Fluid Mechanics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

FLUID MECHANICS

DEFINITION

Fluid mechanics is the study of the behaviour of fluids, either at rest (fluid statics) or in motion (fluid
dynamics).

Fluid mechanics is involved in nearly all areas of civil engineering either directly or indirectly. Some
examples of direct involvement are those concerned with manipulating the fluid:

o Sea and river (flood) defences


o Water distribution/sewerage (sanitation) networks
o Hydraulic design of water/sewage treatment works
o Dams
o Irrigation
o Pumps and turbines
o Water retaining structures

And some examples where the primary object is construction – yet analysis of the fluid mechanics is
essential:

o Flow of air in or around buildings


o Bridge piers in rivers
o Groundwater flow

THE CHARACTERISTICS OF FLUIDS

In everyday life, three states of matter are recognized: solid, liquid and gas. Although different in
many respect, liquids and gases have a common characteristic in which they differ from solids: they
are fluids, lacking the ability of solids to offer a permanent resistance to a deforming force.

DEFINITION OF A FLUID

A fluid is a substance that deforms continuously when subjected to a shear stress, no matter how
small that shear stress maybe. Conversely, it follows that:

If a fluid is at rest, there can be no shearing forces acting and, therefore, all forces in the fluid must
be perpendicular to the planes upon which they act.

B B’ C C’
F

F D
A

1
LIQUID AND GASES

A fluid can be either a liquid or a gas. A liquid is a state of matter in which the molecules can move
relatively freely to each other, but are restricted by cohesive forces to maintain a relatively fixed
volume. Conversely, a gas is a state of matter in which the molecules are practically unrestricted by
cohesive forces. A gas has neither shape nor volume.

In the analysis of the behaviour of fluids the most important difference between liquids and gases is
that, whereas under ordinary conditions liquids are so difficult to compress that they may for most
purposes be regarded as incompressible, gases may be compressed much more readily. The study of
compressible fluids requires the use of the laws of thermodynamics.

THE CONTINUUM

All fluids, whether they are liquids or gases, are composed of molecules in continuous motion,
widely spaced for a gas, closely spaced for a liquid. In both cases, however, the spacing is very large
when compared to the molecular diameter. At this level a fluid property like density has no precise
meaning because the number of molecules occupying a given volume continually changes. In most
engineering applications, however, interest centres on the average conditions of velocity, pressure,
temperature, density and so on. Under these conditions the fluid can be considered as a continuum,
that is a continuous distribution of matter with no empty space.

PROPERTIES OF FLUIDS

The study of properties of fluids is basic for the understanding of flow or static condition of fluids.
The following properties of fluids are of general importance to the study of fluid mechanics.

DENSITY

The density ρ of a fluid is its mass per unit volume. The mass density at a point is determined by
considering the mass δm of a very small volume δV surrounding the point. In order to preserve the
concept of the continuum, it is required that δV approaches a limiting value such that

m
  lim
V  V
Where ε is still large compared to the average distance between molecules.

Density is highly variable in gases, and increases nearly proportionally to the pressure level. Density
in liquids on the other hand is nearly constant, and can be considered to be virtually independent of
the pressure level. The most liquids are treated analytically as being incompressible. In general
liquids are about three orders of magnitude more dense than gases at atmospheric pressure. Typical
values at atmospheric pressure (Patm = 1.013 x 105 N/m3): water, 1000 kg/m3; air, 1.23 kg/m3.

SPECIFIC WEIGHT

The specific weight w of a fluid is its weight per unit volume. Just as mass has a weight W  mg ,
density and specific weight are simply related by gravity:

w  g

2
Units: newtons per cubic metre (N/m3)

Typical values: water, 9.81 x 103 N/m3 (9.81 kN/m3); air, 12.07 N/m3

SPECIFIC GRAVITY

The specific gravity (or relative density) s.g. of a fluid is simply the ratio of a fluid density to a
standard reference fluid, usually water at 4oC for liquids, and air for gases. Thus, for liquids

𝑑𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑠𝑢𝑏𝑠𝑡𝑎𝑛𝑐𝑒 𝜌𝑠𝑢𝑏𝑠𝑡𝑎𝑛𝑐𝑒


𝑠. 𝑔. = 𝑜 =
𝑑𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑤𝑎𝑡𝑒𝑟 𝑎𝑡 4 𝐶 𝜌𝑤𝑎𝑡𝑒𝑟 𝑎𝑡 4𝑜 𝐶

Example

Mercury has a density of 13600 kg/m3. What are its specific weight and specific gravity?

SPECIFIC VOLUME

The specific volume of a fluid is simply the reciprocal of the density i.e. it is the volume occupied by a
unit mass of fluid.

V 1
 
m 

TEMPERATURE

Temperature T is related to the internal energy level of a fluid. Although engineers often use Celsius
or Fahrenheit scales for convenience, many applications require absolute (Kelvin or Rankine)
temperature scales.

VAPOUR PRESSURE

Liquids evaporate because of molecules escaping from the liquid surface. If the space above the
liquid is confined a point will be eventually reached where the number of molecules escaping the
liquid is balanced by the number of vapour molecules condensing and returning to the liquid and a
state of equilibrium between liquid and vapour will exist. The partial pressure exerted by the vapour
molecules on the liquid surface when this point is reached is referred to as the vapour pressure.

The vapour pressure of a given fluid depends on the temperature, and increases with increasing
temperature.

When the pressure above a liquid equals the vapour pressure of the liquid boiling occurs. In a
flowing liquid, if the pressure of the liquid is reduced below the value of the vapour pressure then
local boiling, referred to as cavitation, can occur.

COMPRESSIBILITY

All matter is to some extent compressible. The compressibility of a perfect gas can be described by
the perfect gas law. For most purposes a liquid can be considered as incompressible, but for
situations involving either sudden or great changes in pressure its compressibility becomes

3
important. The compressibility of a liquid is expressed by its bulk modulus of elasticity, K, which is
defined by the equation

P
K 
V
V

Where P is the increase in pressure, which when applied to a volume V, results in a decrease in
volume V .

