You are on page 1of 32

TOPIC T1: MASS, MOMENTUM AND ENERGY AUTUMN 2013

Objectives

(1) Extend continuity and momentum principles to non-uniform velocity.


(2) Apply continuity and Bernoulli’s equation to flow measurement and tank-emptying.
(3) Learn methods for dealing with non-ideal flow.

Contents

0. Basics
0.1 Notation
0.2 Dimensionless groups
0.3 Definitions
0.4 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 Conservation of mass
1.3 Flows with non-uniform velocity

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 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 filling and emptying
3.6 Summary of methods for incorporating non-ideal flow

References
Hamill (2011) – Chapters 1, 2, 4, 5, 7
Chadwick and Morfett (2013) – Chapters 1, 2, 3
Massey (2011) – Chapters 1, 2, 3, 4
White (2011) – Chapters 1, 2, 3

Hydraulics 2 T1-1 David Apsley


0. BASICS

0.1 Notation

x  (x, y, z) position; (z is usually vertical)


t time
u  (u, v, w) velocity; (V is often used for the average speed in a pipe or channel)
p pressure
p – patm is the gauge pressure;
p* = p + ρgz is the piezometric pressure
T temperature

Fluid properties:
ρ density (mass per unit volume)
γ  ρg specific weight (weight per unit volume)
s.g.  ρ/ρref specific gravity (or relative density);
“ref” = water (for liquids) or air (for gases)
μ dynamic viscosity (stress per unit velocity gradient)
ν  μ/ρ kinematic viscosity
σ surface tension (force per unit length)
K bulk modulus (pressure change divided by fractional change in volume)
k conductivity of heat (heat flux per unit area per unit temperature gradient)
c speed of sound

0.2 Dimensionless Groups

ρUL UL
Re   Reynolds1 number (viscous flow)
μ ν
U
Fr  Froude2 number (open-channel hydraulics)
gL
U
Ma  Mach3 number (compressible flow)
c
ρU 2 L
We  Weber4 number (surface tension)
σ

U and L are representative length and velocity scales. You should always state which ones
you are using for a particular flow (e.g. average velocity and diameter for pipe flow).

Topic T3 (Dimensional Analysis) will introduce other important dimensionless groups.

1
Osborne Reynolds (1842-1912); appointed first Professor of Engineering at Owens College (later to become
the University of Manchester); his apparatus is on display in the basement of the George Begg building.
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.

Hydraulics 2 T1-2 David Apsley


0.3 Definitions

A fluid is a substance that flows; i.e. continues to deform under a shear stress.
A solid will deform initially but then reach static 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 (i.e. pressure changes due to motion don’t cause significant 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:
du
τμ
dy
μ is the viscosity. Most fluids of interest (including air and water) are Newtonian, but there
are some important non-Newtonian ones (e.g. mud, blood, paint, polymer solutions).

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
ρUL UL
Re   (1)
μ ν
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 between laminar and turbulent flow are:
pipe flow: ReD  2300 (based on diameter and average velocity)
circular cylinder: ReD  3105 (based on diameter and approach velocity)
flat plate: Rex  5105 – 3106 (based on distance and free-stream velocity)

Important: The Reynolds number and its critical value depend on the particular flow and on
which velocity and length scale are chosen to define it: see the different values above. The
particular choices should, therefore, be stated. (For example, you could choose to use either
radius or diameter as the length scale for flow in a pipe, and they would 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 majority of civil-engineering flows are fully turbulent.

Hydraulics 2 T1-3 David Apsley


0.4 Principles of Fluid Mechanics

Hydrostatics

In stationary fluid, pressure forces balance weight. Hence, pressure increases with depth.

Hydrostatic Equation
dp
Δp  ρgΔz or  ρg (2)
dz

Pressure also varies in the same way with depth in a moving fluid if there is no vertical
component of acceleration or, as a good approximation, if the vertical acceleration << g.

For a constant-density fluid these can also be written


Δ( p  ρgz )  0 or p  ρgz  constant
p + ρgz is called the piezometric pressure.

Thermodynamics

For compressible fluids thermodynamics and heat input are important and one requires, in
addition, an equation of state connecting pressure, density (or volume) and temperature; e.g.

Ideal Gas Law

p = ρRT (3)

p is the absolute pressure, whilst T is the absolute temperature in Kelvin:


T (K)  T (C)  273.15 (4)
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. The
equation is a rearrangement of the form used in chemistry: pV  nR0T , where V is volume
and n is the number of moles of gas.

Fluid Dynamics

Continuity (Mass Conservation)


Mass is conserved.
For steady flow: (mass flux)in = (mass flux)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 Flow

For incompressible flow the energy equation is a purely mechanical equation and can be
derived from the momentum principle.

