in rising of
level on the rear end and lowering on the front end of the vessel, as shown in the figure. For
constant acceleration the free surface takes the inclined position and stays there. The pressure at
any point can be obtained by treating the fluid at rest. That is, the vessel moves but not the fluid,
which remains static with respect to the vessel; effect of the vessel movement is restricted to
change in slope of the free surface. For this reason this subject is studied with fluid statics. If the
vessel moves with no acceleration then the free surface will remain horizontal.

Fig. 1.5.1 Liquid in a container subjected to a constant acceleration

We may determine the angle θ by considering a fluid element ABCD whose enlarged view is
shown in Fig. 1.5.1. Let p is the pressure at the centre of the element considered whose sides are
δy and δz. Because the fluid is continuous, pressure at the faces AB and CD can be written using
the Taylor series while dropping the higher order terms; the pressures are shown in the Fig. for y-
direction and similarly we may write along the z-direction. Using Newton’s second law of
motion, we can write the following equation of motions along y and z directions for the element
taking unit width in x-direction.

We can integrate Eq. 1.5.1a to give p at any depth.

p = –ραy y + f(z)

= f′(z) = –ρ(αz + g) (from second Eq.).

Integration with respect to z gives f. Thus

f = –ρ(αz + g)z + p0,

where p0 is the constant of integration, the pressure at the free surface. This gives

Equation of the free surface is obtained by letting p = p0 in Eq. 1.5.2 to give

The negative sign shows that the slope of the surface is negative in case the vessel is
accelerating, as expected from the logic of inertia. The shape of the vessel does not affect the
pressure distribution so long whole fluid is connected. If there are more than one free surfaces, as
in case of a U tube type vessel, all the free surfaces will have same slope, since the fluid is
connected throughout. We conclude from Eq. 1.5.3 that the free surface will remain horizontal if
ay is zero or there is no acceleration at all. If the vessel moves vertically with some acceleration
then again the free surface will remain horizontal. This means ay is responsible for slanting of the
free surface and in that case az also affects the slope or the slant of the free surface.

Example 1.5.1 A rectangular tank 3 m long, 2 m wide and 2 m deep contains water to a depth of
1.25 m. If it is accelerated horizontally at 3 m/s2 in the direction of its length, find (a) (i) the
inclination of the water surface with the horizontal, (ii) depth of water at the two ends, (iii) total
force on the sides at the two ends of the tank and (iv) net horizontal force on the fluid mass.

(b) What maximum acceleration is permissible in this case?

(c) How much water would spill out if the tank is accelerated at 8 m/s2?

Fig. E 1.5.1 Horizontal acceleration of a tank

Solution (a) Equation 1.5.3 gives the slope of the free surface; the angle θ is tan–1 (3/9.81) = 17°.

∴ AA′ = BB′ = (3/2) tan θ = 0.4587 m.

Depth of water at the rear end = 1.5 + 0.4587 = 1.7087 m and

Hydrostatic force on rear side of the tank = w A z = 9810 × (2 × 1.7087) × 1.7087/2 = 28.642 kN

Depth of water at the front end = 1.5 – 0.4587 = 0.7913 m

Hydrostatic force on front side of the tank = 9810 × (2 × 0.7913) × 0.7913 /2 = 6.143 kN.

∴ Net horizontal force = 28.642 – 6.143 = 22.499 kN, opposite to the direction of αy.

(b) The tank can be accelerated further till the water level at the rear side just touches the top of
the tank. (Further increase in acceleration will spill out the water from the tank.) For this
condition, the slope of the free surface is tan θ′ = (2 – 1.25)/1.5 = ½. This gives ay = g × tan θ′ =
4.905 m/s2.

(c) In this case tan θ ′′ = 8/9.81, and the free surface will be CD. As seen in the Fig., there is no
water on the front end. From the triangle OCD, OD = OC/tan θ” = 2.4525 m.

∴ Volume of water in the tank = ½ × OD × OC × width of tank = ½ × 2.4525 × 2 × 2 = 4.905 m3.


Volume of water originally present = 3 × 1.25 × 2 = 7.5 m3.

→ Volume of water spilled out from the tank = 7.5 – 4.905 = 2.595 m3.

Example 1.5.2 The U-tube shown in the figure is moved horizontally with an acceleration of g/3
m/s2. Find the difference of water level in its two limbs.

Fig. E1.5.2 Acceleration of U-tube

Solution Slope of the free surface in the U-tube is tan θ = (g/3)/g = 1/3. Also from the figure, the
slope is 2y/0.9, where y is the displacement of the liquid in the limb due to acceleration. Equating
the two, we get, y = 0.15 m. Therefore, difference of water level in the two limbs is 2 × y = 0.3
m.

Example 1.5.3 A tank shown in Fig. E1.5.3 contains oil of specific gravity 0.8. If it is given an
acceleration of 5 m/s2 along a 30° inclined plane in the upward direction, determine the slope of
the free surface and pressure at b.

Fig. E 1.5.3 Acceleration on inclined plane

Solution Components of the acceleration along the horizontal and vertical directions are

ay = 5 cos 30 = 4.33 and az = 5 sin 30 = 2.5 m/s2.

Slope of the water surface is then tan θ = ay/(az + g) = 0.3519


→ θ = 19.38°.

Depth at b = 2 – (3/2) tan θ = 1.472 m.

The pressure at b now can be determined using the equation for static fluid. However, there is
acceleration in the z-direction too that would add to the pressure. We may use Eq. 1.5.2 without
considering the first term in y as effect of ay has already been accounted while using the depth
calculated above. Moreover, while going down, z is negative. With these comments, we get

pb/w = (1 + az/g) × 1.472 = 1.847 m.

or,

pb = 1.847 × 9787 × 0.8 = 14.463 kN/m2 (gauge).

1.5.2 Container Subjected to a Constant Rotation

This is another example wherein the law of fluid statics can be used to determine pressure at any
depth. Figure 1.5.2 shows a vertical cylinder rotating at some axis, not necessarily the axis of
symmetry. Rotation causes centripetal acceleration that increases pressure along radially outward
direction in the same horizontal plane. This implies that the liquid’s free surface will no longer
remain horizontal but would take a shape such that its height increases as we move out, as is
shown in Fig. 1.5.2, with minimum depth at O, the point where the axis of rotation meets the free
surface. The fluid rotates as if it is a rigid body and such motion is called forced vortex motion.
We take the origin at the base of the cylinder where the axis of rotation meets and the radial and
z-directions as shown in the figure.
Fig. 1.5.2 Constant velocity rotation of liquid

Consider an element of radial length δr rotating with angular velocity ω and subtending angle δθ
at the axis of rotation; its two enlarged views are also shown with dimensions and the pressure
acting at various faces. We use Taylor series to get pressure at other faces. We may balance the
forces in the radial and axial directions to give the following equations:

For r – direction: p. rδθ.δz – . (r + δr). δθ.δz + 2p δr δz = –ρω2r.δ r.δθ.δz

∂p/∂r = ρω2r, and

For z – direction: p.rδθ.δr – . rδθ.δr – ρg(rδθ.δr.δz) = 0

We may integrate these equations, as explained in section 1.5.1, to give pressure at any point
within the fluid. We get
We have used the condition z = z0, p = p0 on the free surface at the point O, where z0 is the height
of the point O from the origin. If we let p = p0 in Eq. 1.5.5 then we get the equation for the free
surface as

which is the equation of a paraboloid of revolution with vertex at O. If the axis of rotation
coincides with the axis of the cylinder then the paraboloid will be symmetrical about its axis. If
we substitute z = zR at r = R then Eq. 1.5.6 can be written as

Finally recalling that the volume of a paraboloid of revolution is equal to half that of the
circumscribing cylinder. i.e;

volume of paraboloid of revolution = 1/2 πR2 (zR – z0).

When the cylinder is not rotating then the free surface is horizontal. If z1 is the height from the
origin then volume of this additional liquid shall be equal to the volume of paraboloid of
revolution. This gives

πR2 (z1 – z0) = πR2 (zr – z0)/2

z1 = (zR + z0)/2

or,

z0 = 2z1 – zR.

We may express Eq. 1.5.7 in terms of z1

or,
It is observed that the rise near the surface is equal to the fall in level at the centre from the
original level. If we increase the angular velocity then the water may spill out from the cylinder
brim if its height is not sufficient. Slope of the free surface at any radius r can be obtained by
differentiating Eq. 1.5.6. with respect to r. This gives

Fig. 1.5.3 Pressure variation along stream line and radial direction

In fact, forced vortex flow is one case of a more general flow along the curved streamline in
which the velocity does not change on the same streamline and there exists radial acceleration
only. However, in any flow along a curved stream line, the velocity may also change along the
streamline. Figure 1.5.3 shows a streamline with r as radius of curvature having the centre of
curvature at O. To determine variation of pressure along the streamline, consider a cylindrical
element of cross sectional area da along it. Let v be the velocity then v2/r and v dv/ds will be the
radial and the stream wise acceleration respectively at the point under consideration. Equilibrium
of the element demands

Resultant force on the element along the streamline = mass × acceleration

or, pda – (p + dp) da = (w/g) da × ds × v dv/ds → dp/dv = –(w/g) v,

Where w is the specific weight of the fluid. Integrating this equation in the same horizontal
plane, we get

We shall learn in the following chapters that Eq. 1.5.10 is the Bernoulli’s equation and shows
that the sum of the pressure head and kinetic head remains constant along the streamline. On the
other hand, if we take the element along the radial direction, as shown in Fig. 1.5.3 (b), then the
radial equilibrium leads to
If the relation between v and r is known then Eq. 1.5.11 can be integrated. Substituting v = wr in
Eq. 1.5.11 and carrying out integration from r1 to r2, we get

You may get this result from Eq. 1.5.5 as well. This shows that both pressure and velocity
increases with increase in radius and the change in pressure head is equal to the change in the
velocity head in the same horizontal plane. Let us determine pressure at the point P at a depth z
from the free surface, as shown in Fig. 1.5.3 (c), where point P lies in the horizontal plane
passing through the point O. Pressure at the point P from hydrostatics is pa + w z whereas from
Eq. 1.5.12, it is pa + w ω2 r2/(2g), because at O the pressure is atmospheric. Equating pressure at
the point P obtained in two different ways, we get

pa + w ω2 r2/(2g) = pa + w z

z = ω2 r2/(2g)

which is Eq. 1.5.7, as expected. The origin has been taken at the free surface for which z0 = 0.

Example 1.5.4 A closed cylindrical vessel, 0.3 m diameter and 0.6 m high, two-third filled with
oil of specific gravity 0.8 is rotated about its vertical axis. Determine the speed of rotation when
(a) the oil starts touching the lid and (b) the point at the centre of the base is just clear of oil.
How the results will change if the cylinder is open at the top?

Solution Given for case (a): z1 = 0.6 × 2/3 = 0.4 m, zR = 0.6 m (see Fig. E1.5.4 a), and R = 0.15
m.

Using Eq. 1.5.8, we get ω = 18.68 rad/s. The liquid free surface is shown in Fig. E1.5.1.
Fig. E 1.5.4 Rotation of a cylinder

(b) In this case the liquid will collect more and more near the wall as it cannot spill out. Let R be
the radius at the cylinder brim when the liquid has touched the bottom of the cylinder. Initially,
1/3rd cylinder is not filled with liquid. Therefore, volume of the paraboloid of revolution should
be equal to it. This gives

Volume of paraboloid of revolution = volume of unfilled space initially.

or,

1/2 (πR2 × 0.6) = 1/3 {π(0.3/2)2 × 0.6

R = 0.122 m.

Now

zR = 0.6 m, z0 = 0 m and R = 0.122 m,

Eq. 1.5.8 gives ω = 28.1 rad/s.

If the cylinder is open then liquid will spill out as ω increases beyond 18.68 rad/s. When the
liquid touches the bottom of the cylinder then z0 = 0 m, zR = 0.6 m and R = 0.15 m. Eq. 1.5.8
gives ω = 22.9 rad/s. Further, the volume of the oil left in the cylinder = ½ A × 0.6 = 0.3 A, here
A is the cross sectional area of the cylinder. In the initial condition, the volume occupied is A ×
0.6 × 2/3 = 0.4 A.

Therefore, percentage of oil left in the cylinder = 0.3 A/(0.4A) × 100 = 75%. Thus the effect of
removing the lid from the cylinder top is that the oil remaining in the cylinder is decreased by
25% at the time liquid touches cylinder bottom.

