You are on page 1of 34

TOPIC T1: MASS, MOMENTUM AND ENERGY AUTUMN 2006 (REVISED)

Objectives

(1) Extend the basic principles of mass and momentum to flows with non-uniform velocity
and pressure.
(2) Extend the continuity principle to time-varying flows.
(3) Apply continuity and Bernoulli’s eqn for flow measurement and tank-emptying problems.
(4) Learn methods for quantifying losses.

Contents

0. Revision of basic concepts


0.1 Notation
0.2 Dimensionless parameters
0.3 Definitions
0.4 Basic principles of fluid mechanics
0.5 Physical constants
0.6 Properties of common fluids

1. Continuity (conservation of mass)


1.1 Mass and volume fluxes
1.2 Flows with non-uniform velocity
1.3 Time-dependent flow

2. Forces and momentum


2.1 Control-volume formulation of the momentum principle
2.2 Fluid forces
2.3 Boundary layers and flow separation
2.4 Drag and lift coefficients
2.5 Calculation of momentum flux
2.6 Calculation of pressure forces
2.7 The wake-traverse method for measurement of drag

3. Energy
3.1 Bernoulli’s equation
3.2 Fluid head
3.3 Static and stagnation pressure
3.4 Flow measurement
3.5 Tank emptying
3.6 Summary of methods for incorporating losses

References
Hamill (2001) – Chapters 1, 2, 4, 5, 7
Chadwick and Morfett (2004) – Chapters 1, 2, 3
Massey (1998) – Chapters 1, 2, 3, 4
White (2002) – Chapters 1, 2, 3

Hydraulics 2 T1-1 David Apsley


0. REVISION OF BASIC CONCEPTS

0.1 Notation

Geometry
x ≡ (x, y, z) position; (z is usually vertical)
t time
Field Variables
u ≡ (u, v, w) velocity (also use V for average velocity in a pipe or conduit)
p pressure
p – patm is the gauge pressure
p* = p + gz is the piezometric pressure
T temperature
Fluid Properties
density
 dynamic (or absolute) viscosity
 ≡ /
 kinematic viscosity
 ≡ g specific weight (weight per unit volume)
s.g. ≡ / ref specific gravity (or relative density);
“ref” = water (for liquids) or air (for gases)
 surface tension (force per unit length)
K bulk modulus (pressure change divided by volumetric strain)
c speed of sound
k conductivity (heat flux per unit area divided by temperature gradient)

0.2 Dimensionless Parameters

UL UL
Re ≡ ≡ Reynolds1 number (viscous flow)
 

U
Fr ≡ Froude2 number (open-channel hydraulics)
gL
U
Ma ≡ Mach3 number (compressible flow)
c
U 2L
We ≡ Weber4 number (surface tension)


fL
St = Strouhal5 number (vortex shedding; f = shedding frequency)
U

Here, U is a representative velocity scale and L is a representative length scale. There are
many other important dimensionless combinations (see Topic T3: “Dimensional Analysis”).

1
Osborne Reynolds (1842-1912); appointed first Professor of Engineering at Owens College (precursor of the
University of Manchester).
2
William Froude (1810-1879), British naval architect; developed scaling laws for the model testing of ships.
3
Ernst Mach (1838-1916), Austrian physicist and philosopher.
4
Moritz Weber (1871-1951), developed modern dimensional analysis; actually named the Re and Fr numbers.
5
Vincenz Strouhal (1850-1922), Czech physicist; investigated the “singing” of wires.

Hydraulics 2 T1-2 David Apsley


0.3 Definitions

A fluid is a substance that continuously deforms under a shear stress, no matter how small.
A solid will reach equilibrium under such a stress.

Fluids may be liquids (definite volume; free surface) or gases (expand to fill any container).

Hydrostatics is the study of fluids at rest.


Hydrodynamics is the study of fluids in motion.

Hydraulics is the study of the flow of liquids (usually water).


Aerodynamics is the study of the flow of gases (usually air).

All fluids are compressible to some degree, but their flow can be approximated as
incompressible (that is, hydrodynamic pressure changes don’t give rise to density changes)
for velocities much less than the speed of sound (∼ 1480 m s–1 in water, 340 m s–1 in air).

An ideal fluid is one with no viscosity. It doesn’t exist, but it can be a good approximation.

A Newtonian fluid is one for which viscous stress is proportional to velocity gradient (rate of
strain):
du
= 

dy
where is the viscosity. Most fluids of interest in hydraulics (including air and water) are


Newtonian, but there are some important non-Newtonian fluids (e.g. mud, blood, paint).

Real flows may be laminar (adjacent layers slide smoothly over each other) or turbulent
(subject to random fluctuations about a mean flow). If the viscosity is too small to maintain a
smooth, orderly flow, then a laminar flow undergoes transition and becomes turbulent.
Although transition to turbulence is dependent on a number of factors, including surface
roughness, the primary determinant is the Reynolds number
UL UL
Re ≡ ≡ (1)
 

U and L are typical velocity and length scales of the flow. In general:
“high” Re turbulent

“low” Re laminar


Typical critical Reynolds numbers for transition are:


pipe flow: ReD ≈ 2300;
circular cylinder: ReD ≈ 3×105;
flat plate: Rex ≈ 5×105 – 3×106.
Important. The Reynolds number and its critical value depend on which velocity and length
scale are used to define it – which should be stated. For example, you could choose to use
either radius or diameter for flow in a pipe, and they would obviously give a factor-of-2
difference in Reynolds number even though the flow is the same. (Why are the values quoted
for circular cylinder and flat plate above so much larger than that for pipe flow?)

The vast majority of civil-engineering and environmental flows have high Reynolds numbers
and are fully turbulent.

Hydraulics 2 T1-3 David Apsley


0.4 Basic Principles of Fluid Mechanics

Hydrostatics

In stationary fluids, pressure forces balance weight.

Hydrostatic Principle
dp
p=− g z or =− g (2)


dz

The same equation holds in a moving fluid if there is no vertical component of acceleration.

For a constant-density fluid the hydrostatic equation can also be written


p + gz = constant


p + gz is called the piezometric pressure, p*. It represents the combined effect of pressure
and weight.

Thermodynamics

For compressible fluids thermodynamics and heat input are important and one requires, in
addition, an equation of state; e.g.

Ideal Gas Law


p = RT (3)

The gas constant R is a constant for any particular gas and is given by R = R0 / m, where R0 is
the universal gas constant and m is the mass of one mole. For dry air, R = 287 J kg–1 K–1.