Bernoulli’s Equation

Steady incompressible flow without losses:


p  ρgz  12 ρU 2  constant a long a streamline

More generally, in steady incompressible flow,


p
Δ(  gz  12 U 2 )  work done ( per unit mass) by non  conservative forces (5)
ρ

Δ( ) means “change in” and the RHS of (5) represents energy input to the flow by pumps or
removed from the flow by turbines or friction.

(ii) Compressible Flow

For compressible flow 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) is replaced by:
p
Δ(e   gz  12 U 2 )  heat supplied to fluid  work done on fl uid (6)
ρ
The quantity e  p/ρ is called enthalpy.

0.5 Physical Constants

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


Universal gas constant: R0 = 8.314 J mol–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).

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: ρ = 1.20 kg m–3 (ρ = 1.29 kg m–3 at 0 ºC)
Specific weight: γ = 11.8 N m–3
Dynamic viscosity: μ = 1.8010–5 kg m–1 s–1 (or Pa s)
Kinematic viscosity: ν = 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: ρ = 998 kg m–3 (ρ = 1000 kg m–3 at 0 ºC)
Specific weight: γ = 9790 N m–3
Dynamic viscosity: μ = 1.00310–3 kg m–1 s–1 (or Pa s)
Kinematic viscosity: ν = 1.00510–6 m2 s–1
Surface tension: σ = 0.0728 N m–1
Speed of sound: c = 1482 m s–1

Mercury
Density: ρ = 13550 kg m–3

Ethanol
Density: ρ = 789 kg m–3

Fluid properties – especially viscosity – change significantly with temperature.

Water Air
T (°C) ρ (kg m–3) μ (Pa s) ν (m2 s–1) ρ (kg m–3) μ (Pa s) ν (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
μ  T   T0  S 
    (7)
μ 0  T0   T  S 
The constants differ between gases. For air: T0 = 273 K, μ0 = 1.7110–5 Pa s, S = 110.4 K.

Hydraulics 2 T1-6 David Apsley


1. CONTINUITY (CONSERVATION OF MASS)

1.1 Mass and Volume Fluxes

Mass flux or mass flow rate is the mass crossing a given surface per unit time.

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

Volume flux: Q  uA (m3 s–1)


A
u
–1
Mass flux: ρQ  ρuA (kg s )

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


flowout

1.2 Conservation of Mass control


volume
Conservation of mass can be applied to the fluid in, or passing
through, an arbitrary control volume: flowin

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

d
Unsteady flow: (mass)  (mass flux) in  (mass flux) out
dt

If the density is uniform and constant then these can be applied to volume as well as mass.

In steady flow: 1
Q1  Q2 2
u1 u2
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
 uA
in
  uA
out

Unsteady flows will be considered later in the context of the tank-emptying problem.

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 part (c) after Section 2 on the Momentum Principle).

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

1.3 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 breaking the section down into infinitesimal
areas dA, across each of which the velocity is 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 (or bulk velocity) is that constant velocity which would give the same
total flow rate; i.e. Q  u av A or
Q flow rate
Average velocity: u av   (9)
A area

To find the average velocity for a non-uniform velocity profile first find the flow rate Q.

Hydraulics 2 T1-8 David Apsley


1.2.1 Two-Dimensional Flow w
u(y)
Velocity is a function of one Cartesian coordinate: dy
u  u(y)
This is often the case in a wide channel.

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

Two-Dimensional Flow

Quantity of flow: Q  w u dy (10)

or

flow per unit width: q   u dy (11)

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

Solution.
(a)

h
Q   u dA  w
y 1/ 7
 u 0 ( ) dy (w = 0.8 m, h = 0.2 m, u0 = 8 m s–1)
 0 h
To simplify the integral make a change of variables: Y = y/h, dY = dy/h. Then
1 1
 7  7
Q  u 0 wh Y 1 / 7 dY  u 0 wh   Y 8 / 7   u 0 wh
0 8 0 8
 1.12 m 3 s 1

(b)
flow rate Q 7
u av    u0
area wh 8
1
 7 ms

Hydraulics 2 T1-9 David Apsley


1.2.2 Axisymmetric Flow u(r)
r dr
Velocity is a function of the radial coordinate r:
u  u(r )
Examples include pipes and jets.