Example 1.5.5 A vertical cylindrical tank 0.5 m diameter and 1.5 m high is closed at the top. It
contains water to a depth of 1.2 m, the air above the water surface is at a pressure of 39.24 kPa.
Determine the pressure head on the bottom of the tank at the axis and the side, when the tank is
rotated at 200 rpm about its axis.
Fig. E 1.5.5 Pressure at the base

Solution When the tank is rotated, the liquid free surface takes parabolic shape ABC, as shown in
Fig. E1.5.5. Let r is the radius of the point A from the cylinder axis and the origin be at B.
Volume of the air above the water level before rotation remains same. The volume after and
before the rotation are: ½ π r2 × z and π (0.5)2 × 0.3, where z is the vertical height of paraboloid
of rotation and the space above the water surface before rotation is 0.3 m (given). Equating these
two, gives

From Eq. 1.5.7, the equation of the paraboloid of revolution, we get, noting that in this case z0 =
0,

Solving Eqs. A and B, we get z = 0.917 m and r = 0.202 m. If O is the mid point of the cylinder
base then OB = 1.5 – 0.917 = 0.583 m. Further, pressure head at point B due to air pressure
(using Eq. p = ρgH) is = 39.24 × 1000/9810 = 4 m of water. Pressure head at point D, lying on
the cylindrical surface in the horizontal plane passing through the point B, due to centrifugal
force can now be obtained using Eq. 1.5.12. This gives


Now pressure head at O and E can be determined by using Eq. 1.3.2.

Pressure head at O = pressure head at B + 0.583 = 4 + 0.583 = 4.583 m.

Pressure head at E = pressure head at D + 0.583 = 5.397+ 0.583 = 5.980 m.

Summary

When the fluid is stationary or moving with constant velocity then the pressure variation with
depth of the liquid is given by –ρgdz, where ρ is density of the fluid, g is the acceleration due to
gravity and dz is the depth of the point. The negative sign appears because we take z-direction
positive in the upward direction and g acts in the –ve direction. Integration of the equation from
one point to another is possible if the relation between density of the fluid and the height is
known. We have considered incompressible fluid and isothermal conditions, linear variation of
temperature with altitude or isotropic conditions for compressible fluid to evaluate the pressure
variation. Fluid statics leads to the development of manometers used for pressure measurement.
The procedure for writing the manometric equation is discussed.

We have then examined the force acting on the plane and the curved surfaces dipped fully or
partially in the liquid. The important conclusion is that the surface area can be resolved along the
vertical and the horizontal planes to find out the horizontal and the vertical components of the
resultant force acting on the surface. The vertical component turns out to be the weight of the
liquid displaced by the immersed surface. The centre of pressure is always below the centre of
gravity of the area that depends on the second moment of the area, inclination of the surface with
the free surface and the depth of the centre of gravity from the free surface. The depth of the
centre of pressure is given by Eq. 1.3.10. This study has helped in determining the forces and its
position on the hydraulic structures so that these can be designed properly. We have also studied
the stability consideration of the dam.

Floating bodies are of prime importance as ships and boats have been used for a long time for
transportation. Their stability is of concern to us and it depends on the position of the metacentre
vis-a-vis the centre of gravity. It is established that when the metacentre lies above centre of
gravity then the body is stable. The other requirement is that once the body is displaced angularly
because of some disturbance in the surrounding then it should come back to the original position
with minimum number of oscillations. In this respect, we learn that the rolling motion needs
attention and the moment of inertia about the fore-aft axis in the water line section should be
large.

PROBLEMS

Review Questions

1. Differentiate between (i) liquids and gases, (ii) real and ideal fluids, (iii) specific weight
and specific gravity, (iv) compressible and incompressible fluids.
2. Define the terms: density, specific gravity, specific volume, and vacuum pressure.
3. What is the difference between dynamic and kinematic viscosity? State their units in
different system of units as well as their dimensions.
4. Enunciate Newton’s law of viscosity. Explain its importance in fluid motion. What is the
effect of temperature on viscosity of water and gas and explain the difference in their
behaviour with temperature?
5. Cite examples where surface tension plays a prominent role.
6. 1 litre of crude oil weighs 9.6 N. Calculate its specific weight and specific gravity of the
oil.
7. Explain atmospheric pressure. What is the value of atmospheric pressure head in terms of
mercury column and in terms of water column? Explain how atmospheric pressure varies
with altitude?
8. Differentiate between: (i) absolute and gauge pressure, (ii) simple manometer and
differential manometer, and (iii) piezometer and pressure gauge.
9. What do you mean by vacuum pressure?
10. What is a manometer? How are they classified?
11. What are the common liquids used in manometers? What conditions should it satisfy
before we choose a manometric liquid?
12. Suggest manometric arrangements to measure a pressure difference of (i) 50 N/m2, 0
kN/m2, and (iii) 100 kN/m2 between the two sections of a pipe carrying a liquid? How
will the arrangement differ if the pipe carries air?
13. What considerations govern the diameter of the glass tube to be used in a manometer?
14. What will happen if two limbs of U-tube manometer are of different diameters?
15. What is the difference between U-tube differential manometers and inverted U-tube
differential manometers? Where are they used?
16. Describe briefly with sketches the various methods used for measuring pressure exerted
by fluids.
17. What do you understand by hydrostatic equation? With the help of this equation, derive
the expression for the total thrust on a submerged plane area and the buoyant force acting
on it.
18. The diameter of a ram and the plunger of a hydraulic press are 300 mm and 50 mm
respectively. find the load that can be lifted with a force of 400 N.
19. What do you understand by the term, ‘temperature lapse rate’? Obtain an expression for
the temperature Lapse rate. Derive an expression for the pressure at a height Z form the
sea level for a static air when the compression of the air is assumed isothermal. The
pressure and temperature at sea level are p0 and T0 respectively.
20. What is hydrostatic pressure distribution? Give two examples each where pressure
distribution is hydrostatic and the pressure distribution is non-hydrostatic.
21. What do you understand by the total pressure and centre of pressure?
22. Prove that the centre of pressure of a completely submerged plane surface is always
below the centre of gravity of the submerged surface or at the most coincides with it.
When will that happen?
23. Prove that the total pressure exerted by a static liquid on an inclined submerged plane
surface is the same as the force exerted on a vertical plane surface as long as the depth of
the centre of gravity of the surface is unaltered.
24. Explain how would you find the resultant pressure on a curved surface immersed in a
liquid?
25. Why are different techniques adopted to calculate the total force and its point of
application, in case of plane and curves immersed surfaces?
26. What are the conditions of equilibrium of a floating body and a submerged body?
27. Define the terms metacentre, centre of buoyancy, metacentric height, gauge pressure and
absolute pressure.
28. A water body is subjected to acceleration in the vertically upward direction. At what
acceleration will the pressure difference between two points, separated by a vertical
distance h, be zero?
29. What is the Archimedes principle? A body floats in between two fluids of specific
weights γ1 and γ2. What will be the expression for buoyancy force?

Problems for Practice

1. Determine the torque and power required to turn a 10 cm long, 5 cm diameter shaft at 500
revolutions per minute in a 5.1 cm diameter concentric bearing flooded with a lubricating
oil of viscosity 100 centipoise. [0.1029 Nm, 5.387 W]
2. A thrust bearing consists of a 10 cm diameter pad rotating on another pad separated by an
oil film, μ = 80 cP, by 1.5 mm. Compute the power dissipated in the bearing if it rotates
at 100 rpm. [0.0574 W]
3. A skater weighing 800 N skates at 54 km/hr on ice at 0°C. The average skating area
supporting him is 10 cm2 and the effective dynamic coefficient of friction between the
skates and the ice is 0.02. If there is actually a thin film of water between the skate and
the ice, determine its average thickness. Assume water has dynamic viscosity at 0°C of 1
cP.
4. A U-tube is made up of two capillaries of diameters 1 mm and 1.5 mm respectively. The
U tube is kept vertically and partially filled with water of surface tension 0.0075 kg/m
and zero contact angle. Calculate the difference in the level of the menisci caused by the
capillarity. [0.68 mm]
5. Calculate the gauge pressure at A, B, C, and D in the Fig. in N/m2, kg/cm2, and m of
W.G.

Fig P 1.5. Pressure at different points

6. In order to measure the pressure in a pipe containing an oil of specific gravity 0.8, four
U-tubes are connected in series as shown in Fig P1.6. The gauging liquid in all the U-
tubes is mercury of specific gravity 13.6. Between the columns of mercury the liquid in
the U-tubes is water. Find the pressure at the centre of the pipe.
Fig P 1.6. Multi-tube manometer

7. In the arrangement shown in Fig. P 1.7, find the difference of pressure of water in the
pipes A and B.

Fig P 1.7. Differential manometer

8. Find the pressure difference between the containers A and B shown in Fig. P 1.8.
Fig P 1.8. Pressure difference

9. Determine the total pressure and the depth of the centre of pressure for the vertical lamina
shown in Fig. P 1.9.

Fig P 1.9. Pressure on the lamina

10. Figure P 1.10 shows a vertical semicircular lamina. Determine (i) the total pressure on the
lamina and (ii) the depth of the centre of pressure for the lamina.
Fig P 1.10. Pressure on the lamina

11. A square plate of size 2a is placed vertically with its upper edge horizontal and at a depth
a below the liquid surface. A square hole of size a is cut out of the plate as shown in Fig.
P 1.11. Find the total pressure on one face of the plate and the depth of the centre of
pressure.

Fig P 1.11. Pressure on the lamina

12. A rectangular gate shown in Fig. P 1.12 should tip automatically when the water rises
above the certain level. Determine that level in terms of h. [1.732 m]

Fig P 1.12. Height of water for gate to trip

13. A sluice gate, shown in Fig. P 1.13, is 2 m wide and 1.2 m high. It is hinged at the
bottom. On the upstream side there is seawater extending to a height of 1.6 m above the
top of the gate, and on the downstream side there is fresh water up to the top of the gate.
Find (i) the resultant pressure force acting on the gate, (ii) the position of the centre of
pressure, and (c) the least force acting at the top of the gate, which will open the gate.
Fresh water and sea water weigh 9.81 kN/m3 and 10.05 kN/m3, respectively.

Fig P 1.13. Force on sluice gate

14. The gate AB hinged at B, as shown in Fig. P 1.14, is 4.5 m wide and is connected by a rod
and pulley to a sphere of concrete. Find the diameter of the sphere so that the gate
remains just closed. Take specific gravity of the concrete = 2.4. [3.221 m]

Fig P 1.14. Force on gate

15. A slab AB 4 m × 2 m supports water as shown in Fig. P 1.15. The slab is hinged at the
bottom A. The middle point D of AB is connected to a prop DE. If the slab is at 60° with
the horizontal and the prop is at 45° with the slab, calculate the thrust in the prop. [64 kN]
Fig P 1.15. Thrust on prop

16. A tank ABCDE contains water up to a depth of 1 m and is 2 m wide. The curve AB is
defined by z = x2 and curve CD is a quadrant of a circle of radius 0.3 m. Calculate the
magnitude of the resultant forces on surface AB and CD. Find also the angle that the
resultant force makes with the horizontal. [AB: 16.29 kN, 52.96°; CD: 7 kN, 42°]

Fig P 1.16. Force on curved surfaces

17. Find the horizontal and vertical components of the force on the radial gate shown in Fig.
P 1.17. What force F is required to open the gate? Ignore the weight of the gate. The gate
is 2 m wide. [196. 2 kN, 218.6 kN, 0 kN]

Fig P 1.17. Force on radial gate


18. Find the horizontal and the vertical components of the hydrostatic forces on the
segmental gate shown in the Fig. P 1.18. The gate is 3m long. [42.376 kN, 12.1 kN]

Fig P 1.18. Force on segmental gate

19. Figure P 1.19 shows a semi-circular gate ABC retaining water on its two sides. The face
ABC is cylindrical while the face AC is a flat vertical surface. Calculate per metre run of
the gate, and the vertical and horizontal components of the force exerted by the water.
[15.41 kN, 14.715 kN]

Fig P 1.19. Force on semi-circular gate

20. A dam section is 8 m high, the maximum depth of water impounded being 7.5 m, the top
width of the section is 1 m, the weight of the masonry and water is 22000 and 9810 N/m3,
respectively. Find the minimum width required so that the dam does not slip under the
hydrostatic force. Take the coefficient of friction between masonry and masonry as 0.5.
The water face of the dam is vertical. [4.48 m]
Fig P 1.20. Slipping of dam

21. A cylinder of 1 m diameter and 2 m length stays in equilibrium as shown in Fig. P 1.21.
Calculate the specific gravity of the material of the cylinder. [797.3 kg/m3]