T is the absolute temperature in Kelvin:


T (K ) = T (°C) + 273.15 (4)

Fluid Dynamics

Continuity (Mass Conservation)


Mass is neither created nor destroyed.
For steady flow, flow in = flow out

Momentum Principle
Force = rate of change of momentum.
For steady flow, force = (momentum flux)out – (momentum flux)in
Energy
Change in energy = heat supplied + work done
For incompressible flow, change of kinetic energy = work done

Hydraulics 2 T1-4 David Apsley


The Energy Equation

(i) Incompressible Fluids

For incompressible fluids the energy equation is a purely mechanical equation, equivalent to,
and directly derivable from, the Momentum Principle.

Bernoulli’s Equation

Steady incompressible flow without losses:


p
+ gz + 12 U 2 = constant along a streamline

More generally, in steady incompressible flow,


p
( + gz + 12 U 2 ) = work done on fluid (5)

Here, ( ) means “change in” and the RHS of (5) represents the energy (per unit mass) input
by pumps or removed by turbines or friction.

(ii) Compressible Fluids

For compressible fluids the energy equation involves thermodynamics. The energy per unit
mass is supplemented by the internal energy e and energy can also be transferred to the fluid
as heat. (5) becomes
p
(e + + gz + 12 U 2 ) = heat supplied to fluid + work done on fluid (6)

The quantity e + p/ is called enthalpy.




0.5 Physical Constants

Gravitational acceleration: g = 9.81 m s–2 (at British latitudes)


Universal gas constant: R0 = 8314 J kg–1 K–1
Standard atmospheric pressure: 1 atmosphere = 1.01325×105 Pa = 1.01325 bar

Standard temperature and pressure (s.t.p):


IUPAC (International Union of Pure and Applied Chemistry): 0° C (273.15 K) and 105 Pa.
ISO (International Standards Organisation): 0° C (273.15 K) and 1 atm (1.01325×105 Pa).
Since two major international organisations can’t agree it’s probably better to specify
reference conditions explicitly.

Hydraulics 2 T1-5 David Apsley


0.6 Properties of Common Fluids
Properties are given at 1 atmosphere and 20 ºC unless otherwise specified.

Air
Density: = 1.20 kg m–3 ( = 1.29 kg m–3 at 0 ºC)
Specific weight:  = 11.8 N m–3
Dynamic viscosity:  = 1.80×10–5 kg m–1 s–1 (or Pa s)
Kinematic viscosity:  = 1.50×10–5 m2 s–1
Specific heat capacity at constant volume: cv = 718 J kg–1 K–1
Specific heat capacity at constant pressure: cp = 1005 J kg–1 K–1
Gas constant: R = 287 J kg–1 K–1
Speed of sound: c = 343 m s–1

Water
Density: = 998 kg m–3 ( = 1000 kg m–3 at 0 ºC)
Specific weight:  = 9790 N m–3
Dynamic viscosity:  = 1.003×10–3 kg m–1 s–1 (or Pa s)
Kinematic viscosity:  = 1.005×10–6 m2 s–1
Surface tension:  = 0.0728 N m–1
Speed of sound: c = 1482 m s–1

Mercury
Density: = 13550 kg m–3

Ethanol
Density: = 789 kg m–3

Fluid properties – especially viscosity – change significantly with temperature.

Water Air
T (°C) (kg m–3) (Pa s)  (m2 s–1) (kg m–3)  (Pa s)  (m2 s–1)
0 1000 1.788×10–3 1.788×10–6 1.29 1.71×10–5 1.33×10–5
20 998 1.003×10–3 1.005×10–6 1.20 1.80×10–5 1.50×10–5
50 988 0.548×10–3 0.555×10–6 1.09 1.95×10–5 1.79×10–5
100 958 0.283×10–3 0.295×10–6 0.946 2.17×10–5 2.30×10–5

As temperature increases:
viscosities of liquids decrease;
viscosities of gases increase.
(Explain why.)

The viscosity of gases may be approximated by Sutherland’s law:


3/ 2
T T0 + S
= (7)
0 T0 T +S
For air: T0 = 273 K, 

0 = 1.71×10–5 Pa s, S = 110.4 K.

Hydraulics 2 T1-6 David Apsley


1. CONTINUITY (CONSERVATION OF MASS) mout

The rate at which something crosses a surface is called its flux. “Mass flux”
(or “mass flow rate”) is the mass crossing a given surface per unit time. m
Conservation of mass can be applied to the fluid in or passing through the
surface of an arbitrary control volume: min

Steady flow: (mass flux)in = (mass flux)out

d
Unsteady flow: (mass ) = (mass flux) in − (mass flux) out
dt

1.1 Mass and Volume Fluxes

If velocity u is uniform over a section and normal to area A, then


A
Volume flux: Q = uA (m3 s–1) u
Mass flux: m = Q = uA (kg s–1)

Q is also called the quantity of flow, (volumetric) flow rate or discharge.

A stream tube is a bundle of streamlines across which there is no flow.

For an incompressible flow, is constant along a streamline and mass conservation implies
volume conservation; i.e.:
Q1 = Q 2 1
2 u2
u1
If u is uniform over the cross-section then
u1 A1 = u 2 A2

If u is not uniform or if there is more than one inlet or outlet, then


total flow in = total flow out
Q in = Q out
and the total volume flux or mass flux must be obtained by summation or, for continuously-
varying quantities, by integration.

Hydraulics 2 T1-7 David Apsley


Example.
The figure shows a converging two-dimensional duct in which flow enters in two layers. A
fluid of specific gravity 0.8 flows as the top layer at a velocity of 2 m s–1 and water flows
along the bottom layer at a velocity of 4 m s–1. The two layers are each of thickness 0.5 m.
The two flows mix thoroughly in the duct and the mixture exits to atmosphere with the
velocity uniform across the section of depth 0.5 m.

0.5 m 2 m/s

0.5 m

0.5 m 4 m/s

p1=15 kN/m2

(a) Determine the velocity of flow of the mixture at the exit.


(b) Determine the density of the mixture at the exit.
(c) If the pressure p1 at the upstream section is 15 kPa, what is the force per unit width
exerted on the duct? (Do this part after Section 2 on the Momentum Principle).

Answer: (a) 6 m s–1; (b) 933 kg m–3; (c) 7.8 kN

1.2 Flows With Non-Uniform Velocity

The continuity principle may be extended to cases where u varies over a cross-section (e.g.
flow in pipes or flow in a boundary layer) by considering the flow to be broken down into
infinitesimal areas dA, across each of which the velocity is approximately constant:
dQ = u dA

The total quantity of flow is found by summation or, in the limit of small areas, integration:

Volume flow rate: Q= u dA (8)

The average velocity (sometimes called the bulk velocity) is that constant velocity which
would give the same total flow rate; i.e. Q = u av A or
Q
Average velocity: u av = (9)
A

In other words, to find the average velocity from a non-uniform velocity profile you will first
have to find Q.