A small element of area over which the velocity is uniform is an infinitesimal hoop of radius
r and thickness dr:
dA  2πr dr (12)
Axisymmetric Flow

Quantity of flow: Q   u 2πr dr (13)

Example.
Fully-developed laminar flow in a pipe of radius R has velocity profile:
u  u0 (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 2πr dr  2π u 0 (1  r 2 /R 2 )r dr
0 0
For convenience, make a change of variables: s = r/R, ds = dr/R . Then
1

1

1
s2 s4 
Q  2πR u 0  (1  s ) s ds  2πR u 0  ( s  s 3 ) ds  2πR 2 u 0   
2 2 2

0 0 2 4 0
1
 πR 2 u 0
2
Hence,
Q
u av 
πR 2
1
 u0
2

Note: For velocity profiles measured in an experiment (where the integral would be
evaluated graphically as the area under a curve), it is unnatural and inaccurate to have the
integrand (u 2πr) vanishing at the centre, r = 0, since this is where velocity is highest; in this
case (13) can be rewritten, making the change of variables s = r2, ds = 2r dr, as
 
Q  π u ds or Q  π u dr 2 (14)
 
i.e.
quantity of flow = π  (area under a u vs r2 graph)

Hydraulics 2 T1-10 David Apsley


2. FORCES AND MOMENTUM

Momentum Principle

Force = rate of change of momentum (15)

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

An ideal fluid is one without viscosity. No such fluid actually exists, but it can be a useful
approximation. The momentum equation for an ideal fluid is called the Euler equation5. The
momentum equation for a real fluid is called the Navier6-Stokes7 equation.

2.1 Control-Volume Formulation of the Momentum Principle

The equation of motion (15) can be expressed mathematically in many ways, including partial
differential equations, but for many problems in hydraulics it can be applied by considering
the change of momentum as fluid passes through a control volume.

For a steady flow, fixed control volume and uniform inflow and outflow velocities:
force  rate of change of momentum
uout
mass entering  change in velocity

time
 mass flux  change in velocity
i.e. F
F  ρQ(uout  uin ) u in

This mathematical expression – as used in Hydraulics 1 – is fine when the inflow and outflow
velocities are uniform, but not when they vary over a cross-section. More generally, define:
Momentum flux = mass flux  velocity = ρQu (uniform flow) (16)
or  ρQu (non-uniform flow)

Momentum Principle For Steady Flow

Force = (momentum flux)out – (momentum flux)in (17)

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

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. Before tackling this there follows a brief,
qualitative discussion of fluid forces and important flow phenomena.

5
Leonhard Euler (1707-1783), Swiss mathematician, later Professor of Physics at the St Petersburg Academy;
tackled many problems in fluid mechanics and mathematical physics.
6
Claude Navier (1785-1836), French civil engineer; also known for his strong political views, including
opposition to Napoleon’s military aggression.
7
George Gabriel Stokes (1819-1903), Irish mathematician and Lucasian Professor of Mathematics at
Cambridge; many important works in hydrodynamics.

Hydraulics 2 T1-11 David Apsley


2.2 Fluid Forces

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


 body forces (proportional to volume); e.g.
– weight;
– centrifugal and Coriolis forces8 (apparent forces in rotating frames);
 surface forces (exerted by adjacent fluid and proportional to area); e.g.
– pressure forces;
– viscous forces;
 reactions from solid boundaries.

Weight acts whether the fluid is moving or not and would be balanced by a hydrostatic
pressure distribution. It can be excluded from the analysis if we consider only departures
from the hydrostatic state and work with the piezometric pressure.

Since surface forces are proportional to area they are usually expressed in terms of stresses:
force
stress  or force  stress  area (18)
area

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


the control volume shown the net pressure force in the x pLA pRA
direction from pressures on the left (L) and right (R) faces is A
( pL  pR ) A
Since the net force only involves the difference in pressures it does not matter whether
absolute, gauge or other relative pressure is used, as long as one is consistent.
A T A
Shear stresses (τ) act tangentially to surfaces. For the control
volume shown the net force in the x direction from shear stresses on
the top (T) and bottom (B) faces is BA
τT A  τ B A
(What is denoted τ here is strictly τxy; in complex flows other components such as τxx, τyz, …
may be important. By convention, τxy 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


y
Shear stresses arise from two sources: viscous forces and, in turbulent flow, u(y)
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.

For Newtonian fluids, viscous stress is proportional to velocity gradient. If


velocity u is a function of y only then, in laminar flow:
du
τμ (19)
dy

8
Very important in environmental flows: winds and ocean currents; also arise in rotating machinery, e.g. pumps.

Hydraulics 2 T1-12 David Apsley


This defines the dynamic viscosity, μ. (In complex flow a more general stress-strain
relationship is required, but this is beyond the scope of this course.)