The reciprocal of bulk modulus is termed the compressibility, implying that

1
𝑐𝑜𝑚𝑝𝑟𝑒𝑠𝑠𝑖𝑏𝑖𝑡𝑦 =
𝐾

Typical values of bulk modulus: water, K = 2.05 x 109 N/m2; oil, K = 1.62 x 109 N/m2

SURFACE TENSION

Surface tension σ is caused by the force of cohesion at the free surface. A liquid molecule in the
interior of the liquid mass is surrounded by other molecules all around and is in equilibrium. At the
free surface of the liquid, there are no liquid molecules above the surface to balance the force of the
molecules below it. Consequently, as shown below, there is a net inward force on the molecule. The
force is normal to the liquid surface. At the free surface a thin layer of molecules is formed. It is
because of this film that a small needle can float on the free surface.

Surface tension is usually expressed in N/m. The formation of bubbles, droplets and free jets are due
to the surface tension of the liquid.

4
PRESSURE INSIDE A WATER DROPLET

There is a natural tendency for liquids to minimize their surface area. For this reason, drops of liquid
tend to take a spherical shape in order to minimize surface area.

For such a small droplet, surface tension will cause an increase of internal pressure P in order to
balance the surface force.

From free body diagram, we have


𝜋
i. 𝑃𝑟𝑒𝑠𝑠𝑢𝑟𝑒 𝑓𝑜𝑟𝑐𝑒 = 𝑃 × 𝑑 2
4
ii. Surface tension force acting around the circumference = 𝜎 × 𝜋𝑑

Under equilibrium conditions these two forces will be equal and opposite i.e.
𝜋
𝑃 × 𝑑 2 = 𝜎 × 𝜋𝑑
4
4𝜎
=> 𝑃=
𝑑

Example

In order to form a stream of bubbles, air is introduced through a nozzle into a tank of water at 20 oC.
If the process requires 3.0 mm diameter bubbles to be formed, by how much the air pressure at the
nozzle must exceed that of the surrounding water?

Take surface tension of water at 20 oC = 0.0735 N/m.

CAPILLARITY

Rise or fall of a liquid in a capillary tube is caused by surface tension and depends on the relative
magnitude of cohesion of the liquid and the adhesion of the liquid to walls of the containing vessel.

Liquids rise in tubes they wet (adhesion > cohesion) and fall in tubes they do not wet (cohesion >
adhesion).

5
WETTING AND CONTACT ANGLE

Fluids wet some solids and do not others.

The figure shows some of the possible wetting behaviours of a drop of liquid placed on a horizontal,
solid surface. Fig (a) represents the case of a liquid which wets a solid surface well. The angle θ
shown is the angle between the edge of the liquid surface and the solid surface, measured inside the
liquid. This angle is called the contact angle and is a measure of the quality of wetting. For perfect
wetting, in which the liquid spreads as a thin film over the surface of the solid, θ is zero.

Fig (c) represents the case of no wetting. If there was exactly zero wetting, θ would be 180o.
However, the gravity force on the drop flattens the drop, so that 180 o angle is never observed.

Capillarity is important (in fluid measurements) when using tubes smaller than about 10 mm in
diameter.

The figure above represents capillary rise (or depression) in a tube, which is given approximately by

2 cos 4 cos
h or h
gr gd

Where

6
h = height of capillary rise (or depression); σ = surface tension; θ = wetting angle; d = diameter of
tube

Example

Water has a surface tension of 0.4 N/m. In a 3 mm diameter vertical tube a liquid rises 6 mm above
the liquid outside the tube, calculate the contact angle.

VISCOSITY

Viscosity, μ, is the property of a fluid, due to cohesion and interaction between molecules, which
offers resistance to shear deformation. Different fluids deform at different rates under the same
shear stress. Fluid with a high viscosity such as syrup deforms more slowly than fluid with a low
viscosity such as water.

When a fluid is in motion shear stresses are developed if the particles of the fluid move relative to
one another. When this happens adjacent particles have different velocities. If fluid velocity is the
same at every point then there is no shear stress produced; the particles have zero relative velocity.

Consider the flow in a pipe in which water is flowing. At the pipe wall the velocity of the water will
be zero. The velocity increases towards the centre of the pipe. This change in velocity across the
direction of flow is known as velocity profile and is shown graphically in the figure below:

NEWTON’S LAW OF VISCOSITY

Consider a 3-D rectangular element of fluid, like that in the figure below:

7
The shearing force F acts on the area on the top of the element. This area is given by A  z  x .
Thus,

𝐹
𝑠ℎ𝑒𝑎𝑟 𝑠𝑡𝑟𝑒𝑠𝑠, 𝜏 =
𝐴

The deformation which this shear stress causes is measured by the size of the angle  and is known
as shear strain.

It has been found experimentally that shear stress is directly proportional to the rate of shear strain .

If the particle at point E (in the figure above) moves under the shear stress to point E’ and it takes
time t to get there, it has moved the distance x. For small deformations
𝑥
𝑠ℎ𝑒𝑎𝑟 𝑠𝑡𝑟𝑎𝑖𝑛, ∅ =
𝑦
𝑥 𝑢
𝑟𝑎𝑡𝑒 𝑜𝑓 𝑠ℎ𝑒𝑎𝑟 𝑠𝑡𝑟𝑎𝑖𝑛 = =
𝑡𝑦 𝑦
𝑥
Where = 𝑢 is the velocity of the particle at E.
𝑡

Using the experimental result that shear stress is proportional to rate of shear strain, then
𝑢
𝜏 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 ×
𝑦
𝑢
The term is the change in velocity with y, or the velocity gradient, and may be written in the
𝑦
𝑑𝑢
differential form . The constant of proportionality is known as the dynamic (absolute) viscosity, μ,
𝑑𝑦
of the fluid. Thus

𝑑𝑢
𝜏= 𝜇
𝑑𝑦

This is known as Newton’s law of viscosity. In the equation, τ is the shear stress in N/m2, 𝑑𝑢⁄𝑑𝑦 is
the velocity gradient or rate of shear strain, and has units of radians/s, μ is the coefficient of dynamic
viscosity.

COEFFICIENT OF DYNAMIC VISCOSITY

The coefficient of dynamic viscosity, μ, is defined as the shear force per unit area, required to drag
one layer of fluid with unit velocity past another layer a unit distance away.