Hydraulics 2 T1-8 David Apsley


b
1.2.1 Two-Dimensional Flow
u(y)
dy
Velocity is a function of height:
u ≡ u(y)
This is often the case in a wide rectangular channel.

A small element of area over which the velocity is uniform has the form of a rectangle, width
b and height dy:
dA = b dy

Quantity of flow: Q = b u dy (10)

or:

Flow per unit width: q= u dy (11)

Example.
The distribution of velocity in a rectangular channel of width b = 800 mm and depth h = 200
mm is given by
1
y 7
u = u0
h
–1
where u0 = 8 m s . What is (a) the quantity of flow; (b) the average velocity?

Solution.
(a)
h
y
Q= u dA = b u 0 ( )1 / 7 dy (b = 0.8 m, h = 0.2 m, u0 = 8 m s–1)
0 h
Simplify the integral with a change of variables Y = y / h, dY = dy / h :
1 1
7 8/7 7
Q = u 0 bh Y 1 / 7 dY = u 0 bh × Y = u 0 bh
0 8 0 8
−1
= 1.12 m s
3

(b)
flow rate Q
u av = =
area bh
7
= u0
8
= 7 m s −1

Hydraulics 2 T1-9 David Apsley


u(r)
1.2.2 Axisymmetric Flow
r dr
Velocity is a function of radius:
u ≡ u (r )
Examples include pipes and jets.

A small element of area over which the velocity is uniform has the form of an infinitesimal
hoop of radius r, thickness dr:
dA = 2 r dr (12)

Quantity of flow: Q= u 2 r dr


(13)

Example.
Fully-developed laminar flow in a pipe of radius R has velocity profile:
u = u 0 (1 − r 2 /R 2 )
Find the average velocity in terms of u0.

Solution.
The average velocity can be found by dividing the flow rate by the area. For the flow rate,
R R

Q= u 2 r dr = 2
 

u 0 (1 − r 2 / R 2 )r dr
0 0
Substitute s = r / R, ds = dr / R for convenience:
1 1 1
s2 s4
Q = 2 R u0

2
(1 − s ) s ds = 2 R u 0
2 
2
( s − s ) ds = 2 R u 0
3 
2

0 0 2 4 0

1 2
= R u0 

2
Hence,
Q
u av =
R2 

1
= u0
2

Note. For velocity profiles measured in an experiment (where the integral must usually be
evaluated graphically), it is unnatural and inaccurate to have the integrand vanishing at the
centre (since this is where velocity is highest) and (13) can be rewritten as
Q= 

u dr 2 (14)

i.e.
Quantity of flow = 

× (area under a u – r2 graph)

See, for example, the pipe-flow laboratory experiment and the Example Sheet.

Hydraulics 2 T1-10 David Apsley


1.3 Time-Dependent Flow

Examples of time-dependent flows are:


• fluid oscillations; e.g. pressure transients in pipes, vortex shedding, waves;
• moving-boundary problems; e.g. pistons, rotating machinery.

The mass of fluid (m) inside a control volume can vary if either the density of fluid or the size
of the control volume changes. To balance this there must be a net mass flux through the
boundaries of the control volume. Thus, in any time interval
change of mass = mass that has flowed in – mass that has flowed out
or, in rate form,
dm
= (mass flux ) in − (mass flux) out (15)
dt

For incompressible flows, conservation of mass implies conservation of volume:


dV
= Qin − Qout (16)
dt

Example. (White, 2002) An incompressible fluid is being squeezed outwards between two
large circular discs by the uniform downward motion V0 of the upper disc. Assuming 1-
dimensional radial outflow, derive an expression for V(r).
V0

r
h(t) V(r)

Solution: For an incompressible fluid, conserving mass is equivalent to conserving volume.


Hence, using a control volume consisting of a cylinder of height h and radius r:
dV
= Qin − Qout
dt
d
( r 2 h) = 0 − (2 rh)V
 

dt
dh


r2 = −2 rhV


dt
r dh
V =−
2 h dt
dh
But = −V0 . Hence,
dt
Vr
V= 0
2h

Hydraulics 2 T1-11 David Apsley


2. FORCES AND MOMENTUM

Momentum Principle
Force = rate of change of momentum (17)

In principle this is a vector equation, but, in practice, often only one direction is relevant.

An ideal fluid is one without viscosity. Ideal fluids don’t exist, but can be a useful
approximation. The momentum equation for an ideal fluid is often called the Euler equation6.

The momentum equation for a real fluid is often called the Navier7-Stokes8 equation.

2.1 Control-Volume Formulation of the Momentum Principle

The equation of motion (17) can be expressed mathematically in many ways, including partial
differential equations, velocity potential (a bit like gravitational potential), vorticity (related to
local angular momentum), or even in terms of complex variables (see White, 2002).
u out
Fortunately, in hydraulics it is usually adequate to work from first principles
by considering the momentum balance for a control volume (CV).

For a steady flow and fixed control volume: F


force = rate of change of momentum u in
mass × change in velocity
=
time
= Q(u out − uin )

This is OK if the velocity at inflow and outflow are uniform, but not very useful if they vary.
Instead we define

Momentum flux = mass flux × velocity = Qu (18)

Momentum Principle For Steady Flow

Force = (momentum flux)out – (momentum flux)in (19)

= (rate at which momentum leaves CV) – (rate at which momentum enters CV)

6
Leonhard Euler (1707-1783), Swiss mathematician, later Professor of Physics at the St Petersburg Academy;
tackled many problems in fluid mechanics and mathematical physics.
7
Claude Navier (1785-1836), French civil engineer; also known for his strong political views, including
opposition to Napoleon’s military aggression.
8
George Gabriel Stokes (1819-1903), Irish mathematician and Lucasian Professor of Mathematics at
Cambridge; many important works in hydrodynamics.

Hydraulics 2 T1-12 David Apsley


Example.
A jet of fluid flows smoothly onto a stationary curved vane which turns it through 45°. The
initial jet has diameter 40 mm and uniform velocity 25 m s–1. The exit jet may be assumed to
have uniform velocity 20 m s–1. Calculate the net force on the vane.

20 m/s

45
40 mm
25 m/s

Solution.
Gauge pressures are zero at inlet, outlet and free surface so that there is no net pressure force.
Let the force on the vane be F = ( Fx , Fy ) . Then the reaction of the vane on the fluid is
− F = (− Fx ,− Fy ) . A suitable control volume cuts incident and deflected jets where the flow is
uniform.