In turbulent flow one is usually interested in the time-averaged mean velocity u . Since
momentum transfer between fast- and slow-moving fluid is dominated by turbulent mixing
rather than viscous stresses the mean shear stress is substantially greater than μ 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 of the latter is boundary-layer separation. Free stream

In real fluids velocity vanishes at solid boundaries (the no-slip


condition). There is a boundary layer where velocity changes rapidly
Boundary layer
from its value in the free stream to zero at the boundary. At high
Reynolds numbers boundary layers are usually very 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 may cause the more-slowly-moving fluid near the boundary to reverse
direction. This backflow leads to flow separation.

... ... slows adv


speeds up down erse
grad pressu
ient re

bac
kflo
w

flow
separation

For bodies with sharp corners flow separation occurs at all but the smallest Reynolds
numbers and causes a large increase in pressure (or form) drag
because the pressures upstream and downstream are very
different. (Upstream pressure is high because flow is brought H L
to rest; downstream pressure is low because velocities in the High Low
recirculating flow are small, so that the pressure is almost pressure pressure
constant and equal to that at the separation point.)

For more gently curved bodies separation may or may not occur.

Turbulence prevents or delays flow separation because it facilitates the transport of fast-
moving fluid from the free stream into the near-wall region, maintaining forward motion.
Thus, counter-intuitively, provoking a boundary layer on a curved surface into turbulence
(e.g. by roughening the surface) may actually reduce drag because it delays or prevents
separation. This is why golf balls have dimples.

Hydraulics 2 T1-13 David Apsley


2.4 Drag and Lift Coefficients

The force on a body in a flow can be resolved into components.

lift F
U0 drag

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 ρU 02 )  area:

Drag and Lift Coefficients

drag lift
cD  , cL  (20)
1
2 ρU 02 A 1
2 ρU 02 A

U0 is a representative velocity scale and is usually taken as the approach-flow velocity. A is a


representative area which depends on the body geometry and the nature of the flow (see
below). Just like the Reynolds number, both should be specified when defining cD or cL.

Drag on Bluff or Streamlined Bodies

Bluff bodies (i.e. flow separation)


 force is predominantly pressure drag U0
 A is the projected area (normal to the flow)
A
 cD is of order 1

Streamlined bodies (i.e. no flow separation)


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

Hydraulics 2 T1-14 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

For uniform velocity:


momentum flux  mass flux  velocity
 (ρQ)u (21)
 (ρuA)u

For 1-d flow in the x direction,


momentum flux  ρu 2 A (22)

If the velocity is not uniform then subdivide into small areas over which the velocity is
uniform and then sum or integrate:

momentum flux   ρu 2 dA (23)

Special Cases

Velocity profile Momentum flux

(i) Uniform
ρU 2 A
Area A

w
(ii) 2-dimensional u(y)
dy 
w ρu 2 dy
dA = w dy 

(iii) Axisymmetric u(r)


r dr  2
 ρu 2πr dr
dA = 2πr dr 

Hydraulics 2 T1-15 David Apsley


Example.
A two-dimensional beam of height b = 100 mm spans a square air-conditioning duct of height
h = 400 mm (see the figure below). The approach flow is uniform (uA = 0.6 m s–1), whilst the
downstream velocity profile is 2-dimensional and given by:
 3 1 2πy
 u B (  cos ), if y  h / 2
u 4 4 h

 uB , if y  h / 2
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 uB;
(b) the difference between pressures at upstream 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) Let w be the width of the tunnel. (In this particular case, w = h because the duct is square.)

By continuity (i.e., the same volumetric flow rate at both sections):



u A hw   u w dy

wake

 h / 2 3 1 2πy 
h

 w  u B (  cos ) dy   u B dy 
0 4 4 h h / 2 
Hence, dividing by width w (equivalent to working per unit width):
 3 1 h 2πy 
h/2
1 

u A h  u B  y   sin  h 
 4 4 2π h  0 2

 u B ( 83 h  12 h)
 78 u B h
Thus,
8 8
uB  uA   0.6  0.6857 m s 1
7 7

Answer: u2 = 0.69 m s–1

Hydraulics 2 T1-16 David Apsley


(b) By Bernoulli’s equation:
p A  12 ρu A2  p B  12 ρu B2
 p A  p B  12 ρ(u B2  u A2 )  12  1.2  (0.6857 2  0.6 2 )  0.06611 Pa

Answer: Pressure difference = 0.066 Pa

(c) Steady-flow momentum principle:


force = rate of change of momentum = (momentum flux)out – (momentum flux)in

Letting F be the magnitude of the force on the beam toward the right, and hence that on the
fluid toward the left,