Fig P 1.21. Cylinder in equilibrium

22. A hollow cylinder open at both the ends has internal diameter of 30 cm, wall thickness of
15 cm and a length of 90 cm. If it weighs 625 N, find whether the cylinder would be
stable while floating in water with its axis vertical. [Unstable].
23. A wooden log of 0.6 m diameter and 5 m length is floating in river water. Find the depth
of the wooden log in the water when the specific gravity of the log is 0.7. [0.395 m.]
24. Find the density of a metallic body, which floats at the interface of mercury of specific
gravity 13.6 and water such that 40% of its volume is submerged in mercury. [6040
kg/m3.]
25. A block of wood of specific gravity 0.7 floats in water. Determine the metacentric height
of the block if its size is 2 m × 1 m × 0.8 m. [0.0288 m.]
26. A body has the cylindrical upper portion of 3 m diameter and 1.8 m deep. The lower
portion is a curved one, which displaces a volume of 0.5 m3 of water. The centre of
buoyancy of the curved portion is at a distance of 4.95 m below the top of the cylinder.
The centre of gravity of the whole body is 1.2 m below the top of the cylinder. The total
displacement of the water is 3.9 tonnes. Determine the metacentric height of the body.
[0.596 m].
27. A rectangular pontoon 10.0 m long, 7 m broad and 2.5 m deep weighs 686.7 kN. It
carries on its upper deck an empty boiler of 5 m diameter weighing 588.6 kN. The centre
of gravity of the boiler and the pontoon are at their respective centres along a vertical
line. Find the metacentric height. Weight density of seawater is 10.104 kN/m3. [0.12 m].
28. A solid cone floats in water with its apex downwards. Determine the least apex angle of
cone for stable equilibrium. The specific gravity of the material of the cone is given 0.8.
[31°.]
29. A rectangular tank is moving horizontally in the direction of its length with a constant
acceleration of 0.4 m/s2. The length, width and depth of the tank are 6 m, 2.5 m, and 2 m
respectively. If the depth of water in the tank is 1 m and tank is open at the top then
calculate: (i) the angle of the water surface to the horizontal, (ii) the maximum and
minimum pressure intensity at the bottom, and (iii) the total force due to water acting on
each end of the tank. [13° 44.6’, 17008.5 N/m2, 2611.4 N/m2, 868.95 N, 36861.8 N]
30. The rectangular of the above problem contains water to a depth of 1.5 m. Find the
horizontal acceleration which may be imparted to the tank in the direction of its length so
that (i) the spilling of water from the tank is just on the verge of taking place, (ii) the front
bottom corner of the tank is just exposed, and (iii) the bottom of the tank is exposed up to
its mid-point. [1.636 m/s2, 49050 N, 12262.5 N, 3.27 m/s2, 49050 N]
31. A rectangular tank of length 6 m, width 2.5 m, and height 2 m is completely filled with
water when at rest. The tank is open at the top and is subjected to a horizontal constant
linear acceleration of 2.4 m/s2 in the direction of its length. Find the volume of the water
spilled from the tank. [11.007 m3.]
32. A ship displaces 49250 kN of sea water of density 1025 kg/m3. The second moment of
inertia of the water line section about axis through the centre is 12000 m4 and the centre
of buoyancy is 2 m below the centre of gravity. If the radius of gyration is 3.7 m,
calculate the period of oscillation. [11.08 s]

Find answers on the fly, or master something new. Subscribe today. See pricing options.

Back to top

 Recommended
 Playlists
 History
 Topics
 Settings
 Support
 Get the App
 Sign Out

© 2019 Safari. Terms of Service / Privacy Policy




Skip to content

 Safari Home

 Resource Centers

 Playlists

 

 History

 Topics

 Learning Paths

 Offers & Deals

 Newsletters


 Highlights


o Settings
o Support
o Sign Out
o
o
o

Table of Contents for


Fluid Mechanics and Hydraulic Machines

 Search in book...

 Toggle Font Controls



o
o
o
o

Prev Previous Chapter


4. Flow Through Pipe and Channel

Next Next Chapter


6. Turbulent Flow

CHAPTER 5
Laminar Flow
OUTLINE

 Introduction–Types of solutions
 Exact Solution of N–S equation
 Low Reynolds Flow
 Boundary Layer
 Integral Solution for Boundary Layer
 Boundary Layer Separation and Control
 Flow over Curved Surfaces

5.1 Introduction—Types of Solutions

We discussed laminar flow in Chapter 2 and noticed that flow takes place as if the fluid is made
of layers or lamina and the interaction between these adjacent layers is through the transfer of
molecules, which results in the various transport properties, namely, the dynamic viscosity,
thermal conductivity, diffusion, etc. We also developed the mathematical model for this flow and
noted that the resulting equations are (i) non-linear, (ii) partial differential equation of second
order, (iii) elliptic, and (iv) set of simultaneous equations. For the solution of these equations, we
have four alternatives, namely

1. Exact solution of exact equation


2. Approximate solution of exact equation
3. Exact solution of approximate equation
4. approximate solution of approximate equation
The first kind of solution was possible in some simple cases, such as Couette, Hagan-Poiseuille
flows, flow through annulus, etc. Stokes flow and hydrodynamic lubrication are a few examples
from the second category of solution. Boundary layer flow fills the third category of the solution.
A large category of flows fall into the last category of solutions.

5.2 Exact Solution of N–S Equation

There are some simple two-dimensional flows in which the governing equations reduce to total
differential equations and their exact solution can be obtained. Flows between two parallel plates
in a pipe or between the rotating cylinders are a few to name as example. These flows can be
idealised as two-dimensional flows, since the flow in the pipe and annulus is axisymmetric while
in the channel two plates of infinite width in z direction form the passage and hence there is no
variation in z direction. While the flow in pipe and annulus is because of pressure difference
from inlet to outlet, flow through a channel may be due to pressure difference as well as moving
the plates relative to each other. We shall now consider the analysis procedure for these three
kinds of flows.

5.2.1 Couette Flow

Couette flow is a flow of the incompressible fluid passing through a channel, as shown in Fig.
5.2.1, because of both—the pressure difference along its length and the movement of the upper
plate with some velocity U while keeping the lower plate fixed. (If the lower plate also moves
with –U velocity then the flow developed is shear flow. If the flow is due to pressure gradient
only then it is called plane Poiseuille flow.) When the plate moves, the fluid layer in its contact
also moves with the same velocity that induces the next layers to move subsequently with some
velocity. This process goes on and ultimately all the layers in the flow region move along the
flow direction from zero velocity at lower plate (it is stationary) to the velocity of the upper plate
U. Had there been no pressure difference across the channel, we might expect the variation of the
velocity to be linear. Pressure gradient affects the velocity profile in the region. Let the
coordinate system be fixed at point O to the lower plate and the height of the channel passage be
h. We also assume that the plates are long enough along the flow direction so that the flow is
fully developed, which means the velocity profile has stabilised and does not change with x
direction. We will now investigate the fully developed flow. Equation 2.4.15 governs this flow
with the boundary conditions: y = 0, u = v = 0 and y = h, u = U, v = 0.
Fig. 5.2.1 Couette or channel flow

For a fully developed flow, x variation of u-component of velocity is zero, that is, ∂u/∂x = 0. The
continuity equation then gives ∂v/∂y = 0. Its integration with respect to y gives v = c. From the
boundary condition at y = o, v = 0, we get constant c = 0. Therefore, v = 0 everywhere in the
fully developed flow, which means its first and higher order derivatives with respect to x or y
will be zero. Now consider the momentum equations, which give after dropping the terms having
v or its derivatives and x-derivatives of u

X–momentum Eq.:

Y–momentum Eq.:

The second equation implies that p is not a function of y and hence ∂p/∂x can be written as dp/dx.
Therefore N–S equation is reduced to

Boundary conditions: u = 0 at y = 0 and u = U at y = h

In the last equation, the partial derivatives have been replaced with the total derivative because u
is function of y alone. We get an ordinary differential equation of second order in u for which
enough number of boundary conditions are given to give an unique solution. So there is unique
solution to these equations. Integration of Eq. 5.2.1 two times with respect to y and using the
boundary condition gives

The constant of integrations a and b were obtained from the boundary conditions. Rest is just
rearranging the terms and putting them in non-dimensional form. Equation 5.2.2 is parametric
with B and hence gives a family of curves for the velocity distribution. The velocity distribution
is shown in Fig. 5.2.1(b). We note that when B = 0, the variation is linear, as mentioned earlier.
For a decreasing pressure (negative pressure gradient dp/dx) along the flow, B > 0 and we detect
that the velocity of the intermediate layers is more. For a large pressure gradient, the velocity of
a few layers close to the upper plate may be more than that of the plate itself. On the other hand,
if adverse pressure gradient exists (pressure increasing along the flow) then B < 0 and velocity of
the layers is less. At some pressure gradient, the velocity profile will have zero slope at the lower
plate and beyond this limit, the layers close to the lower plate move in the opposite direction of
the upper plate.
Let L be the length of the plate in the flow direction. The discharge per unit width of the plates in
x direction, for the case when the upper plate is at rest and the flow is due to pressure gradient
only, is

or,

Equation 5.2.1 can also be obtained by considering a fluid element of length dx and height dy in
the flow region at a distance y from the lower plate, as shown in Fig. 5.2.1 (a). The layers above
it try to pull it in forward direction while the lower one pulls it back with the result the shear
forces act on it in the direction shown in figure, where τ = μ du/dy. Also let p1 and p2 are the
pressure at the two ends of the element. Making balance of forces acting on this element, we get

(p1 – p2) dy = dτ dx or, dp / dx = dτ / dy = μ d2u / dy2

This is the same equation. The two terms refer to the pressure and viscous forces in the NS
equation. We assume absence of the body force, but then what happens to the inertia force is not
known in this kind of analysis.

Example 5.2.1 For steady flow between parallel plates, h = 2.0 mm, μ = 0.05 kg/ms, and dp/dx =
–10 kN/m3. Determine q, maximum shear and maximum velocity.

Solution Equation 5.2.2 gives du/dy on differentiation with respect to y from which we get

Maximum shear occurs at the stationary wall, which is at y = 0.

Therefore,

τmax = –(h/2) dp/dx = 10 N/m2.

Average velocity,
U= = 0.0667 m/s and q = Uh = 0.1334 × 10–3 per m3

Maximum velocity occurs at the centre of the channel, which is at y = h/2. Substituting in Eq.
5.2.2, we get

Example 5.2.2 Water at 60°C flows between two large plates. The lower plate moves to the left
at a speed of 0.3 m/s. The plate spacing is 3 mm and the flow is laminar. Determine the pressure
gradient required to produce zero net flow at a cross section. Take μ = 4.7 × 10–4 Ns/m2 at 60°C.

Solution This is the Couette flow with the difference that the lower plate is moving instead of
the upper plate. Without going into the arguments (you should be able to justify it), we write the
governing momentum equation as μ d2u/dy2 = dp/dx with the boundary conditions y = 0, u = –U
and y = h, u = 0. This gives the solution

and

We are given μ = 4.7 × 10–4 Ns/m2, h = 0.003 m, and U = 0.3 m/s. For Q = 0, we get from
Equation (A)

A negative pressure gradient indicates that the pressure must decrease in the direction of the
plate from the left.

Example 5.2.3 A continuous belt passing upward through a chemical bath at velocity U0, picks
up a liquid film of thickness h, density ρ and viscosity μ. Gravity tends to make the liquid drain
down, but the movement of the belt keep the fluid from running off completely. Assume that the
flow is fully developed and that the atmosphere exerts no shear force at the interface of the film.
State clearly the boundary condition to be satisfied by velocity at y = 0 and y = h. Obtain an
expression for the velocity profile.
Fig. E 5.2.3 Belt passing through a barth

Solution The movement of the belt is shown in Fig. E 5.2.3 with the coordinate system used for
analysis. Under fully developed flow assumption, the velocity component v is zero everywhere
and from continuity equation again ∂u/∂x = 0. So we get inertia term to be zero and this leads to
same equation μ d2u/dy2 = dp/dx with the boundary conditions at y = 0, u = U0 and y = h, du/dy =
0 (no shear force at the free surface). Integration of the equation twice with respect to y and using
the boundary conditions gives

Example 5.2.4 Kerosene oil flows upwards through inclined plates at the rate of 2.0 l/s per meter
width. The distance between the plates is 10 mm and the inclination of plates is 20° with the
horizontal. Determine the difference of pressure between the two sections 10 m apart. Take ρ =
800 kg/m3 and μ = 2x = 10–3 kg/m s.

Solution Given: h = 10 × 10–3 m, ρ = 800 kg/m3, μ = 2 × 10–3 kg/ms, q = 2 × 10–3 m3/s. From this
data, we obtain

m/s and z2 – z1 = 10 sin 20 = 3.42 m.

Using Eq. 5.2.41, we get the difference of piezometric head at the two sections.