Mass flux:


D2
Q = u1
4


× 0.04 2
= 1000 × 25 × = 31.42 kg s -1
4
Momentum in x direction:
− Fx = Q(u 2 cos 45° − u1 )
1
= 31.42 × (20 × − 25) = −341.2 N
2
Momentum in y direction:
− F y = Q(u 2 sin 45° − 0)


1
= 31.42 × (20 × ) = 444.3 N
2

Net force on the vane, (Fx,Fy) = (341, – 444) N, or 560 N at angle 52.5° to the horizontal.

If there is more than one inflow or outflow then the net momentum flux must be obtained by
summation. In particular, for non-uniform flows it is necessary to work out momentum fluxes
(Section 2.5) and fluid forces (Section 2.6) by summation or integration.

Hydraulics 2 T1-13 David Apsley


2.2 Fluid Forces

The total force on the fluid in a control volume is a combination of:


• body forces (proportional to amount of fluid); e.g.
– weight;
• surface forces (proportional to area); e.g.
– pressure forces;
– viscous forces;
• reactions from solid boundaries.

(Note that weight acts irrespective of whether the fluid is moving or not and would be
balanced by a reaction, or, internally, by a hydrostatic pressure distribution. It can, therefore,
be excluded from the analysis if we consider only departures from this hydrostatic state.)

These are the only forces we shall consider here. However, other fluid forces exist; e.g.
– Coriolis forces in a rotating frame (e.g in meteorology);
– surface tension.

As in structural mechanics, surface forces are usually expressed in terms of stresses:


force
stress = (20)
area
or
force = stress × area

Pressure (p) is a normal stress directed inward to a surface. For


the control volume shown the net pressure force in the x direction
from pressures on the left (l) and right (r) faces is p lA prA
pl A − p r A A
Since the net effect of a uniform pressure on all boundaries is
∆x
zero it does not matter whether absolute, gauge or other relative
pressure is used, provided that one is consistent.

Shear stresses ( ) act tangentially to surfaces. For the control


volume shown the net force in the x direction from shear stresses on A τt A
the top (t) and bottom (b) faces is
tA− bA ∆y
What we have called above is strictly xy: for complex flows other τ A
b
components ( xz, xx, yy, …) may be important. By convention, xy is
the force per unit area in the x direction that the fluid on the upper (greater y) side of the
interface exerts on the fluid on the lower (smaller y) side.

Laminar and Turbulent Flow

Shear stresses arise from two sources: viscous forces and, in turbulent flow, the net transfer
of momentum across an interface by turbulent fluctuations (which, as far as the mean flow is
concerned, has the same effect as a force).

Hydraulics 2 T1-14 David Apsley


For Newtonian fluids, viscous stress is proportional to velocity gradient. If y
velocity u is a function of y only then, in laminar flow:
du τ
= (21)
dy
This defines the dynamic viscosity, . (In more complex flow fields a more


general stress-strain relationship is required.)


u
In turbulent flow one is usually only interested in the mean velocity u . Since momentum
transfer between fast- and slow-moving fluid is dominated by the net effect of turbulent
fluctuations rather than viscous stresses the mean shear stress is not equal to du /dy .

2.3 Boundary Layers and Flow Separation

The ideal-fluid (zero-viscosity) approximation is inapplicable if viscous effects have a major


effect on the flow. The most important example is boundary-layer separation.

In real fluids velocity vanishes at solid boundaries; (the no-slip condition). This gives rise to a
boundary layer close to walls where velocity changes rapidly from its value in the free
stream to zero at the boundary. At high Reynolds numbers boundary layer are usually
extremely thin.

In an adverse pressure gradient (where pressure increases and velocity decreases in the
direction of flow; for example, in an expanding channel) the net force in the opposite
direction to flow actually causes the more-slowly-moving fluid near the boundary to reverse
direction. This backflow leads to flow separation.

speeds up ... ... slows down a dv


ers
e
gra pres
die sure
nt

bac
kflo
w

flow separation

Hydraulics 2 T1-15 David Apsley


Turbulence in the boundary layer helps to prevent or delay flow separation because it readily
transports fast-moving fluid from the free stream into the near-wall region, maintaining
forward motion.

For bluff bodies with sharp corners flow separation occurs


at all but the smallest Reynolds numbers and causes a large
increase in pressure (or form) drag. For more streamlined
bodies with convex boundaries separation may or may not
occur.

2.4 Drag and Lift Coefficients


lift F
The force on a body in a flow can be
resolved into streamwise and cross-stream U0 drag
components.

Drag = component of force parallel to the approach flow.


Lift = component of force perpendicular to the approach flow.

The relative importance of drag or lift forces is quantified by non-dimensionalising them by


dynamic pressure ( 12 U 02 ) × area:

Drag Coefficient
drag
cD = (22)
1
2


U 02 A

A is a “representative” area which depends on the body geometry and the nature of the flow
(see below). Like the Reynolds number it should always be specified when defining cD.

A lift coefficient may be defined in the same manner.

Bluff and Streamlined Bodies

Bluff bodies (flow separation)


• force is predominantly pressure drag U0
• A is the projected area (normal to the flow) A
• cD = O(1)

Streamlined bodies (no flow separation)


• force is predominantly viscous drag U0
• A is the plan area (parallel to the flow) A
• cD << 1

Hydraulics 2 T1-16 David Apsley


2.5 Calculation of Momentum Flux

The momentum principle for steady flow may be written for a general control volume:
Force = (momentum flux) out − (momentum flux) in
where
momentum flux = mass flux × velocity
= ( Q)u (23)
= ( uA)u

If velocity is not constant then the momentum flux can be calculated (as for mass flux) by
breaking the surface into small areas over each of which the velocity is uniform. In particular,
if the velocity varies continuously over a cross-section then the sum of contributions from
infinitesimal areas can be obtained by integration; e.g. for the x-component:
momentum flux = u 2 dA (24)

Special Cases

Velocity profile Momentum flux

(i) Uniform
U 2A
Area A

b
(ii) 2-dimensional u(y)
dy
b u 2 dy
dA = b dy

(iii) Axisymmetric u(r)


r dr
u 2 2 r dr
dA = 2 r dr


Note. As for the mass flux, for experimental measurements and graphical integration rather
than theoretical work it is usually more appropriate (and accurate) to write the last of these as
Axisymmetric flow momentum flux = 


u 2 dr 2

= × × (area under a u 2 − r 2 graph)




Hydraulics 2 T1-17 David Apsley


Example. (Examination, January 2003)
A two-dimensional beam of height h = 100 mm completely spans a square air-conditioning
duct of height D = 400 mm (see Figure). The approach flow is uniform (u1 = 0.6 m s–1),
whilst the downstream velocity profile is 2-dimensional and may be represented by:
3 1 y
u 2 ( − cos ) ( y < 2h )
u= 4 4 2h
u2 ( y ≥ 2h )
The pressure is uniform over the height of the duct at both sections. Neglecting drag on the
walls of the duct find:
(a) the value of u2;
(b) the difference between pressures at inlet and downstream sections, assuming that
Bernoulli’s equation holds outside the wake region;
(c) the force on the beam.
Also,
(d) define a suitable drag coefficient for the beam and calculate its value.