𝑑𝑢
𝜇 = 𝜏/
𝑑𝑦

Units: Newton seconds per square metre, Ns/m2 or kilograms per metre per second, kg/m.s or Pa.s.

(Although μ is often expressed in poise, P, where 10P = 1 Ns/m2)

8
Typical values: water = 1.14 x 10-3 Ns/m2, Air = 1.78 x 10-5 Ns/m2, mercury = 1.552 Ns/m2, paraffin oil
= 1.9 Ns/m2.

KINEMATIC VISCOSITY

In engineering practice it is often convenient to introduce the term Kinematic viscosity, ν, defined as
the ratio of dynamic viscosity to mass density.




Units: square metres per second, m2/s

(Also note that ν is often expressed in stokes, st, where 104 st = 1 m2/s)

Typical values: water = 1.14 x 10-6 m2/s, Air = 1.46 x 10-5 m2/s, mercury = 1.145 x 10-4 m2/s, paraffin
oil = 2.375 x 10-3 m2/s.

Example 1

The velocity distribution for flow over a plate is given by u = 2y – y2 where u is the velocity in m/s at a
distance y metres above the plate. Determine the velocity gradient and shear stress at the boundary
and 0.15m from it. Take dynamic viscosity of fluid as 0.9 Ns/m 2.

Example 2

A hydraulic lift used for lifting automobiles has a 25 cm diameter ram which slides in a 25.018 cm
diameter cylinder, the annular space being filled with oil having a kinematic viscosity of 3.7 cm2/s
and relative density of 0.85. If the rate of travel of the ram is 15 cm/s, find the frictional resistance
when 3.3 m of ram is engaged in the cylinder.

NEWTONIAN AND NON-NEWTONIAN FLUIDS

NEWTONIAN FLUIDS

Fluids which obey Newton’s law of viscosity are known as Newtonian fluids. All gases and most
liquids which have simpler molecular formula and low molecular weight such as water, benzene,
kerosene and most solutions of simple molecules are Newtonian fluids.

NON-NEWTONIAN FLUIDS

Fluids which do not obey Newton’s law of viscosity are known as non-Newtonian fluids. Such fluids
are relatively uncommon. They are generally complex mixtures. These include slurries, mud flows,
polymer solutions, blood etc.

9
The figure above shows a relationship between shear stress, τ and velocity gradient, du . For
dy
Newtonian fluids, there is a linear relationship between shear stress and velocity gradient. The slope
of the curve produced is a measure of the absolute viscosity, μ.

In Newtonian fluids, the slope is constant and, therefore, the viscosity is constant. The slope of the
curves for non-Newtonian fluids varies.

Newtonian and non-Newtonian fluids can be represented by the equation

n
 du 
  A  B 
 dy 

Where A, B and n are constants. For Newtonian fluids A = 0, B = μ and n = 1.

Three types of non-Newtonian fluids can be defined:

1. Pseudoplastic or Thixotropic: the plot of shear stress versus velocity gradient lies above the
straight, constant sloped line for Newtonian fluids. The curve begins steeply, indicating a
high absolute viscosity. Then the slope decreases with increasing velocity gradient. Examples
of such fluids are blood plasma, syrups, adhesives, molasses, and inks.

2. Dilatant fluids: the plot of shear stress versus velocity gradient lies below the straight line for
Newtonian fluids. The curve begins with a low slope, indicating a low absolute viscosity.
Then, the slope increases with increasing velocity gradient. Examples are slurries with high
concentrations of solids such as corn starch, starch in water etc.

3. Bingham fluids: sometimes called plug-flow fluids, Bingham fluids require the development
of a significant level of shear stress before flow will begin. Once flow starts, there is an
essentially linear slope to the curve indicating a constant viscosity. Examples of Bingham
fluids are chocolate, mayonnaise, toothpaste and sewage sludge.

10
HYDROSTATICS

Hydrostatics is the study of fluid problems in which there is no relative motion between fluid
elements. With no relative motion between individual elements (and thus no velocity gradients), no
shear can exist, whatever the viscosity of the fluid is. Accordingly, viscosity has no effect in static
problems and exact analytical solutions to such problems are relatively easy to obtain. Hence, all
free bodies in fluid statics have only normal pressure forces acting on them.

INTRODUCTION TO PRESSURE

Atmospheric pressure, gauge pressure and absolute pressure

The pressure of the atmosphere at the earth’s surface is measured by a barometer. At sea level the
atmospheric pressure averages 101 kN/m2 and is standardised at this value. There is a decrease in
atmospheric pressure with altitude; for instance, at 1500 m it is reduced to 88 kN/m 2.

As most pressures encountered in hydraulics are above atmospheric pressure, it is convenient to


regard atmospheric pressure as the datum (Patm = 0). Pressures are then referred to as gauge
pressures when above atmospheric and vacuum pressures when below it. The pressure in a vacuum
is called absolute zero, and all pressures referenced with respect to this zero pressure are termed
absolute pressures.

Many pressure measuring devices measure not absolute pressure but only difference in pressure.
The pressure difference is normally between the pressure in the fluid and the pressure in the
atmosphere.

Thus

Pabs = Pgauge + Patm

Where

Pabs = absolute pressure

Pgauge = gauge pressure

Patm = atmospheric pressure

The diagram below is a graphical representation of these pressures.

11
Gauge Pressure
(+ve pressure)
Absolute
Pressure

Atmospheric Pressure (Patm = 0)

Vacuum Pressure
(-ve pressure)

Absolute zero

For example, if a pressure of 50 kN/m2 is measured with a gauge referenced to the atmosphere and
the atmospheric pressure is 100 kN/m2, then the pressure can be expressed as either P = 50 kN/m 2
(gauge) or P = 150 kN/m2 (absolute).

If on the other hand, a gauge indicates a vacuum pressure of 31 kN/m 2, then this can be stated as 70
kN/m2 (absolute), or -31 kN/m2 (gauge), assuming that the atmospheric pressure is 101 kN/m 2
(absolute).

Pressure intensity

The pressure intensity or more simply the pressure on a surface is the pressure force per unit area.