Example 5.2.5 Laminar flow takes place between the parallel plates 10 mm apart. The plates are
inclined at 45° with the horizontal. For oil of viscosity 0.9 kg/ms and mass density 1260 kg/m3,
the pressure at two points 1.0 m apart are 80 and 250 kPa when the upper plate moves at 2.0 m/s
velocity relative to the lower plate but in opposite direction to flow, see Fig. 5.2.5. Determine (a)
velocity distribution, (b) the maximum velocity and (c) shear stress on the opposite plate.

Fig. E 5.2.5 Flow through inclined paralled plates

Solution Given: μ = 0.9 kg/ms, ρ = 1260 kg/m3, U = –2.0 m/s, h = 0.01 m, difference of
piezometric head between the two given sections 1 and 2,

where

γ = 1.26 × 9.81 = 12.36 kN/m3

L = distance between two points = 1/sin 45 = 1.414 m.

Since h1 > h2, the flow is in downward direction for which ∂h/∂x = –14.76/1.414 = –10.438 and
∂p/∂x = γ∂h/∂x = –129 kN/m2. Substituting the values in Eq. 5.2.2 and simplifying, we get

u = 516.49 y – 71.65 y2.

Differentiating it with respect to y and equating to zero gives location of the layer from the
surface where the velocity is maximum. It gives, y = 3.6 mm = 3.6 × 10–3 m. Substituting it in the
above equation, we get the maximum velocity = 0.93 m/s.

Shear stress on the plate = = –824.837. N/m2.

5.2.2 Hagen–Poiseuille Flow

Hogen and Poiseuille analysed the flow through a circular pipe due to the pressure gradient in the
direction of the flow. Naturally, the cylindrical polar coordinate system would be the most
suitable for analysis of this flow. The flow through pipe would be axisymmetric, since there is no
variation in velocity or pressure along the circumferential direction. This means uθ = 0. We
further take the pipe axis as x-axis. Assuming the fluid to be incompressible and the flow to be
steady, you must be able to show from N–S equation or balancing the forces in a pipe flow2, as
shown in Fig. 5.2.2, that for such a flow

Fig. 5.2.2 Balance of forces in a pipe flow

The constant of integration has to be zero, since the velocity is finite at r = 0. Note the term A/r
→ ∞ as r → 0. A/r will be finite only if A = 0. One more integration gives

where c is constant of integration obtained from the boundary condition at r = R, u = 0. This


gives

Following the procedures of Section 5.2.1, we may determine the average velocity, maximum
velocity, or discharge, etc. The results are

and
If we plot the velocity at any cross section with r, we would observe that the velocity is
symmetrically distributed around the central axis; the velocity is maximum at the axis and zero
on the walls. The velocity distribution is paraboloid of revolution3.

Some power is required to maintain the flow and can be determined from the product of the force
F opposing the flow and the average flow velocity U. If pressure at two sections is p1 and p2 then
the force F = (p1 – p2) πR2.

Power = F × U = (p1 – p2) × πR2 × U = (p1 – p2) Q

The total drag force can also be determined and is given by

D = τw × (2πRL) = πμR2L (– dp/dx),

since

τw = μ(du/dy)y = 0 = – μ(du/dr)y = 0 = (– dp/dx) × R/2.

where y = R – r is the distance of the point from the pipe surface.

The drag or skin-friction coefficient therefore is

Darcy–Weisbach defined a friction factor by the relation , where τ0 is the shear


stress on the pipe wall and U is the average velocity of the flow through the pipe. Though they
defined this empirical relation for a turbulent flow yet it is equally applicable to the laminar flow.
Thus, f = 4 Cf. Or, f = 64 / Re, where Re is the Reynolds number based on diameter of the pipe
and average flow velocity.

For fully developed turbulent flow it is difficult to determine such a relation. Friction factor, as
found from carefully carried out experiments, depends on the Reynolds number and some
measure of the surface roughness, usually expressed as non-dimensional number made with the
hydraulic diameter. Since the surface roughness would vary, we use average surface roughness.

From experimental observations, we find that the flow in the pipe remains laminar up to Re =
2000 beyond which waviness starts developing in the flow and becomes complete turbulent at
some higher Reynolds number. However, if we start decreasing the Reynolds number and see the
limit at which the flow pattern resembles a laminar flow, it is found that this number is 4000.
Thus there exists two critical numbers, called the lower critical and the upper critical numbers.
We assume the flow to be laminar till Re < 2000. However, the flow has been assumed to remain
laminar even up to Re = 2300. That is, the upper limit of Reynolds number for a laminar flow in
a circular pipe is taken from 2000 to 2300.

We defined the momentum and energy correction factors in Chapter 2. We can determine these
using the velocity distribution profile given by Eq. 5.2.6. This is simple integration. The results
of integration, after using the expression for average velocity as well, are

α = 4/3 and β = 2.

Flow through porous media has drawn our attention for a long time. Percolation of water through
the pores in the ground, flow of water within the rock from one to another location within the
ground, filters used in water filtration plant, pharmaceutical and chemical industries are a few
examples of flow through the porous medium which are of immense importance for us. These
pores are small enough but matter a lot. The pores can be modelled as circular pipes through
which the flow takes place because of gravity or may even be due to pressure difference inside
the earth surface or in filters. The large number of such pipes vary in diameter and length and
need introduction of some constant to determine the volume flow rate per unit area of solid,
expressed in terms of average velocity = – (k / μ)Δp*.4 k is called the permeability and has the
dimensions of length squared. It is the average of the square of the pore diameters. The unit of
permeability is Darcy.

Example 5.2.6 Determine the optimum diameter of the pipe required to carry 100 l/s of crude oil
(ρ = 950 kg/m3, μ = 8 × 10–2 kg/ms) and still maintain the laminar flow. Also determine the
power required for its transport.

Solution Given Q = 100 × 10–3 = 0.1 m3/s, ρ = 950 kg/m3, μ = 8 × 10–2 kg/ms

For laminar flow, the Reynolds number for flow through a pipe is around 2000 to 2300; say we
take it 2100. This determines the optimal pipe diameter.

Using Eq. 5.2.8, we get the pressure drop in length L = 1 km = 1000 m.

and

P = Δp × Q = 0.1213 kW
Example 5.2.7 For steady laminar flow through a conical pipe, as shown in figure, determine an
expression for the pressure loss across it.

Fig. E 5.2.7 Flow through conical pipe

Solution Let D be the diameter of the pipe at a distance x from the entrance then

D = D2 + (D1 – D2) (L – x)/L

This gives, dD/dx = – (D1 – D2)/L

For a laminar flow in the pipe, pressure loss is given by Eq. 5.2.8. This gives on integration

Example 5.2.8 A round wire of radius R1 and length l is coated continuously by drawing it at a
uniform velocity U through a pipe of radius R2 containing the coating liquid. If liquid is
maintained at constant pressure everywhere, obtain an expression for power required to draw it.
Fig. E 5.2.8 Power required for coating a wire

Solution This process is used in coating PVC over the copper or aluminum wire, enameling of
the wire, etc. We need to calculate power required for pulling (drawing) the wire that will depend
on the shear force exerted by the surrounding liquid on the moving wire. We should use the
cylindrical polar coordinates and an axi-symmetric, steady laminar flow conditions around the
wire may be assumed. Therefore NS equations can be used with the boundary conditions: at r =
R1, u = U and r = R2, u = 0, where u is the velocity component in the direction z of movement of
the wire. We may further assume that there is no body force. Under these conditions,

∂/∂θ = ∂/∂t = ∂/∂z = 0, the NS equation would reduces to

On integration twice with respect to r and using the boundary conditions at r = R1, u = U and at r
= R2, u = 0, we get

and P = work done/s =

Example 5.2.9 When heated oil is pumped from the oil well to its destination through an non-
insulated pipe line, the temperature of oil Tx at any distance x from the inlet decreases along the
pipe length due to heat transfer to surrounding environment. As a first approximation assume
that the temperature variation can be expresses as Tx/T0 = exp (–kx/D) where T0 is temperature at
x = 0 and k is a constant. Further, dynamic viscosity of oil can be assumed to vary with
temperature as μ ~ T–N where N is a constant. Obtain an expression for pressure loss in length x if
p0, T0, μ0 represent the pressure, temperature and the dynamic viscosity at inlet for discharge Q
flowing in the pipe of diameter D.

Solution Given:Tx = T0 e–kx/D and → μx = μ0 ekNx/D

Loss of pressure in length dx will be

pressure drop Δp is given by its integration from x = 0 to x = x


5.2.3 Flow Between Rotating Cylinders

Let us consider some fluid present between two infinite long concentric cylinders and that any or
both of them be rotating in any direction with say angular velocities ω1 and ω2. In this case, the
radial and axial component of velocity will be zero; only the circumferential component of the
velocity will exist. We may argue as before and get the governing following equations from N–S
equations in cylindrical polar coordinate, with the cylinder axis as x-axis.

and

The first equation shows that the centrifugal force is caused because of the centripetal force
produced by increase in pressure with r. The second equation can be integrated with respect to r
two times to give the solution with two constants of integration, obtained from the boundary
conditions r = R1, uθ = ω1 R1 and r = R2, uθ = ω2 R2. This gives the circumferential velocity
distribution

Our interest is to find out the torque acting on the two cylinders as well. This torque is because of
the shear force acting on the annulus surfaces by the rotating fluid. Shear stress is related to shear
strain rate in cylindrical polar coordinates by the relation

If cylinder 1 is at rest (ω1 = 0) then the torque exerted on the second cylinder is equal to moment
of the shear force on it. Using relations from Eq. 5.2.10,
The same torque would act on cylinder 1 but in opposite direction. This also means that external
torque of this magnitude is to be applied to keep it fixed. This arrangement has been used to
determine viscosity of the fluid and was discussed in Chapter 1. F. M. White has discussed how
to handle the flow in the space between the two eccentric circular cylinders, called eccentric
annulus. The procedure is to use conformal transformation to change it into concentric annulus.
The transformation used for the purpose is

w = y + iz = M tan ς/2, where ς = ξ + iη, and F = (a2 – b2 + c2)/(2c)

Here a, b, and c respectively are the radii of outer and inner cylinders and the eccentricity in their
axes. Muskhelishvili proposed this transformation.

5.3 Low Reynolds Flow

It is possible to neglect inertia term in case of flows at low Reynolds number, since its magnitude
may be very small compared to the viscous force. This case belongs to the second kind of
solution—approximate solution of the exact equation. The resulting equations will then be linear
in nature and can be handled more comfortably. A few such cases are discussed below.

5.3.1 Stokes Flow

Stokes studied flow around a sphere of radius ‘a’ placed in a uniform flow having low velocity
U∞, for which the Reynolds number is around 1, such that the inertia terms can be neglected.
Only gravity is the body force and can be included with pressure term, represented by p* = p +
ρgz. N–S equation takes the form

Such a flow is axisymmetric and use of spherical coordinates because of the geometry of the
object is recommended. This equation requires two times integration and use of the given
boundary conditions to get the unique solution. Without going into the details of the solution
technique (that you should be able to do) the results are

is measured at far off distance in the undisturbed flow. Equation 5.3.3 reveals that the
pressure on the front side of the sphere is more as compared to its rear side pressure. Thus, the
pressure force acts along the direction of the flow. Viscous force also acts on the surface but in
the opposite direction. Both the pressure and the viscous forces can easily be determined. You
should try and verify that the viscous force is only 1/3rd of the total drag force on sphere. We
have learned how to calculate the pressure force on the surface. Similarly, the viscous force is
determined by determining the shear stress from Newton’s law of viscosity. The total force on
the sphere comes out to

This relationship is the Stokes law that holds true when the Reynolds number is less than 0.1. It
helped in experimental determination of the electric charge on an electron in Millikan’s oil drop
experiment and Boltzman constant using Einstein theory on Brownian movement. Stokes law is
also helpful in determining viscosity of fluid in laboratory. A small spherical steel ball is made to
fall down under gravity in a tank full of the liquid whose viscosity is to be determined. The
gravity moves the ball down which is opposed by the buoyant and the viscous forces. As the fall
velocity of the ball increases the later force increases; the net force on the ball is zero when the
ball attains constant velocity. The ball then attains the terminal velocity, which remains constant
during subsequent motion and is measured by recording the time for a known vertical fall. Under
these conditions, the equilibrium equation is

We will just make a comment that the drag force does not depend significantly on the shape and
orientation of the body in the flow. The analysis carried out on a circular disc of diameter d
shows that the drag force is 16 μa U∞ and (32/3)μa U∞ when placed perpendicular and parallel to
the flow direction, respectively. Comparison of their magnitudes confirms our comments.

Example 5.3.1 A viscometer consisting of a long vertical tank containing the fluid at the desired
ambient conditions and a small sphere is employed to determine the viscosity of oil. If a steel
sphere of specific gravity 8 and 1 cm diameter falls with a constant velocity of 0.05 m/s through
a lubricating oil of specific gravity 0.9, calculate its viscosity.