Take the density of air as 1.2 kg m–3.

0.6 m/s
400 mm

100 mm

Solution.
(a) Continuity (per unit width):
2h D
3 1 y
u1 D = u dy = u 2 ( − cos ) dy + u 2 dy
0 4 4 2h 2h
wake
2h
3 1 2h y 3
= u2 y− sin + D − 2h = u 2 ( h + D − 2 h )
4 4 2h 0 2
1
= u 2 ( D − h)
2
Hence
D 400
u 2 = u1 = 0 .6 × = 0.6857 m s −1
D − 12 h 350

Answer: u2 = 0.69 m s–1

Hydraulics 2 T1-18 David Apsley


(b) By Bernoulli’s equation:
p1 + 12 u12 = p 2 + 12 u 22
p1 − p 2 = 1
2


(u 22 − u12 ) = 12 × 1.2 × (0.6857 2 − 0.6 2 ) = 0.06611 Pa

Answer: Pressure difference = 0.066 Pa

(c) Momentum principle (for steady flow):


force = rate of change of momentum = (momentum flux)out – (momentum flux)in
p1 D 2 − p 2 D 2 − F = D u 2 dy − u12 D 2
wake

1
F = D 2 p1 − p 2 + (u12 − 

u 2 dy )
D
wake

Now
2h D
3 1 y 

u 2
dy = u ( − cos ) 2 dy +
2
2 u 22 dy
0 4 4 2h 2h
wake
2h
1 y 

y 

=u 2
2 (9 − 6 cos + cos 2 ) dy + D − 2h
16 0 2h 2h
2h
1 

y 1 1 y 

= u 22 (9 − 6 cos + + cos ) dy + D − 2h
16 0 2h 2 2 h
2h
1 19 2h y 1h y  

=u 2
2 y − 6 sin + sin + D − 2h
16 2 2h 2 

h 

19
= u 22 ( h + D − 2h)
16
13
= u 22 ( D − h)
16
Hence
13 h
F = D 2 p1 − p 2 + [u12 − u 22 (1 − )]
16 D
13 1
= 0.4 2 0.06611 + 1.2 0.6 2 − 0.6857 2 (1 − × ) = 0.007759 N
16 4

Answer: Force on the beam = 0.0078 N

(d) Drag coefficient:


F 0.007759
cD = 1 2 = = 0.90
2 u1 Dh
1
2 × 1.2 × 0.6 2 × 0.4 × 0.1

Answer: Drag coefficient = 0.90

Hydraulics 2 T1-19 David Apsley


Example. (Examination, January 2004)
A long T-shaped element, of depth h / 4 and oriented symmetrically as shown below,
completely spans a wind-tunnel duct of depth 2h, where h = 0.2 m. The velocity upstream is
uniform: U0 = 40 m s–1. The velocity distribution is measured at a position downstream and is
found to be
2
1 y2
U ( y ) = U max 1− 1− 2
2 h
where y is the distance from the centreline.

y
h

U0 U(y)

(a) Sketch the expected pattern of flow around the T-shaped element, indicating, in
particular, separation and reattachment points, recirculating flow regions and the
direction of flow.
(b) Calculate Umax.
(c) If the pressure drop from the upstream to the downstream section is 200 Pa, find the
force per unit span on the T-shaped element.
(d) Define a suitable drag coefficient for the element and calculate its value.

Take the density of air as 1.2 kg m–3.

Answer: (b) 54.6 m s–1; (c) 36.5 N; (d) 0.76

Example. (Examination, January 2002)


Water enters a horizontal pipe of diameter 20 mm with uniform velocity 0.1 m s–1 at point A.
At point B some distance downstream the velocity profile becomes fully-developed and
varies with radius r according to:
u = u 0 (1 − r 2 /R 2 )
where R is the radius of the pipe. The pressure drop between A and B is 32 Pa.

(a) Find the value of u0.


(b) Calculate the total drag on the wall of the pipe between A and B.
(c) Beyond point B the pipe undergoes a smooth contraction to a new diameter DC.
Estimate the diameter DC at which the flow would cease to be laminar.

[The critical Reynolds number for transition in a circular pipe, based on average velocity and
diameter is 2300. Take the density and kinematic viscosity of water as = 1000 kg m–3 and
= 1.1×10–6 m2 s–1 respectively.]

Answer: (a) 0.2 m s–1; (b) 0.0090 N; (c) 15.8 mm

Hydraulics 2 T1-20 David Apsley


2.6 Calculation of Pressure Forces
p
When pressure changes with position the total pressure force can also be
found by summing over small contributions dA
p dA
Be very careful about direction in specific situations.

Example.
Find the hydrostatic force per unit width and the centre of
y s
pressure on a plane wall inclined at to the vertical.

h
Solution. L
The water depth is h and the total inclined length is L. Local
depth y and inclined distance s are defined in the diagram: θ
y = s cos 

From hydrostatics the gauge pressure is


p = gy = gs cos
 

The total pressure force can be found by summing over contributions from small inclined
areas dA = ds × 1 . The force per unit width f is then
f = p dA
L
=


gs cos ds
s =0

= 1
gL2 cos


The centre of pressure, s , is found by equating moments (force × distance, or p dA × s):


f ×s = s × p dA
L
1
2 gL2 cos × s = gs 2 cos ds
s =0

= 1
3 gL3 cos
s = 23 L

Notes.
(i) All pressures are relative to a convenient reference – here, atmospheric pressure.

(ii) The total force per unit width can be written:


f = ( 12 gL cos ) × L


= 1
2


gh × L
= average pressure × area ( per unit width)

Hydraulics 2 T1-21 David Apsley


(iii) The x component of total force (per unit width) is
f cos = 12 gh × h = average pressure × depth
The y component of total force is
f sin = 12 gh × L sin = ( 12 hL sin ) g = weight of water above
These are general results, even for curved surfaces, as can be seen by balancing forces
on the volume of fluid above the plate.