P h

In the figure above, the vertical downward pressure force acting on the horizontal plane is equal to
the weight of the prism of fluid which is vertically above it plus the pressure intensity at the interface
with another fluid. In the case of an incompressible liquid in contact with the atmosphere, the gauge
pressure, P is given by

P  gh

  density of the liquid

g = acceleration of gravity

h = depth below the free surface

h is referred to as the pressure head and is generally stated in metres of liquid. The units of pressure
are N/m2 or kg/m.s2. This unit is also known as Pascal, Pa, i.e. 1 Pa = 1 N/m2. Also frequently used is
the alternative SI unit, the bar, where 1bar = 1 x 10 5 N/m2.

12
Pascal’s law for pressure at a point

Pascal’s law states as follows:

“The intensity of pressure at any point in a liquid at rest is the same in all directions.”

To prove Pascal’s law for pressure at a point, consider a small fluid element in the form of a
triangular prism ABCDEF surrounding a point in the fluid. A relationship can be established between
the pressures Px in the x-direction, Py in the y-direction, and Ps normal to any plane inclined at any
angle  to the horizontal at this point.

Px is acting at right angle to ABEF, and Py at right angle to CDEF, similarly Ps at right angle to ABCD.
Since there can be shearing forces for a fluid at rest, and there will be no accelerating forces, the
sum of forces in any direction must be zero. The forces acting are due to the pressures on the
surrounding and the gravity force.

Force due to Px = Px  Area ABEF = Pxyz

Horizontal component of force due to Ps = - (Ps x Area ABCD) sin() = - Pssz y/s = -Psyz

As Py has no component in the x direction, the element will be in equilibrium, if

Pxyz-Psyz) = 0

 Pxδyδz = Psδyδz

i.e. Px = Ps

Similarly in the y direction, force due to Py = Pyxz

Component of force due to Ps = - (Ps x Area ABCD) cos() = - Pssz x/s = - Psxz

Force due to weight of element = - mg = - gV = -  (xyz/2) g

13
Since x, y, and z are very small quantities, xyz is negligible in comparison with other two
vertical force terms, and the equation reduces to,

Py = Ps

Therefore, Px = Py = Ps

Which proves that pressure at a point is the same in all directions. This also shows that pressure
intensity is independent of the angle of inclination of the elemental surface.

PRESSURE VARIATION IN A STATIC FLUID

The fundamental equation of fluid statics is that relating pressure, specific weight and vertical
distance in a fluid. This equation may be derived by considering the static equilibrium of a typical
differential element of fluid in the figure below. The z-axis is in the direction parallel to the
gravitational force field (vertical).
Q

 p  p A
l
P

z  z
pA
z

Datum

Let the pressure at the end P be p and that at the end Q be p+δp. The force on the end P is therefore
pδA and the force on the end Q is (p+δp)δA. If the length of the cylinder is δl its volume is δAδl and
its weight ρgδAδl where ρ represents the density and g the acceleration due to gravity. Since no
shear forces are involved in a fluid at rest, the only forces acting on the sides of the cylinder are
perpendicular to them and therefore have no component along the axis.

For equilibrium, the algebraic sum of the forces in any direction must be zero. Resolving in the
direction QP:

 p  p A  pA  gAl cos  0


Now if P is at a height z above some suitable horizontal datum plane and Q is at height z+δz, then
the vertical difference in level between the ends of the cylinder is δz and δlcosθ=δz. The equation
above simplifies to

p  gz  0

Implying that

14
p
  g
z

The minus sign indicates that the pressure decreases in the direction in which z increases, that is,
upwards.

The equation above may be integrated to find

p2 z2
 p1
p   g  z
z1

p 2  p1   g z 2  z1 

Variation of pressure vertically in a fluid under gravity

Consider a hypothetical differential cylindrical element of fluid of cross sectional area A and height
(z2 - z1 = h) below.

Upward force due to pressure P1 on the element = P1A

Downward force due to pressure P2 on the element = P2A

Force due to weight of the element = mg = A(z2 - z1)g

Equating the upward and downward forces,

P1A = P2A + A(z2 - z1)g

P2 - P1 = - g(z2 - z1)

Or p 2  p1   gh

Thus in any fluid under gravitational acceleration, pressure


decreases, with increasing height z in the upward direction.

Equality of pressure at the same level in a static fluid

Consider a horizontal cylindrical element of fluid in the figure below, with cross-sectional area A, in
a fluid of density ρ, pressure P1 at the left hand end and pressure P2 at the right hand end.

The fluid is at equilibrium so the sum of the forces acting in the x-direction is zero.

15
Equating the horizontal forces, P1A = P2A (i.e. sum of the horizontal forces must be zero)

Equality of pressure at the same level in a continuous body of fluid

Pressures at the same level will be equal in a continuous body of fluid, even though there is no
direct horizontal path between P and Q provided that P and Q are in the same continuous body of
fluid.

We know that, PR = PS

PR = PP + gh  1

PS = PQ + gh  2

MEASUREMENT OF FLUID PRESSURE

In the case of liquids with a free surface the pressure at any point is represented by the depth
below the surface. When the liquid is totally enclosed, such as in pipes and pressure conduits, the
pressure cannot readily be ascertained and a suitable measurement device is required. There are
three principal types:

(a) Piezometer
(b) Manometer, and
(c) Bourdon gauge

PIEZOMETER

If a tapping is made in the boundary surface and a sufficiently long tube connected, the liquid will
rise in the tube until balanced by atmospheric pressure. The pressure in the main body of the liquid
is represented by the vertical height of the liquid column. Clearly, the device is only suitable for
moderate pressures, otherwise the liquid will rise too high in the piezometer tube.

16
Thus, pressure at A = Pressure due to column of liquid of height h, or

P  gh

A barometer is a special type of a piezometer. It is used for measuring atmospheric pressure. A


simple barometer consists of a tube more than 760 mm long inserted in open container of mercury
with a closed tube end at the top and an open tube end at the bottom and with mercury extending
from the container up into the tube. The tube is evacuated at the top and the only pressure causing
the mercury to rise in the tube is that of the atmosphere.

The atmospheric pressure is calculated from the relation

Patm  gh

Where ρ is the density of fluid in the barometer.

MANOMETERS

Piezometers cannot be employed when large pressures in the lighter liquids are to be measured,
since this would require very long tubes, which cannot be handled conveniently. Furthermore, gas
pressures cannot be measured by the piezometers because a gas forms no free atmospheric surface.
These limitations can be overcome by the use of U-tube manometers.