( p A  p B )hw  F  w  ρu 2 dy  ρu A2 hw

wake


 F  ( p A  p B )hw  ρu A2 hw  ρw  u 2 dy

wake
Now,
h
 2  2
 u dy   u dy
 0
wake
h
h/2
2 πy 2 
 2 3 1
 u B (  cos ) dy   u B2 dy
0 4 4 h h / 2
 h / 2 1 2 πy 2 πy h
 u 
2
(9  6 cos  cos 2 ) dy  
0 16
B
h h 2
 1 h/ 2 2 πy 1 1 4 πy h
 u   (9  6 cos
2
  cos ) dy  
16 0
B
h 2 2 h 2
 1  h 2 πy 1 1 h 4πy 
h/2
h 
 u  9 y  6 sin
2
 y   sin  
h  0
B
16  2π h 2 2 4π 2 
 u B2  6451
h
where cos 2 θ  12 (1  cos 2θ) is used to integrate cos2. Returning to the expression for F:
F  hw( p A  p B  ρu A2  ρu B2  64
51
)
 0.4  0.4  (0.06611  1.2  0.6 2  1.2  0.6857 2  64
51
)  0.007759 N

Answer: Force on the beam = 0.0078 N

(d) Drag coefficient:


F 0.007759
cD  1 2   0.90 (beam face area = b  w)
2 ρu A (bw)
1
2  1.2  0.6 2  0.1  0.4

Answer: Drag coefficient = 0.90

Hydraulics 2 T1-17 David Apsley


Example. (Examination, January 2008)
A cylinder spans a wind tunnel of rectangular cross section and height h = 0.3 m, as shown in
the figure. The spanwise width w = 0.6 m and the cylinder diameter is 90 mm. The upstream
velocity is uA and is uniform. The velocity profile measured a short distance downstream of
the cylinder is symmetric about the centreline and is given by
  2
 y  
3

  y 
 
 10  1  6    4 , if y  0.1
u   0.1   0.1  
  
 u
 B , if y  0.1
–1
where u is the velocity in m s and y is the distance from the centreline in m.

cylinder y
uA
h
uB

(a) Assuming that the downstream velocity profile has no discontinuities, what is the
value of uB?
(b) Calculate the upstream velocity uA.
(c) Assuming that Bernoulli’s theorem is applicable outside the wake of the cylinder
calculate the pressure difference between upstream and downstream sections.
(d) Neglecting drag on the walls of the tunnel, calculate the total drag force on the
cylinder.
(e) Define a suitable drag coefficient and calculate its value.

Answer: (a) 30 m s–1; (b) 23.3 m s–1; (c) 213 Pa; (d) 26.6 N; (e) 1.51

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 ρ = 1000 kg m–3 and
ν = 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 T1-18 David Apsley


2.6 The Wake-Traverse Method for Measurement of Drag

Objects in a flow experience a force. If


Force on BODY Force on FLUID
the fluid exerts a force F on the body then
the body exerts a reaction force of the F
same magnitude but opposite direction on
F
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. In both cases upper and lower boundaries are streamlines, across which there is no
flow. Fluid passing close to the body forms a wake of low velocity downstream.

If the flow is constrained by boundaries (e.g.


in a wind tunnel) then the velocity outside the body
wake must increase slightly (in order to pass inflow wake
the same volume flow rate through the same
area), with a compensating fall in pressure.
This is called a blockage effect.

In the unconstrained case, upper and lower


boundaries should be sufficiently far away
that pressure is equal to that in the free body
stream. (Otherwise the pressure force on the inflow wake
curved boundaries will have a net component
in the x direction.)
streamline
The steady-state momentum principle gives:
force on fluid = (momentum flux)out – (momentum flux)in
   
 F   p dA   p dA   ρu 2 dA   ρu 2 dA (24)
   
in out out in

The two integrals on the RHS tend to be individually much larger than F, but almost cancel
each other out. For experimentally-determined velocity profiles this cancellation means that a
small error in either will lead to a huge error in F. To avoid this, it is common to rewrite the
RHS as a single integral over the wake. Provided that the inflow velocity is uniform (uin):
 2   
 ρu dA  u in  ρu in dA  u in  ρu dA   ρuu in dA
   
in
in
 out out
mass flux

Substituting in (24) and rearranging gives



F   [ρu (u in  u )  ( pin  p)] dA (25)

out
(Sometimes this is used as the starting point for the momentum principle, since ρu dA(uin – u)
is the rate of change of momentum in an individual stream tube carrying mass flux ρu dA.
However, this is only useful if the inflow velocity is uniform.)

