You are on page 1of 31

Hydraulics 2 T2-1 David Apsley

TOPIC T2: FLOW IN PIPES AND CHANNELS AUTUMN 2004



1. Pipe flow
1.1 Introduction
1.2 Governing equations for circular pipes
1.3 Laminar pipe flow
1.4 Turbulent pipe flow
1.5 Expressions for the friction factor f
1.6 Other losses
1.7 Pipeline calculations
1.8 Energy and hydraulic grade lines
1.9 Pipe networks

2. Open-channel flow
2.1 Normal flow
2.2 Hydraulic radius and the drag law
2.3 Friction laws Chzy and Mannings formulae
2.4 Open-channel flow calculations
2.5 Optimal shape of cross-section

Appendix


References

Massey (1998) Chapters 6, 7
Chadwick and Morfett (2004) Chapters 4, 5
White (2002) Chapters 6, 10
Hamill (2001) Chapters 6, 8


Hydraulics 2 T2-2 David Apsley
1. PIPE FLOW

1.1 Introduction

The flow of water, oil and gas in pipelines is of great importance to engineers. An
understanding of the relationship between quantity of flow (or discharge) and head loss is
vital in designing reservoir and distribution systems.

Laminar and Turbulent Flow

In 1883, Osborne Reynolds demonstrated the occurrence of
two regimes of flow laminar or turbulent according to the
size of a dimensionless parameter later named the Reynolds
number. The conventional definition for round pipes is
Re
VD
(1)
where
V = average velocity (flow rate / cross-sectional area)
D = diameter
= kinematic viscosity (= / )

For smooth-walled pipes the critical Reynolds number at which transition occurs is typically
Re
crit
2300 (2)
but in commercial pipes transition from intermittent to fully-turbulent flow may occur over a
range 2000 < Re < 4000.


Development Length

At inflow, the velocity profile is often nearly
uniform. Boundary layers develop on the sides
because of wall friction. These grow with distance to merge at the centre of the pipe. Beyond
this distance the velocity profile becomes fully-developed (i.e., doesnt change any further
with distance). Typical correlations for this development length are:

=
) (turbulent Re 4.4
) laminar ( Re 06 . 0
1/6
D
L
devel
(3)

The kinematic viscosity of air and water is such that most pipe and duct flows in civil
engineering have high Reynolds numbers and are fully turbulent. The development length
can usually be ignored as a function of total length. However, there are some duct flows
notably laboratory wind tunnels where one would prefer the development length to be much
longer than the working section.

Example.
water
= 1.010
6
m
2
s
1
. Calculate the Reynolds numbers for average velocity
0.5 m s
1
in pipes of inside diameter 12 mm and 0.3 m. Estimate the development length in
each case.

Answer. Re = 6000 and 1.510
5
. L
devel
= 0.23 m and 9.6 m.
laminar
turbulent



Hydraulics 2 T2-3 David Apsley
1.2 Governing Equations For Circular Pipes

Fully-developed pipe flow is determined by a balance between three forces
pressure
weight
friction

It will be shown that the combined effect of pressure and weight can be expressed as a
piezometric pressure:
p
*
= p + gz (4)

Changes in pressure can also be written in terms of changes in head:
z
g
p
g
p
H H g p or
*
*
+ = = = (5)


For a circular pipe of radius R, consider the forces with components along the pipe axis for an
internal cylindrical fluid element of radius r < R and length l.

mg

p+ p
p


Note that:
(1) p is the average pressure over a cross-section; for circular pipes this is equal to the
centreline pressure, with equal and opposite hydrostatic variation above and below;
(2) The arrow drawn for the shear stress indicates the conventional positive direction for ,
corresponding to the force exerted by the outer fluid on the inner. In the present example,
is clearly negative.

Balancing forces in the direction of the pipe axis:

0 ) 2 ( sin ) )( ( ) (
2 2
= + + +

friction
weight
force pressure
l r mg r p p r p
From the geometry,
l
z
l r m
sin
2
=
=

Hence:
0 ) 2 ( ) (
2 2
= + l r z g r r p


Hydraulics 2 T2-4 David Apsley
Dividing by the volume, r
2
l,
0 2
) (
= +
+

r l
gz p

Hence, letting l 0, the pressure gradient is related to the shear stress by

l
p
r
d
*
d
2
1
= where p
*
= p + gz (6)

Thus, pressure and weight only appear in combination as the piezometric pressure. To
maintain flow against frictional forces requires either a pressure drop, a component of weight
along the pipe axis (arising from a slope) or both.

Since p* is independent of radius (since there is no acceleration across the pipe) and dp*/dl is
independent of distance l (since is independent of l in fully-developed flow) we may write,
using G for (downstream) pressure gradient,
constant
l
p
G = =
d
*
d
(7)
Hence, the shear stress varies linearly across the pipe:
Gr
2
1
= (8)


The maximum shear stress occurs at the walls:

4
) (
GD
R
w
= = (9)
where D = 2R is the pipe diameter.


Equations (8) and (9) apply to all fully-developed pipe flow regardless of whether it is
laminar or turbulent. For laminar flow we can use it to establish an actual velocity profile,
because we can write in terms of u/r (Section 1.3). For turbulent flow an analytical
velocity profile is not available, but gross parameters (quantity of flow, head loss) may be
obtained if the wall shear stress
w
can be related (empirically) to the dynamic pressure
( V
2
) by means of a friction factor (Section 1.4).