A U-tube manometer consists of a glass tube bent in U-shape, one end of which is connected to a
point at which pressure is to be measured and the other end remains open to the atmosphere.

The following are the different types of U-tube manometers:

(a) Simple U-tube manometer


(b) Inverted U-tube manometer

17
(c) U-tube with one limb enlarged
(d) Two fluid U-tube manometer
(e) Inclined U-tube manometer
(f) Differential manometer

SIMPLE U-TUBE MANOMETER

Pressure in a continuous static fluid is the same at any horizontal level so,

For the left hand arm

For the right hand arm

As we are measuring gauge pressure we can subtract giving

DIFFERENTIAL MANOMETERS

A U-tube gauge is arranged to measure the pressure difference between two points in a pipeline. As
before, the principle involved in calculating the pressure difference is that the pressure at the same
level in the two limbs must be same, since the fluid in the bottom of the U-tube is at rest.

18
For the left-hand side:

Pc  PA  gha

For the right-hand side:

PD  PB  g hb  h   m gh

Since PC = PD

Then PA  gha  PB  g hb  h   m gh

Pressure difference, PA  PB  g hb  ha   hg m   

Example

Determine the pressure difference between the water pipe and the oil pipe shown in the figure
below.

19
U-TUBE MANOMETER WITH LIMBS ENLARGED (MICRO-MANOMETER)

Small differences in liquid levels are difficult to measure and may lead to significant errors in
reading. Using an arrangement as shown in below, the reading may be amplified. For improved
accuracy the manometer fluid density should be close to that of the fluid used for measurement.

The equilibrium equation for this arrangement is:

 y  y
p1   A g h  z    B g  z  z    p 2   A g h  z    B g  z  z     c gy
 2  2

The amount of liquid on each side remains constant. Therefore

20
y a y
a1z  a2  z  2 
2 a1 2

Substituting for z

  a  a 
p1  p 2  gy C   B 1  2    A 2 
  a1  a1 

Since a2 is usually very small compared with a 1, then

p1  p2  C   B gy

Example

A manometer with two enlarged limbs is to be used to find the pressure difference of air flowing in a
pipeline between two points A and B. The air density is 1.2 kg/m 3. The manometric fluid is having a
specific gravity of 1.1 and the filler fluid is water. Under measuring conditions, the manometric fluid
movement on the pressure side is 5 cm. Determine the pressure difference between the two points
A and B, if the area of the well chamber is 10 times that of the tube.

Air

A B

Filler
fluid

y
Manometric
fluid

INCLINED-TUBE MANOMETER (MICRO-MANOMETER)

As shown below, the differential reading is proportional to the pressure difference. If the pressure
difference is very small, the reading may be too small to be measured with good accuracy. To
increase the sensitivity of the differential reading, one leg of the manometer can be inclined at an

21
angle θ, and the differential reading is measured along the inclined tube. As shown below,
h2  l 2 sin  , and hence

PA  PB   2 gl2 sin  3 gh3  1 gh1

Obviously, the smaller the angle θ, the more the reading l2 is magnified.

Example

For the inclined manometer containing mercury, shown in the figure below, determine the pressure
in pipe B if the pressure in pipe A is 10 kPa. Pipe A has water flowing through it, and oil is flowing in
pipe B.

22
TOTAL PRESSURE FORCE ON AN IMMERSED SURFACE

Since pressure is defined as force per unit area, then the force acting on a small area, δA will be

F  PA
As the fluid is at rest, the force will act at right-angles to the surface.

Consider a plane surface immersed in a liquid. The total area is made up of many elemental areas.
The force on each elemental area is always normal to the surface, but in general each force is of
different magnitude as the pressure usually varies.

F1=P1δA1

F2=P2δA2

Fn=PnδAn
δA1

δAn

δA2

All the forces on the small elements can be represented by a single force, called the resultant force,
acting at right angles to the plane through a point called the centre of pressure.

Thus, Resultant force, R = sum of forces on all elemental areas

R  P1A1  P2A2  .......... ..........  PnAn   PA

HORIZONTALLY SUBMERGED PLANE

For a horizontal plane submerged in a liquid (or a plane experiencing uniform pressure over its
surface), the pressure, P will be equal at all points of the surface.

Consider a plane horizontal surface immersed in a liquid. Let

A = area of immersed surface

x = depth of horizontal surface from the liquid

23
The total pressure force on the surface

F = weight of the liquid above the immersed surface = specific weight of liquid x volume of liquid
= specific weight of liquid x area of surface x depth of liquid

 F  gAx

VERTICALLY SUBMERGED SURFACE

Consider a plane vertical surface of arbitrary shape immersed in a liquid.

x
b
x
dx h

Let

A = total area of the surface

G = centre of the area of surface

x = depth of centre of area

h = distance of centre of pressure from free surface of liquid

C = centre of pressure

24
TOTAL PRESSURE FORCE, F

Consider a thin horizontal strip of the surface of thickness dx and breadth b. Let the depth of the
strip be x. Let the intensity of pressure on strip be P; this may be taken as uniform as the strip is
extremely small. Then P  gx

Total pressure force on the strip = PA=P.bdx = gx.bdx


Total pressure force on the whole area, F  gx.bdx  g bdx.x 
But  bdx.x = moment of the surface area about the liquid level = Ax

Therefore, F  gAx

Thus, the total pressure force on a surface is equal to the area multiplied by the intensity of pressure
at the centre of area of the plane.

The expression F  gAx holds for all surfaces whether flat or curved.

LOCATION OF CENTRE OF PRESSURE, h

The intensity of pressure on an immersed surface is not uniform, but increases with depth. As the
pressure is greater over the lower portion of the plane, the resultant pressure force on any
immersed surface will act at some point, below the centre of gravity of the immersed surface and
towards the lower edge of the object.