Solution We are given a = 0.5/100 = 0.005 m, ρs = 8000, ρf = 900 and U∞ = 0.05 m/s.
Substitution in Eq. 5.3.4. (a) gives μ = 7.74 Ns/m2. Let us also check the Reynolds number in this
case, which is

Value of Re < 1 justifies use of Eq. 5.3.4(a).

5.3.2 Hydrodynamic Lubrication

This is another flow that has considerable importance in industry and considers determination of
pressure and velocity field in the fluid confined between two non-parallel plates with one plate
fixed and the second moving. Moreover, the gap between the plates is much less as compared to
other plate dimensions and the inclination between the plates is also very small. Situation in
bearings, where the shaft (slipper) rotates in its bearing, can be approximated as fluid confined
between the two plates inclined at an angle α to each other having gap h1 and h2 at the ends, as
shown in Fig. 5.3.1. This is so because the radius of the shaft is much larger than the gap
between the shaft and the bearing.

Fig. 5.3.1 Modelling of a bearing

For analysis, let us take the coordinate system as shown in the figure. Then the gap h at any
distance x is h1 – (h1 – h2) x/L, where L is the length and h/L << 1. We assume the lower plate
moving with velocity U and the upper plate fixed. Let us have an estimate of the inertia and
viscous force to check if the inertia force may be neglected as compared to the viscous force.
This is done in the following manner

Thus the inertia force, though not completely absent as in Couette flow, is negligible in this case.
This difference separates this flow from the Couette flow and there is difference too in the
boundary conditions, which are:

y = 0, u = U, at y = h (x), u = 0, at x = 0, p = p0 and at x = L, p = p0.

To be exact, the pressure variation at the inlet and exit sections will not be constant and there
would be some variation; this variation however is ignored because of small magnitude of
variation owing to small film thickness.

We may also neglect the terms ∂2u/∂x2 as compared to the term ∂2u/∂ y2 because of its low
magnitude. Further, the v component of velocity and its derivatives will almost be non-existent,
the y-momentum equation is satisfied automatically. With these simplifications, the momentum
equation is reduced to
We infere that in this case too, pressure is not a function of y. Like before, we can integrate it
with respect to y and get the following results using the given boundary conditions

or,

Equation 5.3.7 has two unknowns namely B and Q, which are obtained using the conditions at x
= 0 and x = L. This gives

Equation 5.3.8 gives variation of p with x, which is shown in Fig. 5.3.1. The point of maximum
pressure can be determined using calculus. At the point of maximum pressure dp/dx = 0.
Therefore, Eq. 5.3.5 gives . Thus, at this section velocity varies linearly.

Because of this pressure, a force F acts in the upward direction on the upper plate that supports
it. The force F as well as the shear force D per unit width of the plate can be evaluated as follows

We note that h1 – h2 should be small in order that the supporting force may be large. The Michell
thrust bearing works on this principle of hydrodynamic lubrication. As said, this is an
idealisation, since the actual curves of the shaft will be curved.
Example 5.3.2 A journal bearing supports a circular shaft of diameter D = 10 cm turning at 3600
rpm. The bearing length L = 10 cm. The gap between the shaft and the bearing, h = 0.1 mm, is
filled with a lubricant whose viscosity is μ = 6.7 × 10–5 Pas. Calculate the torque applied to the
shaft to overcome the friction in the bearing and the power P consumed in the bearing by the
friction.

Solution Since the gap at the bearing end is not given, we shall assume it to be constant. In that
case the problem is changed to that of Couette flow with the shaft as moving plate of length 10
cm. The shear stress exerted on the shaft having radius R = D/2 is

τs = μU/h = μωR/h and T, torque on the shaft = τs (2πRL) × R

= πμωLD3/(4h) = = 1.994 × 10–2 Nm

P = Tω = 7.479 W

5.4 Boundary Layer

When the fluid flows over the surface then the layer of the fluid in contact with the surface
acquires the velocity of the surface, a condition called no-slip condition. The velocity of the
successive layers goes on increasing along normal to the surface, though the rate of increase
decreases as we move away from the surface. That is, the velocity plot has maximum slope at the
surface and it ultimately reduces to zero at some point away from the surface; theoretically this
point will be at infinity. But in reality variation is fast initially and after some distance, a few
mm, the velocity of the layers is almost same. In other words, the presence of the surface is felt
by the fluid layers up to certain distance normal to the surface and all the layers above it move as
if no surface is present.

Fig. 5.4.1 Boundary layer development over a flat plate

Figure 5.4.1 shows the experimental results of Prandtl obtained for a flow over a flat plate.
Velocity component u (along the flow direction) was measured at a section in the direction
perpendicular to the surface and was plotted at that section. Such measurements were made and
plotted at different sections of the plate; the figure shows these at a few sections. These plots
support the observations mentioned earlier. Such behaviour is explained recognising the fact that
the effect of viscosity is to oppose the motion. Thus the fluid layer velocity is decreased as we
approach the surface. We can also explain the growth of the boundary layer from the fact that the
slowness of layer decreases the mass flow rate and in order to satisfy the mass conservation, the
area of flow should increase and that’s why thickening of the layer along the flow direction.

Prandtl argued that the effect of viscosity is limited to some distance normal to the surface and
he divided the flow into two zones—the boundary layer close to the surface and potential flow
zone away from the surface. To demarcate the two zones, he took arbitrarily a point at a distance
such that the u component at this point is equal to 99 percent of the free stream velocity U and
called distance of this point from the surface as the boundary layer thickness at that section. Thus
he obtained a number of such points at different sections and joined these to give a demarking
line, as shown in the figure. We make the following observations:

1. Boundary layer grows along the flow direction. This is because the layers in contact with
the surface becomes slow and therefore more area is required to satisfy the mass
conservation.
2. The flow within the boundary layer is laminar up to certain distance at which section the
Reynolds number is about 3.2 × 105. At this point, the boundary layer becomes wavy and
remains so till the Reynolds number is about 5.5 × 105. Finally flow in the boundary layer
turns to turbulent flow. These two Reynolds number based on the distance x of the
section from the leading edge are known in literature as lower and upper critical numbers
and the flow between the two sections as transition region.
3. As we move down the plate, Reynolds number increases and the slope of the velocity
profile near the surface decreases; at some point it is zero. At this point the boundary
layer is ready to leave the surface and this point is called the point of separation.
Separation occurs because the fluid particle does not have enough kinetic energy to
overcome the frictional force. At the point of separation, ∂u/∂n = 0, where n represents
the normal to the surface; for a flat plate n is y.
4. Further downstream, there is another sub-layer close to the surface in which velocity is
negative. The fluid particles leave the main flow surface and circulate in this sub-layer
trying to go with the main flow getting energy from it.
5. Velocity profile in laminar boundary layer is parabolic whereas in turbulent boundary
layer it is logarithmic.
6. The boundary layer thickness varies with Reynolds number in the following manner For

laminar flow and for turbulent flow

5.4.1 Displacement, Momentum and Energy Thickness

We discussed formation of the boundary layer zone near the surface placed in the flowing fluid,
which has thickness that increases in the direction of the flow. At the end of the boundary layer
velocity of the layer is 0.99 U, where U is the mainstream velocity. Since the layers in the
boundary layer region move slow, there is loss of mass, momentum and energy flux rate. For that
reason, we think of an imaginary distance by which the plate may be moved normal to it in an
uniform flow field having velocity U such that the decrease in the these quantities (mass,
momentum, or energy) due to displacement of the plate is equivalent to that caused by the
decrease in velocity in the boundary layer flow.

Fig. 5.4.2 Displacement thickness

Let us consider a boundary layer flow over the flat plate of unit width along z-direction, which is
shown in Fig. 5.4.2, and a layer of thickness dy at a distance y from the surface. Flow velocity at
this layer is u whereas the fluid velocity in the potential flow region is U. Thus reduction in mass
flow rate through this strip due to loss in velocity is ρ(U – u) dy, equal to the area of the hatched
strip in the figure. The total loss of mass flow rate is then obtained by integrating this term from
y = 0 to ∞. This gives:

where δ* is the distance by which the plate has been shifted into the flow so that the resulting loss
in the mass flow rate is equivalent to that caused by the formation of the boundary layer. This
thickness is called displacement thickness for the obvious reason. We may now define similarly
the momentum and energy thickness, represented by δm and δe.5 Note that the loss in momentum
and energy corresponding to the loss in mass ρ (U – u) dy in the boundary layer is ρ (U – u) u dy
and ½ ρ (U – u) u2 dy, respectively whereas that by shifting the plate in uniform flow is ρU2 dy
and ½ ρU3 dy, respectively. Equating these and simplifying we get

Note: Though the integration is from 0 to ∞, the integral will be always zero when y > δ, since
the factor (1 – u/U) is negligible. For all practical purpose, the integration limit may be taken up
to δ.

These thickness are not arbitrary like the boundary layer thickness and are related to the velocity
distribution in boundary layer flow. Dependence of u/U on δ brings in dependence of various
displacement thicknesses on δ; a rough or approximate idea can be obtained from the following
table:
Type of boundary layer Displacement thickness Momentum thickness
Laminar δ/3 δ/8
Turbulent δ/8 7δ/72

We shall determine thickness for different velocity distributions while studying the solution of
the boundary layer equation.

5.4.2 Prandtl Boundary Layer Equations

Prandtl argued that the distances in the boundary layer region are very small compared to that in
the flow direction and that the velocities range from 0 to U. So he thought of comparing the
order of magnitude of various terms in the governing equations. He observed that relative
magnitude of many terms in these equations is negligibly small compared to other terms. This
simplified the equations and the resulting equations are known as boundary layer equations.

We shall restrict our discussions to a two-dimensional-steady flow of an incompressible fluid


over a plate just for simplicity to explain the procedure; in fact it is equally applicable for a three-
dimensional flow, compressible fluids, or the flow on a curved surface. We non-dimensionalise
the lengths x and y along x and y-direction with the length of the plate L, the velocity components
u and v by the potential flow velocity U and the pressure by the dynamic pressure ρU2, in order
to reduce the governing equations to a non-dimensional form. Many authors prefer to use
different symbols for the non-dimensional variables to avoid confusion but still others have
retained the same symbols; the form of equation itself speaks off. Let us rewrite the governing
equations in the non-dimensional form retaining the old symbols for non-dimensional variables:

Re is the Reynolds number = UL/ν whose appearance reflects that the equation is written in non-
dimensional form. For the boundary layer, x varies from 0 to 1 and we say it is of the order 1,
written as x ~ O (1). y << x while y varies from 0 to δ/L and we say y ~ O (δ). Similarly u ~ O
(1). By the order of the magnitude, we do not mean that all the terms in Eq. 5.4.3 necessarily will
have the maximum value at the same point. There is no reason for all of them to be maximum at
one point only. This raised a grave question about the validity of comparing the order of the
terms in the equations, since Eq. 5.4.3 is applicable at a point whereas its terms have maximum
magnitude, which are compared in order to delete small terms, at different points at a section.
This issue has been dealt with Meksyn in his book. Nonetheless, the closeness of the results
obtained from Prandtl boundary layer theory for different flows gives us confidence in his
arguments and we shall proceed to discuss his solution.
We may now get the order of various terms in the equations. Order of the derivatives of u are O
(∂u/∂x) ~ 1, O (∂u/∂y) ~ 1/δ, O (∂2u/∂x2) ~ 1, and O (∂2u/∂y2) ~ 1/δ2. For example, take ∂2u/∂y2.
Here u ~ O (1) and therefore the order of change in u is also 1 as it changes from 0 to 1.
Similarly the order of y is δ and so is the order of change in y as it changes from 0 to δ.
Therefore, O (∂2u/∂y2.) = 1/δ2. This way order of all the terms in the equation can be obtained.

Substituting the order in continuity equation we get, O (∂v/∂y) should also be 1 to satisfy this
equation since O (∂u/∂x) is 1. This means O (v) ~ δ, since y is of the order δ. Now we can
determine the order of the derivatives of v, these are: O (∂v/∂x) ~ δ, O (∂v/∂y) ~ 1, O (∂2v/∂x2) ~
δ, and O (∂2v/∂y2) ~ δ × (1/δ2) = 1/δ. Order of each term has been written below the term in Eq.
5.4.4.

X-momentum:

Y-momentum:

We observe from the x-momentum equation that

1. O (∂2u/∂y2) >> O (∂2u/∂x2), since ∂2u/∂y2 is of O (1/δ2) whereas ∂2u/∂x2 is of O (1).


Therefore ∂2u/∂x2 can be neglected from the viscous force terms.
2. The other terms in this equation are of the order 1. This implies that 1/Re should be of O
(δ2) so that viscous force term may be of O (1) so as to retain it in the equation. This
means the flow should have high Reynolds number.