(iv) The centre of pressure, on the line of action of the resultant, is at 2/3 of total depth.

The pressure force on a small element is directed normal to that element. The pressure force
on the 2-d element of length ds shown has individual x and y components
df x = ( p ds ) cos
= p dy θ
= pressure × projected area normal to the x direction ds dy
df y = ( p ds ) sin
= p dx
dx
= pressure × projected area normal to the y direction

Thus, for inclined surfaces, one can find components of force by either:
calculating the force, then taking components in x and y directions;
or
multiplying pressure by x and y components of area.

The latter is more convenient for curved surfaces. e.g. on


pU(x)
aerofoil sections the pressure force (per unit span) in the
upward y direction is
p L ( x) dx − pU ( x) dx = ( p L − pU ) dx
pL(x)
lower surface upper surface
x
This can be used to calculate the lift force.

2.7 The Wake-Traverse Method for Measurement of Drag

Objects in a flow experience a Force on BODY Force on FLUID


hydrodynamic force. If the fluid
exerts a force F on the body then
F
F
the body exerts a reaction force
–F on the fluid. By measuring
the change in momentum and
pressure one can use the momentum principle to deduce the force on the body.

Suitable control volumes for constrained (e.g. wind tunnel) and unconstrained flow are shown
below. Upper and lower boundaries are streamlines, across which there is no mass or
momentum flux.

Hydraulics 2 T1-22 David Apsley


(i) constrained (change in free-stream velocity)

Fluid passing close to a body forms a


body
wake of low-velocity downstream. If
inflow wake the flow is constrained by boundaries
then the velocity outside the wake
must increase slightly, with a
compensating fall in pressure. In the
unconstrained case, upper and lower
boundaries should be sufficiently far
(ii) unconstrained (no change in free-stream velocity)
away that pressure is equal to that in
the free stream.
body
inflow wake

streamline

The steady-state momentum principle gives:


force = (momentum flux)out – (momentum flux)in
−F + Pin dA − p dA = u 2 dA − U in2 dA (25)
in out out in
Since the inflow velocity is uniform (Uin) the last term can be converted to an integral over
the outflow by using continuity:
U in2 dA = U in U in dA = U in u dA = U in u dA
in in out out
Substituting in (25) and rearranging gives
F= [ u (U in − u ) + ( Pin − p)] dA (26)
out

For unconstrained flows, pressures upstream and downstream are equal, as are the free-
stream velocities: Uin = Uout. In the constrained case, it can be shown that, provided the wake
is narrow (so that the difference in free-stream velocities is small), any pressure difference
approximately9 compensates for the change in free-stream velocity U . In both cases, then,
F= u (U ∞ − u ) dA (27)
out
Thus, hydrodynamic forces may be deduced indirectly by measuring momentum deficit in the
wake, rather than directly using a force balance. This is called the wake traverse method and
you will have an opportunity to use it in the wind-tunnel laboratory experiment.

9
By Bernoulli, Pin − Pout = 1
2

2
(U out − U in2 ) ; the error in (27) is 1
2


(U in − U out ) 2 dA , which is second


order in the (small) free-stream velocity difference Uin – Uout.

Hydraulics 2 T1-23 David Apsley


3. ENERGY

3.1 Bernoulli’s Equation

Consider a thin stream tube, with varying cross-sectional area 1


A. In the absence of thermal effects the energy of fluid passing 2
u1 u2
through it is changed by the work done by pressure and
viscous forces from the adjacent fluid and energy supplied or
removed by external agents (e.g. pumps and turbines).

Since the sides are locally parallel to the flow, energy only flows in or out of the control
volume through ends 1 and 2. Hence, for steady flow,
(rate of energy passing 2) − (rate of energy passing 1) = rate of doing work (28)
The mechanical energy consists of kinetic energy (½mU2) and potential energy (mgz).

The rate at which forces do work (i.e. power) is given by force × velocity (in direction of
force). The rate of working of pressure at end 1 is therefore (pAU)1 and at end 2 is –(pAU)2.
where U is flow speed. Pressure does no work on the sides because force and velocity are
perpendicular there.

(28) can then be written


Q( 12 U 2 + gz ) 2 − Q( 12 U 2 + gz )1 = ( pAU )1 − ( pAU ) 2 + W
where W consists of the rate of working of friction forces (i.e. losses) and pumps/turbines
etc. Q = AU is the mass flow rate, which must be constant along the stream tube. Dividing
by Q = AU gives
p p W
( 12 U 2 + gz ) 2 − ( 12 U 2 + gz )1 = ( )1 − ( ) 2 +
Q
or, rearranging,

Bernoulli’s Equation With Losses


p
( + 12 U 2 + gz ) = (29)

∆ means “change in” and is the work done per unit mass of fluid passing through this
length of stream tube.

Notes.
(i) represents all non-pressure work done on the system. It is composed of frictional
losses (due to viscosity) and any work done on or by the flow via pumps, turbines etc.

(ii) Each term in the equation represents some form of energy or work done per unit
mass.

(iii) (29) can easily be extended to thermal flows (boilers, condensers, refrigerators, ...) by
adding the internal energy e to the LHS and the rate of heat input QH to the RHS.

Hydraulics 2 T1-24 David Apsley


(iv) If there are no losses and no external sources of energy then (29) reduces to
Bernoulli’s equation10:
p 1 2
+ 2 U + gz = constant (along a streamline) (30)

For incompressible flows, is also constant along a streamline and hence this
equation is often applied as
p + 12 U 2 + gz = constant (along a streamline) (31)
Note the assumptions:
• inviscid (no losses)
• incompressible
• steady
• along a streamline (different streamlines may have a different “constant”)

Example. Water is emptied from a tank through a horizontal pipe with centreline h below the
level of water in the tank. The pipe has a severe constriction where the diameter is D1 and the
water exits to the atmosphere through a nozzle of diameter D2 (> D1).

h
D1 D2

(1) (2)

(a) Assuming no losses, formulate an expression for the gauge pressure at the constricted
section (1) in terms of diameters D1 and D2 and the tank level h.

(b) Cavitation (i.e. the formation of bubbles of vapour) will occur if the absolute pressure
falls below the vapour pressure for water (2.3 kPa at 20 °C). If the tank level h = 2 m
and the constriction diameter D1 = 25 mm, calculate the exit diameter D2 at which this
begins to occur. To avoid cavitation should you increase or decrease D2?

(Take atmospheric pressure = 101 kPa.)