Let C be the centre of pressure of the immersed object

Let h = depth of centre of pressure below free liquid surface

Io = moment of inertia of the surface about free surface

Consider the horizontal strip of thickness dx. Total pressure force on strip = gx.bdx

Moment of this pressure force about free surface = gx.bdxx  gx 2 .bdx


Total moment of all such pressures for whole area, M  g x 2 .bdx

x .bdx  I o = moment of inertia of the surface about the free surface (or second moment of
2
But
area)

Therefore, M  gI o

The sum of the moments of the pressure force is also equal to F  h

Equating the two and solving for h

25
gI o I
h  o
gAx Ax

Also I o  I G  Ax 2

Where IG = moment of inertia of the surface about horizontal axis through its centre of gravity

IG
Therefore location of centre of pressure, h  x
Ax

CENTRE OF GRAVITY AND MOMENT OF INERTIA OF SOME IMPORTANT GEOMETRICAL FIGURES

NAME OF C.G FROM BASE AREA IG ABOUT AN AXIS I ABOUT BASE


FIGURE PASSING THROUGH C.G
AND PARALLEL TO THE
BASE
Triangle h bh bh 3 bh 3
x A IG  I
3 2 36 12
Rectangle d A  bd bd 3 bd 3
x IG  I
2 12 3
Circle
x
d d 2 d 4 _
A IG 
2 4 64
Trapezoid
x
h2a  b 
A
a  b h
IG 

h 3 a 2  4ab  b 2 
3a  b  2 36a  b 
Ellipse xh A  bh 
IG  bh3
y b 4
Semi-ellipse
x
4h
A

bh IG 
9  64 3
2
bh

3 2 72
y b
Parabolic 2 2 8
section x h A bh IG  bh 3
5 3 175
3
y b
8
Semicircle
x
4r
A
1 2
r IG 
9  64 4
2
r

3 2 72

h d d
x x
x

b b
Triangle Rectangle Circle

26
x

x
a

x h
h y y
C.G C.G
h
y
b b
b

Trapezoid Ellipse

x x

r y
y y
C.G
h

b b

Semi-ellipse Semicircle

x
h
C.G
y
y
b

Parabolic section

27
Example

A vertical isosceles triangular gate with its vertex up has a base width of 2 m and a height of 1.5 m. If
the vertex of the gate is 1 m below the free water surface, find the total pressure force and the
position of the centre of pressure.

1.0 m

1.5 m

2.0 m

Example

A sheet piling holds fresh water and salt water (s.g = 1.035) on either sides of it as shown in the
figure below. Find the moment about the base M of the resultant force per unit length of piling.

Salt water 3.5 m

2.5 m Fresh water

28
INCLINED IMMERSED SURFACE

Consider a plane of arbitrary shape, wholly submerged in a liquid in equilibrium. The plane of the
surface makes an angle θ with the horizontal, and the intersection of this plane with the plane of the
free surface (where the pressure is atmospheric) is taken as the x-axis. The y-axis is taken down the
sloping plane. Every element of the area is subjected to a force due to the pressure of the liquid.

h
x
h
y

A yp

G
P

At any element of area δA, at a depth h below the free surface, the pressure is P = ρgh and the
corresponding force is

F  PA  ghA

Because the area is inclined at an angle θ, it is convenient to work in the plane of the area, using y to
denote the position on the area at any depth h. H is related to y by the following expression:

h  y sin 

Where y is measured from the level of the free surface of the fluid along the angle of inclination of
the area. Then,

F  g  y sin  A

The total force on one side of the plane is therefore

F   g  y sin  dA  g sin   ydA

But  ydA is the first moment of area about the x-axis and may be represented by Ay .
29
Therefore,

F  g sin  Ay

But x  y sin 

 F  gAx

CENTRE OF PRESSURE

To find the pressure centre, the moment of the resultant force, F is equated to the moment of the
distributed forces about the y-axis and x-axis, respectively.

The moment of the distributed (small force) force dF with respect to the x-axis is

dM  dF  y

But dF  g  y sin  dA

Then dM  ygy sindA  g sin y 2 dA  


The moment of all the forces on the entire area is found by integrating over the area.

 
M   g sin  y 2 dA  g sin   y 2 dA

y
2
But dA is the second moment of area (moment of inertia) Io of the entire area with respect to
the x-axis.

Therefore,

M  gI o sin

Also moment about the x-axis is given by the resultant force, i.e.

M  F  yp

Equating the two expressions of moment,

F  y p  gI o sin 

gI o sin  gI o sin  I


 yp    o
F g sin   Ay Ay

Also I o  I G  Ay 2

Therefore,

30
I G  Ay 2 I G
yp   y
Ay Ay

In terms of the vertical depth the expression for the centre of pressure becomes,

I G sin 2 
h x
Ax

PROCEDURE FOR COMPUTING THE FORCE OF A SUBMERGED SURFACE

1. Identify the point where the angle of inclination of the area of interest intersects the level of
the free surface of the fluid. This may require the extension of the angled surface or the fluid
surface line.
2. Locate the centroid of the area from its geometry.
3. Determine x as the vertical distance from the level of the free surface down to the centroid
of the area.
4. Determine y as the inclined distance from the level of the free surface down to the centroid
of the area. Note that x and y are related by

x  y sin 

5. Calculate the total area A on which the force is to be determined.


6. Calculate the resultant force from

F  gAx

7. Calculate I G , the moment of inertia of the area about its centroidal axis.
8. Calculate the location of the centre of pressure from

IG
yp  y
Ay

9. If it is desired to compute the vertical depth to the centre of pressure, h either of the two
methods can be used. If the distance y p has already been computed, use

h  y p sin 

or h can be computed directly from

I G sin 2 
h x
Ax

31
Example

A sewer discharges to a river. At the end of the sewer is a circular gate with a diameter of 0.6 m. The
gate is inclined at an angle of 45o to the water surface. The top edge of the gate is 1.0 m below the
surface. Calculate

(a) The resultant force on the gate caused by the water in the river,
(b) The vertical depth from the water surface to the centre of pressure

45o
1.0m

0.6m

Exercise

A rectangular plane 1.2 m by 1.8 m is submerged in water and makes an angle of 30 o with the
horizontal, the 1.2 m sides being horizontal. Calculate the magnitude of the net force on one face
and the position of the centre of pressure when the top edge of the plane is

(a) At the free surface


(b) 500 mm below the free surface
(c) 30m below the free surface

FORCES ON A CURVED IMMERSED SURFACE

On a curved surface the forces F on individual elements differ in direction, and a simple
summation of them may not be made. Instead, the resultant forces in certain directions may be
determined, and these forces may then be combined vectorially. It is simplest to calculate horizontal
and vertical components of the total force.