Hydraulics 2 T1-19 David Apsley


For unconstrained flows, pressures upstream and downstream are equal (pin – p = 0), as are
the free-stream velocities (uin = u∞,out). In the constrained case, it can be shown that, provided
the wake is narrow compared with the duct height (so that the difference in free-stream
velocities is small), any pressure difference approximately compensates for the change in the
free-stream velocity9. Thus, in the case of either unconstrained flow or low blockage ratio,
(25) can be satisfactorily approximated by

F   ρu (u   u ) dA (26)

out
where u∞ denotes the free-stream velocity at outflow. Thus, hydrodynamic forces may be
deduced indirectly by measuring the velocity profile 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  12 ρ(u2 ,out  uin2 ) ; the error in (26) can then, after a lot of algebra, be shown to
1
 2 ρ(u in  u ,out ) dA , which is second order in the (small) free-stream velocity difference.
2
be

Hydraulics 2 T1-20 David Apsley


3. ENERGY

3.1 Bernoulli’s Equation10

Mechanical-energy principle:
change in kinetic energy = work done

In rate form, with work done by conservative forces (here, gravity) rewritten in terms of
potential energy:
rate of change of (kinetic  potential ) energy
 rate of working of non  conservative forces

Consider steady flow through a thin stream tube with varying 1


2 u2
cross-sectional area A. The mechanical energy of fluid u1
passing through it is changed by the work done by pressure
and viscous forces from the adjacent fluid and by energy
supplied or removed by external agents (pumps or turbines).

Since the sides are locally parallel to the flow, energy only flows into or out of this control
volume through ends 1 and 2. Hence, in steady flow:
rate at which energy passes 2  rate at which energy passes 1
(27)
 rate at which work is done on the fluid

The rate at which forces do work (i.e. power) is force  velocity (in the direction of the force).
The rate of working by pressure forces 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 of the stream tube because the
pressure force is perpendicular to the velocity there.

Mechanical energy consists of potential energy (mgz) and kinetic energy (½mU2). Since the
rate at which energy passes a point is mass flux  energy per unit mass, (27) becomes
ρQ( gz  12 U 2 ) 2  ρQ( gz  12 U 2 )1  ( pAU )1  ( pAU ) 2  W
where W is the net rate of working of friction and other external forces (e.g. due to pumps or
turbines). ρQ is the mass flow rate, which must be constant along the stream tube. Dividing
by ρQ = (ρUA)1 = (ρUA)2 gives
p p W
( gz  12 U 2 ) 2  ( gz  12 U 2 )1  ( )1  ( ) 2 
ρ ρ ρQ
or, rearranging,
Bernoulli’s Equation With Losses

p
Δ(  gz  12 U 2 )  W (28)
ρ
 means “change in”. W is the work done by non-conservative forces. It represents all non-
pressure work done on the fluid and is composed of frictional work (always negative) and
any work done on the flow by pumps, turbines etc. Each term in the equation represents
energy or work done per unit mass.
10
Daniel Bernoulli (1700-1782) Swiss-Dutch mathematician, a member of an illustrious family of well-known
mathematicians.

Hydraulics 2 T1-21 David Apsley


Thermal or Compressible Flow

(28) can be extended to thermal flows (boilers, condensers, refrigerators, ...) or compressible
flow by using the total energy equation:
change in (internal + kinetic) energy = work done + heat input

(28) then becomes


p
Δ(e   gz  12 U 2 )  rate of doing work  rate of supplying heat (29)
ρ
p
The RHS is energy input per unit mass. The quantity e  is called the specific enthalpy.
ρ

Incompressible Flow

If there are no losses and no work done by pumps or turbines then (28) reduces to
p
 gz  12 U 2  constant (along a streamline) (30)
ρ
For incompressible flow, density ρ is also constant along a streamline. Then we have:
Bernoulli’s equation (without losses):

p  ρgz  12 ρU 2  constant (along a streamline) (31)


where
p = static pressure
p + ρgz = piezometric pressure
2 ρU
1 2
= dynamic pressure

Note the assumptions:


 along a streamline (different streamlines may have a different “constant”)
 steady
 incompressible
 inviscid (no losses)

Hydraulics 2 T1-22 David Apsley


3.2 Fluid Head

In hydraulics both energy and pressure are often expressed in length units; e.g. “metres of
water” or “millimetres of mercury”. In equation (31) each term has dimensions of pressure, or
of energy per unit volume. Dividing by weight per unit volume (i.e. specific weight) ρg, these
become energies per unit weight. This has dimensions of length and is called fluid head.

p  ρgz  12 ρU 2  energy per unit volume, or total pressure


p U2 (32)
z  energy per unit weight , or total head
ρg 2g

p
= pressure head
ρg
p
z = piezometric head
ρg
U2
= dynamic head
2g

Energy losses due to friction and the change in pressure imparted by pumps are often
specified in terms of head. For pumps the rate of working (i.e. power) is given by
power  ρgQH (33)

where Q is the quantity of flow and H is the change in head. This will be revisited in Topic 4.