D2 4
Answer: (a) p1 − p atm = ρgh[1 − ( ) ] ; (b) 39 mm; decrease D2.
D1

10
Daniel Bernoulli (1700-1782) Swiss-Dutch mathematician, a member of an illustrious family of well-known
mathematicians.

Hydraulics 2 T1-25 David Apsley


3.2 Fluid Head

In Bernoulli’s equation:
( p + gz + 12 U 2 ) = energy change per unit volume
 

Each term has dimensions of pressure, or of energy per unit volume. If one divides by the
specific weight g these become energies per unit weight:
p U2
( +z+ ) = change in fluid head
g 2g (32)
= energy change per unit weight

Energy per unit weight has dimensions of length and is called fluid head.
p
= pressure head
g
U2
= dynamic head
2g
p U2
+z+ = total (or available) head
g 2g
In hydraulics, energy and pressure are both often expressed in length units; e.g. “metres of
water” or “millimetres of mercury”.

Losses due to friction and the capabilities of pumps are typically specified in terms of head;
that is, work done per unit weight. For the latter the rate of working (i.e. power) is given by
power = gQH


(33)

where Q is the quantity of flow and H is the change in head. These will be examined further
in Topic 2 (Pipe flow) and Topic 4 (Pumps).

3.3 Static and Stagnation Pressure


stagnation point
A stagnation point is a point on a streamline where the P0 = P + 12 ρU 2 (highest pressure)
U=0
velocity is reduced to zero. In general, any non-rotating P = P0
solid obstacle in a stream produces a stagnation point
next to its upstream surface, where the flow streamlines
must split to pass around the obstacle.

The stagnation pressure (or Pitot pressure) p0 is that


pressure which would arise if the flow were brought instantaneously to rest. By Bernoulli’s
equation it is given (for incompressible fluids) by p + 12 U 2 . We define:
stagnation pressure p + 12 U 2


static pressure p
dynamic pressure 1
2
U2

The dynamic pressure (and hence the flow velocity) is found by the difference between
stagnation and static pressures (see the wind-tunnel laboratory experiment).

Hydraulics 2 T1-26 David Apsley


3.4 Flow Measurement

3.4.1 Measurement of Pressure – Manometry Principles

Three basic rules apply in a stationary fluid:


(1) Same fluid, same height same pressure
(2) Same fluid, different height p=− g z 

(3) Different fluids: pressure is continuous at an interface

A B
U-Tube Manometer y
By rule (1) the pressure at level C is the same in both arms of the manometer.
By rules (2) and (3) it can be found from pA and pB respectively by summing h
the changes in pressure over the heights of columns of fluid:
left arm right arm C
p A + g (h + y ) = pC = p B + gy + m gh

Cancelling gy and subtracting gh gives the pressure difference p = p A − pB :


Differential Manometer Equation

p=( 

m − ) gh


(34)

If the working fluid is a gas then  

m and (34) can be approximated by


p = m gh

(35)

Inclined Manometer
)
arge
L (l
Differences in head may be small and difficult to measure h (small)
accurately. The movement of the manometer fluid may be θ
amplified by inclining the manometer. It is the vertical difference
in height which is proportional to pressure differences: this is given in terms of the much
larger length L by
h = L sin


(36)

Hydraulics 2 T1-27 David Apsley


3.4.2 Measurement of Velocity

Basic idea: bring the fluid to rest at one point and measure the difference between static and
stagnation (Pitot) pressures:
p0 − p = 1


2
U2
Pitot static dynamic
pressure pressure pressure

Examples.
2
U /2g
(1) Open channel flow. free surface

stagnation point

piezometer Pitot tube

(2) Pitot tube and piezometer U2


2g

(3) Pitot-static tube – measures both stagnation and static pressures in the same instrument.

total head tube


static head tube

static holes
stagnation point

Hydraulics 2 T1-28 David Apsley


3.4.3 Measurement of Quantity of Flow

Venturi Flowmeter 1
2
A venturi is a localised smooth constriction in a duct.

In the throat region:


U increases (by continuity)
p decreases (by Bernoulli)
The difference in pressure between the main flow and
the throat can be measured by a differential manometer and converted to quantity of flow.

Bernoulli: p1 + 12 U 12 = p 2 +


1
2


U 22
p1 − p 2 = 12 (U 22 − U 12 )


There are two unknowns on the RHS but this can be reduced to one (the velocity U1 which is
required) by using continuity:
U 1 A1 = U 2 A2
A
U 2 = U1 1
A2
Hence, substituting for U2 in Bernoulli’s equation:
A
p = 12 U 12 ( 1 ) 2 − 1


A2
Rearranging for U1 and taking Q = U1A1,
1/ 2
2 A12


p
Q= 

(37)
( A1 /A2 ) 2 − 1

Q = constant × ( p)1 / 2


(38)

Thus, the flow rate can be found by measuring the pressure difference p.

In accurate experiments a coefficient of discharge, cd, is included to represent departures from


non-ideality. cd is the ratio of the actual flow rate to the ideal flow rate which would be
computed from Bernoulli’s equation and continuity:

Q = c d Qideal (39)

Design Features
• A large convergence angle is advantageous as it tends to make the flow more uniform.
• A small divergence angle is necessary to prevent flow separation.
• The throat must be long enough for parallel flow to be established.
• For a well-designed flowmeter a typical value of the coefficient of discharge is ~ 0.98.

Hydraulics 2 T1-29 David Apsley


Orifice Flowmeter

An orifice is an aperture of negligible streamwise thickness through


which fluid passes. vena contracta

The streamwise thickness determines frictional losses.

The fluid cannot turn immediately, so that the emerging stream tube
continues to contract up to the vena contracta – the section of
minimum area.

An orifice meter is a means of measuring the flow rate in a


duct by measuing the differential pressure across an orifice.

It is basically an extreme variant of the venturi meter with the


divergent region omitted. The basic premise is that the pressure
throughout the recirculating eddy is essentially equal to that at
the vena contracta.

Advantage: cheap.
Disadvantage: considerable loss of energy.

By the same process as that for the venturi meter one obtains:
1/ 2
2 A12


p
Q= 

( A1 /Av ) 2 − 1
where Av is the area of the vena contracta. Av is not obvious from the geometry. If Av is
replaced by the area of the orifice then this may be compensated for by a coefficient of
discharge, but, in practice, theory is simply used to deduce the form of relationship between


flow rate Q and pressure drop p:


Q = constant × ( p)1 / 2


(40)
with the constant of proportionality determined by calibration.