CASE I

In the diagram below the liquid is resting on top of a curved base.

32
The element of fluid ABC is in equilibrium as the fluid is at rest.

HORIZONTAL FORCES

Considering the horizontal forces, none can act on CB as there are no shear forces in a static fluid so
the forces would act on the faces AC and AB as shown below.

It can be seen that the horizontal force on AC, FAC, must equal and be in the opposite direction to the
resultant force, RH on the curved surface.

As AC is the projection of the curved surface AB onto a vertical plane, it can be generalised that

RH = resultant force on the projection of the curved surface onto a vertical plane = gAx

Therefore, RH will act horizontally on the vertical plane through the centre of pressure of the
projection of the curved surface.

VERTICAL FORCES

The diagram below shows the vertical forces which act on the element of fluid above the curved
surface.

33
There are no shear forces on the vertical edges, so the vertical component can only be due to the
weight of the fluid. Thus,

RV = weight of fluid directly above the curved surface

This force will act vertically downward through the centre of gravity of the mass of fluid.

RESULTANT FORCE

The overall resultant force is found by combining the vertical and horizontal components vectorially.
Thus,

R  RH2  RV2

And this force acts at an angle θ. The angle the resultant force makes to the horizontal is

 RV 
  tan 1  
 RH 

Example

A surface consists of a quarter of a circle of radius 2.0 m. It is located with its top edge 1.5 m below
the water surface. For unit width of the immersed surface, calculate the magnitude and direction of
the resultant force on the upper surface.

1.5 m

2m

34
CASE II

The figure below shows a situation where there is a curved surface which is experiencing fluid
pressure from below.

The calculation of the forces acting from the fluid below is very similar to when the fluid is above.

HORIZONTAL FORCE

From the figure below it can be seen that the only two horizontal forces on the area of fluid, which is
in equilibrium, are horizontal reaction force which is equal and in the opposite direction to the
pressure force on the vertical plane AB.

Thus, the resultant horizontal force of a fluid below a curved surface is:

RH = resultant force on the projection of the curved surface on a vertical plane = gAx

VERTICAL FORCE

The vertical force acting is shown on the figure below. If the curved surface where removed and the
area were replaced by the fluid, the whole system would be in equilibrium. Thus, the force required
by the curved surface to maintain equilibrium is equal to that force which the fluid above the surface
would exert i.e. the weight of the fluid.

35
The resultant vertical force of a fluid below a curved surface is:

RV =weight of the imaginary volume of fluid vertically above the curved surface.

RESULTANT FORCE

The resultant force and direction of application are calculated in the same way as for fluids above
the surface:

R  RH2  RV2

The angle the resultant force makes to the horizontal is

 RV 
  tan 1  
 RH 

EFFECT OF A PRESSURE ABOVE THE FLUID SURFACE

If an additional pressure exists above the fluid or if the fluid itself is pressurised, the effect is to add
to the actual depth a depth of fluid ha .i.e.

P
ha 
g

FORCES ON CURVED SURFACES WITH FLUID ABOVE AND BELOW

The diagram shows a semi-cylindrical gate projecting into a tank containing a liquid. The force due to
fluid pressure would have a horizontal component acting to the right of the gate. This force acts on
the projection of the surface on a vertical plane and is computed in the same manner as before.

36
h

In the vertical direction, the force on the top of the gate would act downward and would equal the
weight of the liquid above the gate. However, there is also a force acting upward on the bottom
surface of the gate equal to the total weight of the fluid, both real and imaginary, above that
surface. The net vertical force is the difference between the two forces, equal to the weight of the
semi-cylindrical volume of fluid displaced by the gate.

Fdown

Fup Fnet

37
PRACTICAL APPLICATIONS OF HYDROSTATICS

WATER PRESSURE ON A DAM

Dams are constructed in order to store large quantities of water, for the purpose of irrigation and
power generation. A dam may be of any cross-section, but the following are important:

(i) Rectangular dams


(ii) Trapezoidal dams

WATER PRESSURE ON RECTANGULAR DAMS

Consider a rectangular dam retaining water on one of its sides as shown below:

G
H

H/3
F W R

Let H = height of water retained by the dam

x = depth to the centroid

Total pressure force, F  gAx

But A=Hx1 (unit width)

gH 2
Therefore, F
2

And this force will act at a height H/3 above the base of the dam.

Let W be the height of the dam masonry per unit length of the dam. W will act downwards through
the centre of gravity of the dam.

Thus, W = mg

Resultant pressure force is given by

R  F 2 W 2

And the inclination of the resultant with the vertical is given by:

38
F
  tan 1  
W 

Example

A retaining wall 6 m high and 2.5 m wide retains water up to its top. Find the total pressure force per
metre length of the wall and the point at which the resultant cuts the base. Also find the resultant
thrust on the base of the wall per metre length. Assume weight of masonry as 23 kN/m 3.

2.5 m

6m G

WATER PRESSURE ON TRAPEZOIDAL DAMS

Consider a dam with a trapezoidal cross-section

H
G
h
θ
F

W R

b
heel

Let a = top width of the dam

b = base width of the dam

39
H = height of the dam

h = height of water column

L = length of dam = 1 m

ab
Weight of masonry, W  mg  g    H 1
 2 

Let the centre of gravity of the section be at a distance x from the vertical face. Now dividing the
trapezium into a rectangle and a triangle and taking moments about the heel on the vertical face, we
get

a 2  ab  b 2
x
3a  b 

gh 2
Total pressure force, F
2

This pressure acts at h/3 from the base of the dam. The direction of action is given by

F
  tan 1  
W 

Example

A concrete dam of trapezoidal section having water on vertical face is 15 m high. The base of the
dam is 10 m wide and top 3 m wide. Find the resultant thrust on the base per metre length of dam,
and the point where it intersects the base. Take the specific weight of masonry as 24 kN/m 3 and
water level coinciding with the top of the dam.

3m

15 m
G
θ

W R

10 m

40
BUOYANCY AND FLOTATION

The resultant force exerted on a body by a static fluid in which it is submerged or floating is called
the buoyant force. The buoyant force always acts vertically upward through the centre of buoyancy.