3.3 Static and Stagnation Pressure

A stagnation point is a point on a streamline


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

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 ρU 2 . Define:
stagnation pressure p  12 ρU 2
static pressure p
dynamic pressure 1
2
ρU 2

The dynamic pressure (and hence the flow velocity) is found by measuring the difference
between stagnation and static pressures.

Hydraulics 2 T1-23 David Apsley


3.4 Flow Measurement

3.4.1 Measurement of Pressure – Manometry Principles

Basic rules of hydrostatics for a connected body of fluid:


(1) Same fluid, same height  same pressure
(2) Same fluid, different height  Δp  ρgΔz
(3) Different fluids: pressure is continuous at an interface

U-Tube Manometer A B
By (1) the pressure at level C is the same in both arms of the y
manometer. By (2) and (3) it can be found from pA and pB
respectively by summing the changes in pressure over the heights

of columns of fluid: h
left arm right arm
   
p A  ρg (h  y)  pC  p B  ρgy  ρ m gh C
where ρ and ρm are the densities of the working fluid and the
manometer fluid respectively. Cancelling ρgy and subtracting ρgh m
gives the pressure difference Δp  p A  p B :
Manometer Equation

Δp  (ρ m  ρ) gh (34)

If the working fluid is a gas then ρ « ρm and (34) can be approximated by


Δp  ρ m gh (35)

Inclined Manometer

Differences in head may be small and difficult to


measure accurately. The movement of the (large)
L
manometer fluid may be amplified by inclining h (small)
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 θ (36)

Hydraulics 2 T1-24 David Apsley


3.4.2 Measurement of Velocity

Method: bring the fluid to rest at one point and measure the difference between static and
stagnation (Pitot) pressures:
p0  p  1
2
ρU 2
Pitot static dynamic
pressure pressure pressure

Examples.

(1) Open-channel flow.


2
U /2g
The free-surface streamline and the streamline free surface
approaching the Pitot tube both have the same
total head, but where the flow is brought to rest at
the front of the Pitot tube the dynamic head U2/2g
is converted into pressure head and subsequently
elevation.
stagnation point

piezometer Pitot tube

U2
2g
(2) Pipe flow – Pitot tube and piezometer

Same principle as above, but the pressure in the


pipe is usually higher than atmospheric, so that
water rises in the piezometer also. U

total pressure tube


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

Hydraulics 2 T1-25 David Apsley


3.4.3 Measurement of Quantity of Flow

General Method

Provide a locally reduced area to change speed, and measure the resulting change in pressure.
The combination of continuity and Bernouilli’s equation yields bulk velocity and flow rate.

1
Venturi Flowmeter (Smooth Constriction) 2
A venturi is a localised smooth constriction in a
duct.

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

Bernoulli: p1  12 ρU12  p2  12 ρU 22  p1  p2  12 ρ(U 22  U12 )

A1
Continuity: U1 A1  U 2 A2  U 2  U1
A2
Eliminating the velocity U2 in the constricted section gives
 A 
Δp  12 ρU 12 ( 1 ) 2  1
 A2 
Rearranging for U1 and using Q = U1A1,
1/ 2
 2 A12 Δp 
Q  (37)
 ( A1 /A2 )  1 ρ 
2

Q  constant  (Δp)1 / 2 (38)

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

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


non-ideal flow. cd is the ratio of the actual flow rate to the flow rate computed from ideal-
flow theory (Bernoulli’s equation and continuity):
Definition of Discharge Coefficient

Q  cd 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 T1-26 David Apsley


Orifice Flowmeter (Abrupt Constriction)
vena contracta
An orifice is an aperture of small streamwise thickness through which
fluid passes.

The streamwise thickness determines frictional losses.

The fluid cannot turn immediately, so the emerging stream tube


continues to narrow as far as the vena contracta – the section of
minimum cross-sectional area.

An orifice meter is a means of measuring the


flow rate in a duct by measuring 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.

Advantages: cheap and small.


Disadvantage: considerable loss of energy.

By the same process as that for the venturi meter one obtains:
1/ 2
 2 A12 Δp 
Q 
 ( A1 /Av )  1 ρ 
2

where Av is the area of the vena contracta (not the aperture). 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 the
relationship between the flow rate Q and the pressure drop Δp:
Q  constant  (Δp)1 / 2 (40)
with the constant determined by calibration.