Similarly, we may go ahead with the y-momentum equation and note that ∂2v/∂x2 may be
neglected compared to ∂2v/∂y2 and that ∂p/∂y should be of the order δ, since all the other terms in
the equation are of this order. This is an important conclusion, as it means the variation for the
pressure in the y-direction can be neglected and pressure at any section will be the pressure in the
potential flow zone, which can be determined using any of the methods discussed in Chapter 3.
In other words, the external pressure is impressed on the boundary layer. This also reduces one
unknown as p is now known and is a function of x alone. Lastly, the terms in y-momentum
equation is of the order δ whereas in x-momentum equation, these are of the order 1. Therefore
we may drop y-momentum equation even. This results, reverting back to the dimensional
coordinates, in

These equations are known as Prandtl boundary layer equations, which are parabolic type. These
equations however are not applicable in the region close to the leading edge where y is
comparable with x. Schmidt and Schroeder made the following objections to this simplified
derivation of the boundary layer equations. These are

1. ∂u/∂x is not of O (1) everywhere when u is increasing rapidly.


2. Even if the order of the terms u ∂u/∂x and v ∂u/∂x be one, does this mean that their sum
will also be of the same order 1.
3. In case of curved surfaces ∂us/∂n need not preserve the sign across the boundary layer.

They too gave a rigorous treatment to the boundary layer flow.

5.4.3 Solution of Boundary Layer Equations

We know the ordinary differential equations have many advantages over the partial differential
equations of the same order because the number of boundary conditions required to get unique
solution is equal to the order of the equation. For a partial differential equation, these are needed
at infinite points of the boundary of the surface enclosing the region, depending on the nature of
the equation. Moreover, the subject of solving the ODE has been exhausted to a great extent and
we have the solution for any equation. Therefore many researchers have made efforts to find out
some independent variable to combine two of them, namely x and y, into one. This possibility
saw the light because the velocity profiles in the boundary layer at different sections are similar.
This mean that similar points at different sections have the same non-dimensional u-component.
The velocity is made non-dimensional using the free stream velocity at that section whereas the
similar points at different sections are obtained using a scale factor g(x) for y-coordinate. This
way, it could be possible to describe a single velocity profile that describes the variation of
velocity across the boundary layer at all the sections in the laminar boundary layer. Schlichting
has discussed this problem in length and has obtained conditions for the type of flow that can
have similar solutions.

Blasius was studying flow over a flat plate. He used the principle of similarity, which demands
that the velocity at two points on any two arbitrary sections at x1 and x2 should satisfy the
equation

g(x1) and g(x2) are the scaling factors for y direction, which, in general, will be different. It is
possible to write


constant is taken as zero. We note from Eq. 5.4.7 that value of scaling factor g depends on x
value of the section, the fluid property ν and flow velocity u. We now differentiate Eq. 5.4.8 to
express different derivatives in the Prandtl b.l. equation as derivative of η. While differentiating
with respect to x, we note that ψ is the product of two functions and care must be taken. Equation
5.4.7 gives

respectively. The process yields

and

The number of primes over f represents the order of derivative with respect to η. Substituting
these in Eq. 5.4.5 and noting that for a flat plate, dp/dx = 0, after simplification we get

This is a third order non-linear ordinary differential equation. The boundary conditions are:

y = 0, u = v = 0 → at η = 0, f = f′ = 0.

At

The boundary conditions on f and its derivatives are obtained from Eq. 5.4.9 with the help of Eq.
5.4.13. For example, u = 0 means f′ = 0 and then v = 0 in association with u = 0 means f = 0 and
so on. This is a two-point boundary value problem, which is solved by matching technique. The
solution was given by Blasius and named after him. He obtained a series solution near η = 0 and
an asymptotic solution in the far field (η → ∞). The power series near η = 0 is assumed to be of
the form

This give the first, second and third derivatives on differentiation with respect to η. Substituting
these in Eq. 5.4.12, we get
This equation is satisfied if the coefficient of powers of η are zero. This gives:

A3 = A4 = A6 = A7 = 0, A5 = – 1/2 , A8 = – 11/4 , etc.

Therefore the series solution near η = 0 is:

In which C0 = 1, C1 = 1, C2, = 11, C3 = 375, C4 = 27897, C5 = 3,817,137, etc. We do not need


more terms in the series.

Now let us try to find out the asymptotic solution at η → ∞. The procedure is to assume an
asymptotic solution of the type

where f2 << f1, f3 << f2, and so on i.e. the successive approximations decrease very fast. We
determine functions f1, f2, etc in succession by first approximating the solution to f = f1 and then
to f = f1 + f2, and so on. Using the b.c. at η = ∞, we may write at a far off point from the surface
in the main flow.

β is a constant to be determined; we note from this equation that . Now include


second term f2 of the asymptotic solution so that f = (η –β) + f2 and substitute in Eq. 5.4.12. We
get

C is a constant of integration that has been chosen in the manner shown in the equation so that
we may rewrite this equation in the following form and then integrate the equation.

We note that the integral in Eq. 5.4.18 is zero when η = ∞. One more integration of Eq. 5.4.18
gives
We may improve the solution further taking the third term in the asymptotic expansion and
proceeding in the similar way. We leave it at this stage. Eq. 5.4.19 has two unknowns K and β
and the power series 5.4.16 has one unknown A2 that are determined by matching the power
series and the asymptotic solutions at some point η = η1 where both the solutions are applicable.
Attempts of Blasius to match the solution and to locate the point led to A2 = 0.332, β = 1.73 and
K = 0.231. (In literature α is used for A2.) Blasius solution in the boundary layer over a flat plate
is reproduced in the following table and is shown in Fig. 5.4.3.

Fig. 5.4.3 Distribution of the velocity functions across b.l.

TABLE 5.4.1 Blasius velocity profile

For engineers, total shear force is of importance. The local shear stress on the wall is
The non-dimensional shear stress at any section x is obtained by using a factor ½ ρU2. Also the
shear force at any section and the average shear force on the plate surface of unit width and
length L in the direction of flow can be determined. The procedure yields

These results match very well with the experimental results. Here ReL refer to Reynolds number
based on the plate length L. From Table 5.4.1, we observe that f′, which represents u/U, is
0.99155 at η = 5. That is, the velocity in the boundary layer is more than 99 percent of the free
stream velocity and η at the boundary layer edge is 5. Using expression for η in Eq. 5.4.7 the
boundary layer thickness is

We may also determine the displacement and momentum thickness using Eq. 5.4.1 and 5.4.2.

We note that the δ* ~ δ/3 and θ ~ δ/8 for a laminar boundary layer, which we mentioned earlier.

5.5 Integral Solution for Boundary Layer


Fig. 5.5.1 Integral solution

Other group of scientists were trying to find the integral solution for the boundary layer. The
procedure of integral solution technique was discussed in Chapter 2. Consider a C.V. as shown in
Fig. 5.5.1 in which the face BC is at a height h from the plate so that it lies in the potential flow
region of incompressible fluid. Further assume that the flow is steady and 2-dimensional and
there is no mass flow across the plate. Since we have learnt the method in details earlier, we
proceed directly to write the Steps for solution taking unit width in the normal direction.

Momentum flow rate respectively across section AB and CD in x-direction is

Mass flow rate across section AB and CD

If mBC is the mass flow rate (cflux) across the section BC then mass conservation demands

Negative sign shows that the mass enters the control volume with velocity U and has the
momentum. The net x-momentum flux per unit time to the control volume is
The external force acting on the control volume in the direction of flow are the shear force at the
wall and the pressure force on the surfaces AB and CD that would be equal to h(dp/dx) dx,
calculated in the way as discussed in Chapter 2. Applying the integral momentum theorem, we
get

But

(Bernoulli equation)

and

The integrals have been replaced by using definition of displacement and momentum thickness.6
Equation 5.5.1 is known as Von Karman integral equation7. For a plane plate, dU/dx is zero and
U is not a function of x. Therefore, Eq. 5.5.1 changes to

The integrals in Eq. 5.5.1 can be evaluated if velocity variation with y is known. The integral
solution is based on choosing this profile and comparing results with experimental results. Thus
the accuracy of the solution depends upon the closeness of the chosen velocity profile with the
actual velocity distribution. The procedure in explained through solved examples. However, in
the following paras Von Karman and Pohlhausen solution is given to stress the complexity
involved in case of flow with pressure gradient.

To evaluate the integrals, Pohlhausen suggested choosing velocity profile that would satisfy the
boundary conditions at the surface and the boundary layer edge. These boundary conditions are:
no slip condition at the surface, continuity of the velocity profile at the b.l. edge, and then the
conditions derived with the help of b.l. equations; these are summarised below.

At

At

A word of caution regarding using the boundary conditions. Boundary conditions derived with
the help of b.l. equation must be used only when the main boundary conditions are not enough to
determine the velocity profile and that too alternately at y = 0 and y = δ. These conditions suggest
that a polynomial of any degree for u/U in y/δ may be chosen and the constants in the polynomial
can be determined using the boundary conditions. Karman took the velocity profile u = a + by +
cy2 + dy3 having four unknowns for the flow over a flat plate and chose the boundary conditions
u = v = 0 at y = 0 and u = U, ∂u/∂y = 0 at y = δ to determine the constants. This gives, after
rearranging the terms,

Substituting these results in Eq. 5.5.1, we get for flat plate (dU/dx = 0)

But at x = 0 (leading edge), δ = 0 → C = 0.


This result differs from Blasius solution for which the numerator was 5. The difference points
out that the velocity chosen is deviating from the actual velocity to some extent. One has to try
different velocity profiles so that the result may match the experimental results. There is no set
procedure to choose the exact velocity distribution at any section. The Von Karman solution of
Eq. 5.5.5 is widely accepted.

Non-zero pressure gradient flows. A very large class of flow exists in which case the pressure
varies in the direction of the flow; the pressure may increase or decrease. Velocity profile is
chosen that might be close to the actual profile in that case. We shall briefly mention the solution
proposed by Holstein and Bohlen who took a 4th order polynomial in the variable η, where η =
y/δ and δ is the boundary layer thickness at a section. They introduced a non-dimensional factor λ

to represent pressure gradient. λ is defined as . The chosen polynomial and the


boundary conditions were

At

Using the boundary conditions and simplifying the velocity profile we get

u/U = F(η) + λG(η),

and

λ is known as the shape factor as it determines the shape of the velocity profile in the boundary
layer for non-zero pressure gradient flows. We now substitute these in Eq. 5.5.1, which generates
a complex equation hard enough to be integrated with respect to x. Som and Biswas have
reproduced the solution that originally is in German. Let us define the non-dimensional
parameters

and

Equation 5.5.1 thus gives an equation in K alone. Using these non-dimensionless parameters, the
integral momentum equation reduces to

or,

Walz suggested that right and side of Eq. 5.5.11 can be represented to a fair degree of accuracy
by a linear relation in K so that

here a = 0.47 and b = 6. With this simplification the Eq. 5.5.11 turns into

This gives on integration with the b.c U = 0 at x = 0

Finally the momentum thickness is

Equation 5.5.14 gives the limit of θ at x = 0 using the L’ Hospital rule. The value is
Example 5.5.1 Assuming the velocity distribution in the laminar boundary layer with zero

pressure gradient may be expressed as where u is the velocity at a distance in the


boundary layer, U is the free stream velocity and δ is the b.l. thickness. Determine the drag
coefficient in terms of the Reynolds number.

Solution We need momentum thickness in Eq. 5.5.2, which is obtained form Eq. 5.4.2.

Substituting in Eq. 5.5.2, we get the equation.

Example 5.5.2 Water flows with a free stream velocity of 2 m/s past a flat plate of length L = 10
cm. Using (i) Blasius solution and (ii) the integral solution with a third order velocity profile,
find (a) the thickness of the velocity boundary layer at a location x = 5 cm, (b) fluid velocity at
the same location at a distance y = 0.0225 cm away from the plate, and (c) the drag force per
meter of depth into the paper. Take for water, ρ = 983.1 kg/m3 and n = 0.4748 × 10–6 m2/s.

1. Blasius profile.

1. → δ = 5 × 0.05 × (210615)–1/2 = 0.05447 cm

2. For x = 5 cm and y = 0.225 cm,


From the Table 5.4.1, f ′ = u/U = 0.629 at this value of η
→ u = 1.258 m/s

3. Cf from Eq. 5.4.23 is

F = 0.002046 × ½ ρU2L = 0.4 N


2. Integral solution. Analysis of third order velocity profile is given in the text. Using Eq.
5.5.5, we get δ = 0.05055 cm. Substituting values of δ and y in the velocity profile, we get
Wall shear stress at any x is

Results obtained by two approaches are close to each other.