The familiar laws of buoyancy (Archimedes’ principle) and flotation are usually stated as follows:

(a) A body immersed in a fluid is buoyed (lifted) up by a force equal to the weight of fluid
displaced.
(b) A floating body displaces its own weight of the fluid in which it floats.

The buoyant force on a submerged body is the difference between the vertical component of
pressure force on its underside and the vertical component of pressure force on its upper side.

F E

C
A

In the figure above, the upward force on the bottom is equal to the weight of liquid, real or
imaginary, which is vertically above the surface ABC, indicated by the weight of liquid within
ABCEFA. The downward force on the upper surface equals the weight of liquid ADCEFA. The
difference between the two forces is a force, vertically upward, due to the weight of fluid ABCD that
is displaced by the solid. In the equation form

FB   f Vd

Where FB = Buoyant force

 f = Specific weight of the fluid

Vd = Displaced volume of the fluid

The same formula holds for floating bodies and Vd is taken as the volume of liquid displaced. When
a body is floating freely, it displaces a sufficient volume of fluid to just balance its own weight.

41
The analysis of problems dealing with buoyancy requires the application of the equation of static
equilibrium in the vertical direction, F v  0 , assuming the object is at rest in the fluid. The
following procedure is recommended for all problems, whether they involve floating or submerged
bodies.

PROCEDURE FOR SOLVING BUOYANCY PROBLEMS

1. Determine the objective of the problem solution. Are you to find a force, a weight, a volume,
or a specific weight?
2. Draw a free-body diagram of the object in the fluid. Show all forces that act on the free-body
in the vertical direction, including the weight of the body, the buoyant force, and all external
forces. If the direction of some force is not known, assume the most probable direction and
show it on the free-body.
3. Write the equation of static equilibrium in the vertical direction, F v  0 , assuming the
positive direction to be upward.
4. Solve for the desired force, weight, volume, or specific weight, remembering the following
concepts:
a) The buoyant force is calculated from FB   f Vd .
b) The weight of a solid object is the product of its total volume and its specific weight;
that is w  V .
c) An object with an average specific weight less than that of the fluid will tend to float
because w  FB with the object submerged.
d) An object with an average specific weight greater than that of the fluid will tend to
sink because w  FB with the object submerged.
e) Neutral buoyancy occurs when a body stays in a given position wherever it is
submerged in a fluid. An object whose average specific weight is equal to that of the
fluid is neutrally buoyant.

Example

Two cubes of the same size, 1 m3, one of specific gravity 0.80, the other of specific gravity 1.1, are
connected by a short wire and placed in water. What portion of the lighter cube is above the water
surface, and what is the tension in the wire?

s.g = 0.80

s.g = 1.1

42
Exercise

How many kg of concrete,   25kN / m3 , must be attached to a beam having a volume of 0.1 m 3
and specific gravity 0.65 to cause both to sink in water?

THE STABILITY OF BODIES IN FLUIDS

When a body floats in vertical equilibrium in a liquid, the forces present are the upthrust FB acting
through the centre of buoyancy and the weight of the body w  mg acting through its centre of
gravity. For equilibrium, FB and w must be equal and act in the same straight line.

FB

The equilibrium of a body may be stable, unstable or neutral, depending upon whether, when given
a small displacement, it tends to return to the equilibrium position, move further from it or remain
in the displaced position.

STABILITY OF SUBMERGED BODIES

A body in a fluid is considered stable if it will return to its original position after being displaced a
small amount.

The condition for stability of bodies completely submerged in a fluid is that the centre of gravity, G,
of the body must be below the centre of buoyancy, B.

If the two points coincide, the submerged body is in neutral equilibrium for all positions. If, on the
other hand, the centre of gravity is above the centre of buoyancy, the moment created when the
body is tilted would produce an overturning moment that would cause a body to overtop.

STABILITY OF FLOATING BODIES

The condition for the stability of floating bodies is different from that for completely submerged
bodies.

The stability is determined by the forces acting when it has been disturbed from the position of
static equilibrium.

The figure below shows a body floating in a fluid.

43
In figure (a), the floating body is at its equilibrium orientation and the centre of gravity, G, is above
the centre of buoyancy, B. A vertical line, called the vertical axis, passes through these points. Figure
(b) shows that if the body is rotated slightly, the centre of buoyancy shifts to a new position B’
because the geometry of the displaced volume has changed. The buoyant force and the weight
produce a righting couple that tends to return the body to its original orientation. Thus, the body is
stable.

The condition for stability is that the centre of gravity should be below the metacentre. The
metacentre (M) is the intersection of the vertical axis of a body when in its equilibrium position and
a vertical line through the new position of the centre of buoyancy, B’, when the body is rotated
slightly.

It is possible to determine analytically if a floating body is stable by calculating the location of its
metacentre. The distance to the metacentre from the centre of buoyancy is called MB and is
calculated from

I
MB 
Vd

Where, Vd = displaced volume of fluid

I = least moment of inertia of a horizontal section of the body taken at the surface of
the fluid.

PROCEDURE FOR EVALUATING THE STABILITY OF FLOATING BODIES

1. Determine the position of the floating body, using the principles of buoyancy.
2. Locate the centre of buoyancy, B; Compute the distance from some reference axis to B,
called ycb. Usually, the bottom of the object is taken as the reference axis.
3. Locate the centre of gravity, G; Compute ycg measured from the same reference axis.
4. Determine the shape of the area at the fluid surface and compute the smallest moment of
inertia I for the shape.
5. Compute the displaced volume Vd.

44
I
6. Compute MB 
Vd
7. Compute y mc  ycb  MB
8. If y mc  y cg , the body is stable
9. If y mc  y cg , the body is unstable

The distance of the metacentre above centre of gravity is known as the metacentric height and for
stability of the body it must be positive (i.e. M above G). Thus

GM  MB  BG  y mc  y cg

Example

A uniform, closed cylindrical buoy, 1.5 m high, 1.0 m diameter, and of mass 80 kg is to float with its
axis vertical in sea water of density 1026 kg/m3. A body of mass 10 kg is attached to the centre of the
top surface of the buoy. Show that, if the buoy floats freely, initial instability will occur.

B
h

45

You might also like