Hydraulics 2 T1-27 David Apsley


3.5 Tank Filling and Emptying

3.5.1 Free Discharge Under Gravity


1
For a tank, discharging as a free jet, with free surface at a
distance h above the discharging fluid, apply Bernoulli’s
equation between the free surface and the jet: h
p1  12 ρU12  ρgz1  p2  12 ρU 22  ρgz 2
Here, p1  p2  patm , U1  0, z1  z 2  h , so that:
2
1
2 ρU 22  ρg ( z1  z 2 )  ρgh

This amounts to an exchange of gravitational potential energy


for kinetic energy and gives
Torricelli’s Formula11

U exit  2 gh (41)

Ideally the discharge would be


Qideal  2 gh  (area of orifice ) (42)
This is not true in practice because of
 frictional effects (small for a sharp-edged orifice)
 contraction (area of vena contracta < area of orifice).
If these are significant then a coefficient of discharge cd may be introduced to compensate for
both of these effects:
Q  cd 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 (i.e. curved) exit to minimise rapid changes in
direction. However, frictional losses are then greater.

Strictly, 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. Also, if the
aperture is large (compared with the height to the free surface), then the value of Uexit 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.

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

Hydraulics 2 T1-28 David Apsley


3.5.2 Continuity Equation

Consider the filling or emptying of a tank (where “tank” is generic; in civil engineering it
could just as well mean “reservoir”).

Volume h(t)
flow in V flow out

By continuity (here, conservation of volume):


rate of change of volume  (volume flow rate) in  (volume flow rate) out
dV
 Qin  Qout (43)
dt

By writing both volume and volume flow rates in terms of the water depth h one can generate
a differential equation for h and solve for the time to empty or fill a tank.

dh
As

The change in volume dV corresponding to a change in water level dh is given, to first order
in small quantities, by
dV  As dh
where subscript s denotes “surface”. Equation (43) can then be written as a first-order
differential equation in the water level h:
dh
As  Qin  Qout (44)
dt

Hydraulics 2 T1-29 David Apsley


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 water-surface area here is constant and is that of a circle with diameter D = 0.5 m:
πD 2
As   0.1963 m 2
4
The inflow rate is zero, Qin  0 . The outflow is gravity-driven flow through the aperture, i.e.
Qout  cd Qideal  cd U ideal Aexit
where cd = 0.6, Aexit = 810-4 m2 and U ideal  2 gh by Torricelli’s formula.

Hence,
dh
As  Qin  Qout
dt
dh
 As  0  cd Aexit 2 gh
dt
dh
 0.1963  2.126  10 3 h (in metre-second units)
dt
Separating variables:
 92.33h 1 / 2 dh  dt

Integrating between t = 0 (where h = 0.8 m) and the emptying time T (where h = 0):
0 T
 
 92.33 h 1 / 2 dh   dt
0.8 0
0
 
  92.332 h1 / 2   T
  0.8
Hence,
T  92.33  2  0.8  165.2 s

Answer: Emptying time = 165 s.

Hydraulics 2 T1-30 David Apsley


In general the water-surface area As is also a function of water depth h.

Example.
A conical hopper of semi-vertex angle 30º contains water to a depth of 0.8 m. If a small hole
of diameter 20 mm is suddenly opened at its point, estimate (assuming a discharge coefficient
cd = 0.8):
(a) the initial discharge (quantity of flow);
(b) the time taken to reduce the depth of water to 0.4 m.

o 0.8 m
30

Answer: (a) 0.996 L s–1; 177 s.

Hydraulics 2 T1-31 David Apsley


3.6 Summary of Methods For Incorporating Non-Ideal Flow

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

In practice, it is necessary to compensate for non-ideal flow. The methods employed include
the following.

Discharge Coefficients

Correct the quantity of flow deduced for ideal flow (no losses; simplified geometry):
Q  cd Qideal

Loss Coefficients

Quantify head or pressure losses in pipes or other conduits:


V2
ΔH   K ( )
2g
or, equivalently (and with the same K):
Δp   K ( 12 ρV 2 )
L
(e.g. pipe friction: K  λ , where λ is the friction factor).
D

Momentum and Energy Correction Coefficients

Correction factors for non-uniform velocity profiles12:


momentum correction coefficien t , β :  ρu 2 dA  β(ρu 2 A)

 av

kinetic energy correction coefficien t , α :  ρu 3 dA  α(ρu 3 A)



 av

12
In some textbooks the kinetic-energy correction coefficient α is called the “Coriolis coefficient”. Since this is
completely unrelated to the much-more-important Coriolis forces (involved in rotational motion) that name will
not be used here.

Hydraulics 2 T1-32 David Apsley

You might also like