Example 5.5.3 Water flows between two parallel walls. The velocity is uniform at the entrance
and in the core region. Beyond a distance Le downstream from the entrance, the flow becomes
fully developed so that the velocity varies over the entire width 2h of the channel. In the
boundary layer region, velocity varies as u = U (x) (y/δ)3 where ; A is a constant.
Determine the acceleration on the axis of symmetry for 0 ≤ x ≤ Le.

Solution We fix the coordinates with the bottom plate at the entrance of the channel measuring y
upward. As the flow proceeds, the boundary layer development takes place. let at any x the
boundary layer thickness is δ. Let U0 and U (x) be the flow velocity at the entrance and the core
region of the channel then the mass flow across the section, by mass conservation, is the mass
entering the channel for steady flow gives the equation

or,

We are given and at x = Le, δ = h (flow is fully developed). This gives A and hence δ at
any section. Substituting these results gives

The acceleration will be convective only as the flow is steady.

5.6 Boundary Layer Separation and Control


In section 5.4, we discussed the reason for formation of boundary layer and learnt that the slope
of the velocity profile at the surface is maximum. However, as flow proceeds, the value of the
slope will, in general, change depending upon the external conditions. For example, if pressure is
increasing, referred as adverse pressure gradient, then the slope of the velocity curve will
decrease and it is possible that at some location it becomes zero for large adverse pressure
gradient. This means that at that section the layer next to the layer in contact with the surface has
zero velocity. Therefore the fluid particle is not able to move forward with the fluid. At the
subsequent sections the slope of the velocity profile is negative and the particles at such section
would move in the opposite direction. That explains the phenomenon of boundary layer
separation and its onset is determined by the condition ∂u / ∂n = 0, where n is the direction
normal to the surface.

Fig. 5.6.1 Variation of velocity, its first and second derivative in b.l. region

From b.l. equation also, we may explain the b.l. Separation. Since at the surface, u = v = 0, Eq.
5.4.5 gives μ(∂2u / ∂y2) = dp / dx. Consider the location where the first derivative ∂u / ∂y is zero,
that is, the point of separation. In adverse pressure situation (dp/dx > 0), the velocity profile in
the b.l. is such that (i) the velocity increases from 0 to U joining the main flow smoothly, (ii) the
first derivative is zero at the surface and the b.l. edge, which means it must have at least one
point of maxima in the b.l. region, and (iii) the second derivative, a measure of the curvature of
the velocity curve, is initially positive at the surface and tends to zero at the free surface. At the
point of maxima of ∂u/∂y curve, ∂2u/∂y2 must be zero. Therefore, the curve for ∂2u/∂y2 starts with
some positive value, passes through the origin, and then tends to zero again at the free surface.
The section, where ∂2u/∂y2 is zero, there exists a point of inflexion on the velocity profile. These
observations suggest the kind of the profiles for u, ∂u/∂y and ∂2u/∂y2, as shown in Fig. 5.6.1 (a).
These profiles are also shown for favourable pressure gradient in Fig. 5.6.1 (b). You should now
be able to explain this type of profile in this case.

We may conclude that the deficiency in kinetic energy of the particles close to the surface,
because of viscosity of the fluid, is responsible for the layers to leave the surface. The fluid
particles leaving the surface thereafter get energy from the main flow that helps in pushing it
forward to join back the main flow. It is not necessary that all the particles or layers will leave
the surface from the same section in a three dimensional surface. There may be, and in fact will
be, number of such locations on the surface. These layers from different locations may later
merge with each other. Thus the flow downstream of the point of separation contains vortices,
called eddies, that disturb the flow pattern and hence the pressure distribution. We shall discuss
this in more details in the next section.

Fig. 5.6.2 Boundary layer control

The idea of b.l. separation control stems from its cause itself. That is, the layer close to the
surface may be accelerated or removed so that the flow remains adhered to the surface. Fig. 5.6.2
shows a few methods that have been used successfully for accelerating, energising, or removing
the boundary layer. In the first and second methods, the fluid near the surface on the upper side is
accelerated either by injecting high velocity fluid (blowing) through the holes made in the
surface or bypassing some of the fluid from other side of the body through the slots made in it.
Whatever the method, the slots should be tangential to the surface where the fluid is entering so
that it least disturbs the flow pattern. Blowing method in which a high velocity fluid from outside
source is injected is a common method used in practice. Yet in another method, the slow moving
fluid is sucked through the slots so that the fluid layers remain in contact with the surface.

Streamlining of the surface. This method employs a streamlined body so that the adverse
pressure gradient may not be large when the fluid flows over its curved surface. Recall that it is
one of the important causes of the separation of the boundary layer. We cannot eliminate the
possibilities of adverse pressure gradient when the flow is over curved surface, but we can
reduce this effect.

5.7 Flow over Curved Surfaces

So far mainly we concentrated on the flows over a flat plate. But when the flow takes place over
a curved surface then the area of flow changes causing either acceleration or deceleration of the
flowing fluid. This affects the boundary layer formation along the surface. While passing over a
convex surface, the flow is initially accelerated till it reaches the point after which the area starts
increasing resulting in increase in pressure downstream. It is during this journey of the fluid that
the boundary layer separates as was discussed above. The point at which the flow will separate
and the way it affects the flow downstream depends on the Reynolds number based on the
mainstream velocity and some linear characteristic length. This effect is shown in Fig. 5.7.1.

Fig. 5.7.1 Influence of Reynolds number on wake zone

When Reynolds number is extremely small, around 1, the flow is symmetrical and there is no
point of separation, as shown in Fig. 5.7.1 (a). For Reynolds number in between 4 and 40, the
flow separates and the point of separations moves upstream along the surface with increasing Re.
Thus the flow downstream is divided into two regions—one the wake zone in which eddies are
present and another the main flow, as seen in Fig. 5.7.1 (b). Flow in the wake zone is
symmetrical and steady, the eddies on the upper and lower side rotate in opposite directions, and
the size of the eddy grows with Re. After crossing Reynolds number = 40, the wake region flow
also becomes unsteady and the wake starts undulating (exhibiting a wave-like phenomenon),
resulting in shedding vortices into the stream. At Re 90 or more, the eddies are shed alternately
from upper and bottom side with a regular pattern of alternately shed clockwise and anti-
clockwise vortices, called the Von Karman street, as shown in Fig. 5.7.1 (c). Strouhal number,
defined as fD/U, expresses frequency f of vortex shedding. Von Karman showed that for stability
of the vortex arrangement, h/L = 0.281, where L is the distance between two successive vortices
and h is the distance between the two rows of vortices; these are shown in Fig. 5.7.1 (d). The
alternating breaking away of the vortices produces a periodic force on the cylinder in the
transverse direction. If the natural frequency of the cylinder in the transverse direction equals the
shedding frequency then resonance occurs. This observation is very useful in the aerodynamic
study of stability of suspension bridges, telephone and power transmission lines, etc. as the wind
blowing past these, generates vortices with certain frequency. For Reynolds numbers beyond
500, multiple frequencies start showing up and the wake becomes turbulent.

We learnt in Chapter 3 that the pressure around a cylinder is symmetrical and therefore there is
no force on it, either along or normal to the flow direction. However, the presence of viscous
force disturbs the pressure distribution and the net effect is a force in the direction of flow. This
force is called the skin-friction drag. Wake formation disturbs the pressure distribution further
adding to the increase in the force, called form drag. The extent of the form drag depends on the
dimension of the body normal to the flow direction. Form drag may be comparable to the skin-
friction force for bluff bodies. The skin-friction and the form drag are taken together and are
labelled profile drag, since both of these forces depend on the profile of the body. When the
body moves in a compressible fluid the compressibility effect would give rise to wave-drag. This
drag is also present when the body is not completely submerged in the fluid and waves are
formed at the free surface due to gravity. In high-speed flow, formation of shock at some
location would cause additional force.

We have used the coefficient of drag and lift in earlier chapters, which are the non-dimensional
numbers expressing the force on a body placed in a flowing fluid as a fraction of the product of
dynamic pressure and some area. This area is the area of the body projected on a plane normal to
the direction of the force under consideration. The projected area, for example, is L × D for a
cylinder of diameter D and length L, L × B for flat plate of length L and width B placed normal to
the flow, (π/4)D2 for sphere of diameter D, etc. The expression for these two coefficients is
reproduced below wherein D is the drag force and L is the lift force on the body.

These coefficients were obtained experimentally through a very carefully carried out
experiments for a very large range of Reynolds number and for flow around a flat plate normal to
flow direction, sphere, cylinder and aerofoil; the results are reproduced in Fig. 5.7.2 on a log-log
graph for the coefficient of drag. We observe from this result that
Fig. 5.7.2 Drag coefficient v/s Reynolds number for two-dimensional bodies

1. It is maximum for the flat plate and least for the aerofoil. This agrees with our
perceptions, since the bluff body disturbs the flow most and a streamlined the least. That
is why, we always try to make a body aerodynamically streamlined.
2. For a flat plate the coefficient is constant over the entire range of Re.
3. For a sphere or cylinder, it decreases monotonically up to Re equal to 2000, remains
almost constant up to Re = 2 × 105, and finally decreases steeply by a factor 5. The
decrease in drag force is due to change in nature of boundary layer from laminar to
turbulent. Once the point of separation is onset, the coefficient increases.
4. For aerofoil, it decreases steeply when Re increases from 1000 to 2 × 105 and then stays
almost constant.
Example. 5.7.1 A television transmission antenna consists of a vertical pipe 20 cm diameter and
30 m high on top of a tall structure. Determine the total drag on the antenna and the bending
moment about the base in a 30 m/s wind at NTP.

Solution For air at NTP, μ = 1.79 × 10–5 Ns/m2 and ρ = 1.22 kg/m3.

At this Re, CD from the graph is read, which is 0.2.

D = 0.2 × × 1.22 × 302 ×(30 × 0.2) = 659 Nand acts at a height of 30/2 = 15 m from the
antenna base.

M = 659 × 15 = 9.9 kNm; M is the bending moment.

Example 5.7.2 A kite has an effective area of 0.5 m2 and mass 0.3 kg. It experiences a drag of 12
N in a wind speed of 40 km/h. (a) Determine the tension in the cord if it makes an angle of 45°
with the horizontal and (b) estimate the lift coefficient for the kite.

Fig. E 5.7.2 Kite example

Solution U = 49 × 1000/3600 = 11.11 m/s, W = weight of the kite = 0.3 × 9.81 = 2.943 N and D
= drag = 12 N. The kite is in equilibrium because of (i) W, (ii) T, the tension in the string, (iii) the
drag force D acting in the direction of wind and (iv) the lift force L produced by the wind flow
across the kite. These forces are shown in Fig. E 5.7.2 For equilibrium

D – T cos 45 = 0 and L – W – T sin 45 = 0 → T = 16.97 N and L = 14.94 N.