Hydraulics 2 T1-30 David Apsley


Example (Massey). A vertical venturi meter carries a liquid of
relative density 0.8 and has inlet and throat diameter of
150 mm and 75 mm respectively. The pressure connection at
the throat is 150 mm above that at the inlet. If the actual rate
of flow is 40 L s–1 and the coefficient of discharge is 0.96 B
0.075 m
calculate (a) the pressure difference between inlet and throat;
(b) the difference in levels in a vertical U-tube manometer
connected between these points, the tubes above the mercury A
being full of the liquid. (Relative density of mercury = 13.56) 0.15 m
C 0.04 m3/s

Solution.
(a) The difference in static pressure comes from Bernoulli’s equation under ideal conditions:
( p + 12 lV 2 + l gz ) A = ( p + 12 lV 2 + l gz ) B
 

p A − p B = 12 l (VB2 − VA2 ) + l g ( z B − z A ) (*)


The velocities to be used in (*) come from the ideal flow rate, which is derived from the
actual flow rate via the coefficient of discharge CD = Q/Qideal:
Q Q/C D
V = ideal 2
=
D /4 D 2/4
3 –1
With Q = 0.04 m s , CD = 0.96, DA = 0.15 m, DB = 0.075 m, this gives
VA = 2.358 m s–1
VB = 9.431 m s–1
The density of the liquid is
−3
l = 0.8 × 1000 = 800 kg m


Hence, substituting in (*) gives:


l (VB − VA ) = 2 × 800 × (9.431 − 2.358 ) = 33350 Pa

2 2 1 2 2

l g (zB − z A ) = 800 × 9.81 × 0.15 = 1177 Pa


And hence
p A − p B = 33350 + 1 177
= 34527 Pa

(b) The height difference h in the U-tube manometer can be established by equating at the
common height C the pressures found by applying the hydrostatic law in the two arms of the
manometer:   

p B + l g ( z B − z C − h) + Hg gh = pC = p A + l g ( z A − z C )
Hence,
( Hg − l ) gh = ( p B + l gz B ) − ( p A + l gz A )
   

= 1
2


l (VB2 − V A2 )
= 33350 Pa
Then,
33350 33350
h= =
( Hg − l ) g (13560 − 800) × 9.81
 

= 0.266 m

Hydraulics 2 T1-31 David Apsley


3.5 Tank Emptying

For an emptying reservoir, discharging as a free jet, with free 1


surface at a distance h above the discharging fluid, apply
Bernoulli’s equation between the free surface and the jet:
   

h
p1 + 12 U 12 + gz1 = p 2 + 12 U 22 + gz 2
Here, p1 = p 2 = p atm , U 1 = 0, z1 − z 2 = h , so that
2
1
2 U 22 = g ( z1 − z 2 ) = gh

Hence, we have
Torricelli’s Formula11
U exit = 2 gh (41)

If the aperture is large (compared with the height to the free surface), then this value will be
different for each streamline passing through the orifice (because each will have a different
value of h). The total discharge would then have to be found by integration.

Ideally the discharge would be


Qideal = 2 gh × (area of orifice) (42)
This is not true in practice because of
(i) frictional effects (small for a sharp-edged orifice)
(ii) contraction (area of vena contract < area of orifice).
If these are significant then a coefficient of discharge cd may be introduced to compensate:
Q = c d Qideal
cd must be measured experimentally. For a sharp-edged orifice, cd ≈ 0.6 – 0.65.

Contraction effects can be reduced by using a bellmouth exit to minimise rapid changes in
direction. However, frictional losses are then greater.

Submerged Orifice

If the discharge is not a free jet but into a reservoir of the


same fluid it is called a submerged orifice. In Bernoulli’s
h1 formula p2 is not then atmospheric, but given by p 2 = gh2 . 

h2
Torricelli’s formula then reads
U exit = 2 g (h1 − h2 ) (43)

Bernoulli’s formula is invalid in the reservoir/orifice problem if the free-surface level


changes rapidly (since it is then time-dependent). If, however, the tank cross section is much
larger than that of the orifice, then a quasi-steady approximation is OK and, by equating the

11
Evangelista Torricelli (1608-1647); Italian mathematician of barometer fame; served as secretary to Galileo.

Hydraulics 2 T1-32 David Apsley


rate at which the volume of fluid in the tank decreases to the rate of discharge from the
orifice,
dV
= −Q (44)
dt
we can find how long it takes to drain the tank.

Example.
A cylindrical tank of base diameter 0.5 m is used to store water. A rupture at the base of the
tank allows water to escape through an aperture of area 8 cm2. A discharge coefficient of 0.6
can be assumed for this orifice. If the depth of water in the tank is initially 0.8 m, how long
does it take to empty the tank?

Solution.
The volume of water in the tank is that of a cylinder of base diameter D = 0.5 m and
(variable) height h. Its volume is, therefore,
D2
V= h
4
This volume is reduced at a rate equal to the flow rate through the aperture, i.e.
Q = c d Qideal = c d U ideal Aexit
where cd = 0.6, Aexit = 8×10-4 m2 and U ideal = 2 gh by Torricelli’s formula.

Hence,
dV
= −Q
dt
d D2
( h) = −c d Aexit 2 gh
dt 4
D 2 dh
= − 2 gh
4c d Aexit dt
D2
h −1 / 2 dh = − dt
4c d Aexit 2 g
Integrating between t = 0 (where h = h0 = 0.8 m) and emptying time T (where h = 0):
0 T
D2
2h 1/ 2
=− t
4c d Aexit 2 g h0 0

2


D h0
− = −T
2c d Aexit 2g
Hence,


D2 h0 

× 0.5 2 0.8
T= = −4
×
2c d Aexit 2g 2 × 0.6 × 8 × 10 2 × 9.81
= 165 s

Hydraulics 2 T1-33 David Apsley


3.6 Summary of Methods For Incorporating Losses

Many theoretical results are derived for ideal fluids, assuming no frictional losses, simplified
geometry and uniform velocity profiles.

In practice, compensation is necessary for non-ideal flow. These include the following

Discharge coefficients
Correct the quantity of flow deduced by Bernoulli’s equation:
Q = c d Qideal

Loss coefficients
Quantify energy or pressure losses in conduits:
V2
H = −K
2g
or, equivalently (and with the same K):
P = − K ( 12 V 2 )


L
(e.g. pipe friction: K =  , where is the friction factor).


Momentum and energy coefficients


Account for non-uniform velocity profiles when computing total flux.
u 2 dA = ( u av2 A)


momentum flux,  

energy flux, 

u 3 dA = ( u av
3


A) 

Hydraulics 2 T1-34 David Apsley

You might also like