PROBLEMS

Review Questions

1. What is the difficulty in mathematical solution of viscous flow problems? Is it possible to


get exact solutions?
2. What is the contribution of Prof. G.G. Stokes in obtaining the N–S equations?
3. In what type of flow is div U dropped from the N–S equation?
4. What terms are dropped from N–S equations if the flow is steady?
5. What are the four types of forces that appear in the final form of the N–S equation for
steady, incompressible, constant viscosity flow? Write down the mathematical expression
of these forces.
6. Mention three of the well-known viscous flow problems for which exact solution of the
N–S equation are available.
7. List out the approximations of N–S equation employed in viscous flow analysis. Justify
the approximations in any one case.
8. In laminar flow between parallel plates, why is the velocity maximum at the point of zero
shear stress?
9. What type of velocity distribution in laminar flow between parallel plates would give a
uniform shear stress across the section?
10. The diameter of a plunger and the radial clearance in a dashpot are 89 mm and 0.5 mm,
respectively. If the velocity of the plunger is 2 cm/s, what is the average velocity of the
fluid through the clearance?
11. What is the justification in dropping inertial forces and body forces while analysing
problems in creeping flows?
12. What is the merit of the hydrostatic lubrication? What is its limitation?
13. The viscous torque in a journal bearing of 10 cm diameter is 10 Nm. All other parameters
being same, what is the viscous torque of a 20 cm diameter bearing?
14. Is Prandtl’s boundary layer theory developed for analysing highly viscous fluid flows?
15. Prandtl’s boundary layer equations are approximated versions of N–S equations. Is this
true?
16. Perfect fluid flow theory is capable of providing solution to the flow characteristic inside
the boundary layer. Comment.
17. What is the most remarkable achievement of boundary layer theory?
18. Viscous friction exercises considerable influence near the boundary. Does it mean that
the viscosity increases as the boundary is approached?
19. The shear stress at the boundary of a flat plate decreases with distance from the leading
edge. Explain the reason, both physically and mathematically considering the case of a
laminar boundary layer as an example.
20. Explain why the boundary shear stress in a turbulent boundary layer is more than in a
laminar boundary layer for the same Reynolds number.
21. Justify the formation of a laminar sub layer inside a turbulent boundary layer.
22. A smooth flat plate is covered by a laminar boundary layer over a part of its length
followed by a turbulent boundary layer. Will the computed drag be more or less than the
two drag if the entire boundary layer is assumed to be turbulent?
23. Distinguish between favourable and adverse pressure gradients.
24. Does the boundary layer get separated only in zones of adverse pressure gradient? Can it
not occur in regions where the pressure gradient is favourable?
25. Does the boundary layer separation occur only on curved boundaries and not on flat
surfaces?
26. Which type of drag is predominant in the flow past a bluff body?
27. What is the objective of stream lining an immersed body?
28. A circular cylinder is of 10 cm diameter and 50 cm length. It is exposed to an air stream
of uniform velocity 10 m/s with its curved surface facing the flow. If Cd = 1.2 and density
of air = 1.2 kg/m3, what is the drag?
29. If the drag coefficient in Stoke’s flow past a sphere is 48, what is the Reynolds number?
30. A ping-pong ball of weight 23.64 milli-Newton and diameter 3.56 cm travels with a
velocity of 30 m/s. If the air density is 1.2 kg/m3, what is the lift coefficient?
31. A circular cylinder of 1 m diameter is rotated to have a peripheral velocity of 19 m/s.
what is the circulation around the cylinder?
32. A double stagnation point occurs on a circular cylinder of 0.5 m diameter when kept in a
uniform flow field of velocity 10 m/s. what is the circulation around the cylinder?
33. The drag of an aeroplane should generally be minimised to achieve economy in fuel
consumption. On what occasion is it necessary to have as much drag as possible?
34. Why is it necessary to run aeroplane on the ground for some distance before take-off?

Problems for Practice

1. Assuming the velocity distribution in the boundary layer as , determine the


thickness of the boundary layer and the shear stress intensity 0.8 m from the leading edge
and the drag force on one side of a plate, 1.2 m long and 1.1 m wide placed in water of
velocity 0.2 m/s. Take μ = 0.001 Ns/m2 for water. [10.95 mm, 0.0365 N/m2, 0.0787 N]
2. In a stream of oil of specific gravity 0.95 and kinematic viscosity 0.92 stoke moving at
5.75 m/s, a plate of 500 mm length and 250 mm width is placed parallel to the direction
of flow. Calculate he friction drag on the one side of the plate. Find also thickness of the
boundary layer and the shear stress at the trailing edge of the plate. [14.747 N, 13.9 mm,
58.99 N/m2]
3. A plate 3 m × 1.5 m is held in water moving at 1.25 m/s parallel to its length. If the flow
in the boundary layer is laminar at the leading edge of the plate, (i) find the distance from
the leading edge where the boundary layer flow changes from laminar to turbulent flow,
(ii) find the thickness of the boundary layer at this section, and (iii) find the frictional
drag on the plate considering both its sides. Take μ = 0.001 Ns/m2. [400 mm, 2.78 mm,
21.8 N]
4. A 1/60 model of a ship was towed in water at a velocity of 0.4 m/s, the model having a
length of 2.5 m and a draft of 0.2 m. Each side of the immersed part of the hull may be
taken to be a plate 2.5 m deep. Calculate the frictional drag on the hull and the thickness
of the boundary layer at the stern of the model. If the total drag on the model is found to
be 1.175 N, find the drag on the prototype. Take μ = 0.001 Ns/m2. [0.3712 N, 59.5 mm,
1971116 N.]
5. The coefficient of drag due to boundary layer over a thin flat plate 4 m long and 0.3 m

wide is given by , where Rl is the Reynolds number at the length l. If the


plate is kept parallel to the flow of water having a free stream velocity of 4 m/s, find the
total drag force considering both sides of the plate. Take kinematic viscosity of water = 1
× 10–6 m2/s.
6. A stream lined automobile with a projected area of 2.75 m2 on a vertical plane at right
angles to the direction of motion, runs at 108 km/hr while wind blows at a velocity of 36
km/hr in the opposite direction. Find the drag resistance due to wind action and the power
required to overcome the resistance. Take CD = 0.42. If however, the wind blows in the
lateral direction at 36 km/hr at right angles to the direction of motion of the automobile,
what would be the drag force tending to topple the vehicle. Area of the vehicle exposed
to wind for this case may be taken as 6 m2 and the corresponding values of CD may be
taken as 0.75. Density of air = 1.27 kg/m3. [1173.48 N, 35.2 kW, 285.75 N].
7. At what distance from centre of the tube of radius r0 does the average velocity occur in
laminar flow? [0.707 r0]
8. The cup anemometer shown in P 5.7 is placed in wind blowing at a velocity of 54 km/hr.
Ignoring bearing friction find the speed of rotation. Take CD = 0.34 for the convex
surface and 1.33 for the concave surface. [35.2 rpm]

Fig. P 5.7 Cup anemometer

9. A paratrooper descends in air of density 1.25 kg/m3 with a parachute, which is


hemispherical having a diameter of 3.25 m. If the weight of the paratrooper and the
parachute is 785 N, find the terminal velocity reached by him. Take CD = 0.5. [17.4 m/s].
10. A vehicle having a projected area of 6 m2 is travelling at 80 km/hr. The resistance to the
motion of the vehicle is 2160 N. If 20% and 10% of the total resistance are due to rolling
and surface friction respectively and the balance resistance is the form drag, determine
the form drag coefficient. Take ρ = 1.25 kg/m3 for air. [0.817]
11. A Clark-Y type aerofoil moves at a velocity of 15 m/s through standard atmosphere at an
elevation of 2.5 km. The aerofoil has a chord length of 2.25 m and a span of 13.5 m, the
angle of attack being 5° 25′. Calculate the weight, which the wing carries, and the power
required to drive the aerofoil. Take corresponding to α = 5° 25′, CL = 0.465 and CD =
0.02. Density of air = 1.25 kg/m3. [137.9 kN. 815.75 kW]
12. Oil flows through a pipe 15 cm diameter and 650 m in length with a velocity of 0.5 m/s.
if the kinematic viscosity of oil is 18.7 stokes; find the power lost in overcoming the
friction. Take specific gravity of oil = 0.9. [24.73 W]
13. Oil of specific weight 910 kg/m3 and kinematic viscosity 0.002 m2/s is pumped through a
15 cm dia 300 m long pipe at the rate of 20 tonnes per hour. Show that the flow is viscous
and find the head lost. [30 m]
14. Oil is pumped along a horizontal 15 cm dia pipe 200 m long. The specific gravity of the
oil is 0.8 and its kinematic viscosity is 1.3 stokes. Flow is laminar so that the fictional
coefficient for the pipe will be 16/Re in which Re is the Reynolds number. It has 12 kW
motor to drive the pump, which has an efficiency of 65%. Find the quantity of the oil
flowing through the pipe in a minute. If the pipe is not horizontal with the other end 10 m
above the level of the oil in the pump, find the power required to deliver 0.03 m3/s.
15. Castor oil of viscosity 9.871 poise flows between two plates 4 mm apart. If the difference
of pressure between the two sections 100 m apart were 0.025 bar, find the rate of flow
between the plates.
16. A container full of oil having viscosity 2 poise, has a horizontal parallel crack in its wall
which is 60 cm wide, 5 cm thick, in the direction of flow. Find the flow of oil leaking per
hour through the crack, which forms a gap of 0.05 cm between parallel surfaces. The
pressure difference between inside and outside faces of the gap is 0.15 bar.
17. In the injection moulding of a certain plastic the heated compound is forced under a
pressure difference of 7 kgf/cm2 through a tube of 0.5 cm diameter and 25 cm length. If
the mean velocity of flow is 40 cm/s, calculate the viscosity of the compound, assuming
laminar flow. If the specific gravity of the compound is 1.5, calculate the Reynolds
number for the flow and verify the assumption. [5.365 Ns/m2]
18. A belt conveyor consists of a flat belt 0.5 m wide which slides at a velocity of 4 m/s
parallel to a surface separated by a 6 cm thick layer of oil of viscosity 0.25 Ns/m.
determine (i) the pressure gradient required to cause no shear stress at the belt surface, (ii)
the average velocity and the discharge of oil to be maintained for the above. [–555.55
N/m2, 2.67 m/s, 0.08 m3/s]
19. A coaxial cylinder viscometer consists of the inner stationary cylinder of diameter 10 cm
and the outer drum of internal diameter 10.2 cm. When the outer drum rotates at 500 rpm,
the torque registered on the torsion meter attached to the inner cylinder is 3 Nm.
Determine the viscosity of the fluid filled in the viscometer to a depth of 20 cm. [0.365
Ns/m2]
20. A viscometer consisting of a long vertical tank containing a fluid at the desired ambient
conditions and a small sphere is employed for determining the viscosity of oil. If a steel
sphere (specific gravity 8.0) of 1 cm diameter falls with a constant velocity of 5 cm/s
through the lubricating oil (specific gravity 0.9), calculate the viscosity of the oil.
21. A 2 m diameter cylinder rotates at 1800 rpm in an air stream (ρ = 1.225 kg/m3) of 25 m/s.
Estimate the theoretical lift per unit length of the cylinder. [11.55 kN]
22. Experiments were conducted in wind tunnel with a wind speed of 50 km/hr on a float
plate 1 m long and 1 m wide. The specific weight of air is 1.15 kg/m3. The plate is kept at
such an angle that the coefficients of lift and drag are 0.75 and 0.15 respectively.
Determine (i) the lift force, (ii) the drag force, (iii) the resultant force, and (iv) power
exerted by the air stream on the plate. [83 N, 16.7 N, 84.7 N, 78.70, 0.231 kW]
23. Determine the final steady rate at which an air bubble, 0.5 mm dia, rises in a jar of honey
with specific gravity 1.3 and dynamic viscosity 0.5 Ns/m2. [0.354 mm/s]
24. What will be the water discharge through 0.60 m thick filter of cross-sectional area 15 m2
if it is made of 0.2 mm diameter uniform sand? Assume the head loss through the filter as
0.2 m, water viscosity 0.001 kg/ms, and porosity of filter 0.5. [0.05445 m3/s]
25. Below are given the wind tunnel measurements of velocity at different elevation from the
boundary

What is the free stream velocity? Determine δ the thickness of nominal boundary layer;
also determine the exponent n in the equation u/U = (y/d)1/n. Is the boundary layer
laminar or turbulent? Comment on the observation and justify the increase in subsequent
stations for the measurement as we move away from the boundary. [10.38 m/s, 5 mm, 6,
turbulent]

26. If velocity distribution in laminar boundary layer over a flat plate is assumed to be given
by u = a + by + cy2, determine its form using the necessary boundary conditions. [u/U =
2(y/δ) – (y/δ)2]
27. A wind tunnel has a cross-section at its inlet of 1.0 m × 1.0 m and a length of 10 m. Wind
at uniform velocity of 15 m/s enters the tunnel at 20°C. Determine the cross-sectional
dimension at the end of the test section which will yield zero pressure gradient along its
length. Assume velocity distribution in turbulent boundary layer to follow the law u / U =
(y / δ)1/5. [1.052 m × 1.052 m]
28. Find the ration of friction drag on the front half and rear half of the flat plate kept at zero
incidence in a stream on uniform velocity, if the boundary layer is laminar over the whole
plate. [2.414]
29. Boeing has a wing of chord length 6.0 m. For standard air, determine the location on the
wing where boundary layer will change from laminar to turbulent when it travels at 260
km/hr. What will be the location of this transition if it travels at the same speed at 20 km
altitude. For standard air: ρ = 1.225 kg/m3 and u = 1.461 × 10–5 m2/s. At 20 km altitude, ν
= 1.599 × 10–5 m2/s. [1.101 m, 1.107 m]
30. If turbulent boundary layer on a flat plate has the velocity distribution given by u / U =
8.74 (Uy / ν)1/7, obtain the relationship for variation of δ/x.
31. Show that in the length of establishment for laminar flow in a circular horizontal pipe, the
frictional force of he pipe wall is given by F = πR2 (p1 – p2 – 1/3 ρU2), where U is the
average velocity of flow in the pipe of radius R, and p1 and p2 are the pressure at sections
1 and 2 respectively.

Find answers on the fly, or master something new. Subscribe today. See pricing options.

Back to top
 Recommended
 Playlists
 History
 Topics
 Settings
 Support
 Get the App
 Sign Out

© 2019 Safari. Terms of Service / Privacy Policy



You might also like