Hydraulics 2 T2-5 David Apsley

1.3 Laminar Pipe Flow

Laminar, pressure-driven flow through a circular pipe is often called Poiseuille flow or
Hagen
1
-Poiseuille
2
flow. It is one of a very small class of simple flows for which exact
solutions of the Navier-Stokes equations can be written down analytically. Others include:
plane Poiseuille flow pressure-driven flow between parallel plane walls;
plane Couette
3
flow flow between parallel plane walls moving at different speeds;
flow between concentric rotating cylinders.
In each case a critical Reynolds number exists, beyond which the laminar flow solution
becomes unstable to small disturbances and the flow becomes turbulent.

In fully-developed pipe flow, shear stresses balance pressure and weight:
Gr
2
1
= (8)
where

L
gh
l
p
G
f
d
d
*
= = (10)
is the piezometric pressure gradient and h
f
= H is the head lost by friction over length L.


Viscous forces occur in flows with velocity gradients. Across any interface the drag of upper
fluid on lower fluid is equal and opposite to the drag of lower fluid on upper. The direction of
the drag force is such as to reduce the velocity difference. With the sign convention adopted
for (stress exerted by outer fluid on inner fluid) we have:
r
u

=
stress = viscosity velocity gradient
Hence,
r
G
r
u
2
1
=


Integrating:
constant r
G
u + =
2
4

Applying the no-slip condition at the wall (u = 0 on r = R) the complete solution is

) (
4
2 2
r R
G
u = (11)


The shape of the velocity profile is parabolic (quadratic in r).


1
G.L.H. Hagen, German engineer who, in 1839, measured water flow in long brass pipes and reported that there
appeared to be two regimes of flow.
2
J.L.M Poiseuille (1799-1869), French physician who was interested in flow in blood vessels.
3
M.F.A. Couette (1858-1943), flow between concentric rotating cylinders - method used to measure viscosity.

Hydraulics 2 T2-6 David Apsley

Example. Find, from the velocity distribution given above,

(a) the centreline velocity;

(b) the average velocity V;

(c) the flow rate in terms of head loss and pipe diameter;

(d) the momentum factor
(
]
(
= A u
A V
d
1
2
2
(= factor by which the actual momentum
flux varies from that assuming uniform velocity with the same mass flow rate).

(e) the wall shear stress;

(f) the friction factor, defined as
2
2
1
V
f
w
= as a function of Re.



















Answers. (a)
4
) 0 (
2
0
GR
u u = = ; (b)
0
2
1
2
8
u
GR
V = = ; (c)
L
gh
D
Q
f
128
4
=

(d) = 4/3; (e) GR
r
u
R r
w
2
1
=

=
=
; (f)
Re
16
= f

The exercise demonstrates some of the key qualitative aspects of laminar pipe flow:
the average velocity is half the maximum velocity (in turbulent flow they would be nearly
equal);
the quantity of flow is proportional to (diameter)
4
(head loss);
the friction factor varies as 1/Re.

Hydraulics 2 T2-7 David Apsley
1.4 Turbulent Pipe Flow

In turbulent flow one is usually interested in time-averaged or mean quantities. The shear
stress is the mean rate of transport of momentum per unit area across any surface drawn in
the flow and is dominated by turbulent fluctuations rather than molecular viscosity.

In turbulent flow there is no longer an explicit relationship between mean stress and mean
velocity gradient u/r (because momentum is transferred more by the net effect of random
fluctuations than by viscous forces). Hence, to relate quantity of flow to head loss we require
an empirical relation connecting the wall shear stress and the average velocity in the pipe.
This is quantified by a friction factor defined as

2
2
1
pressure dynamic
stress shear wall
V
f
w
= = (12)

The problem is to relate the main hydraulic parameters:
head loss h
f

quantity of flow Q
diameter D
Other factors involved are: length of pipe (L); surface roughness (k
s
); viscosity ( ).

L
p
e
r
i
m
e
t
e
r
D
mg

p+p
p

w
area


The following forces are in balance in fully-developed pipe flow:
pressure
weight
friction

For the length of pipe shown, the balance of forces along the axis in fully-developed flow is:

0 sin ) ( = + +

drag wall
w
weight
force pressure net
L perimeter mg area p p area p
Since
L
z
L area m
sin =
=

this gives:
L perimeter area gz p
w
= + ) (


Hydraulics 2 T2-8 David Apsley
Dividing by the volume (area L),
area
perimeter
L
p
w
=
*

Write:

f
gh p
*
= (pressure difference in terms of frictional head loss)
) (
2
2
1
V f
w
= (definition of friction factor)

D area
perimeter 4
= (for circular pipes)

Then, rearranging gives the frictional head loss for circular pipes:

Darcy
4
-Weisbach
5
Equation
)
2
( 4
2
g
V
D
L
f h
f
= (13)


Note the form of this equation. It relates frictional head loss h
f
to the dynamic head V
2
/2g.
The total head loss is proportional to pipe length (intuitively reasonable!) and inversely
proportional to diameter.


It remains to specify f for a given pipe flow. Methods for doing so are discussed in section 1.5
and lead to the Colebrook-White equation. Since f depends on both the relative roughness of
the pipe (k
s
/D) and the flow velocity itself (through the Reynolds number Re VD/ ) either
an iterative solution or a chart-based solution is usually required.

The quantity of flow Q can be related to the average velocity and pipe diameter via
4
2
D
V Q =
At high Reynolds numbers f tends to a constant (determined by surface roughness) for any
particular pipe. Compare:
5
2 2
D
Q
D
V
h
f
(turbulent)
4 2
D
Q
D
V
h
f
(laminar)


Very important note.
Lecturers in this department favour calling f the friction factor. One reason is historical; the
other is that it coincides with the definition of skin-friction coefficient in aerodynamics.
However, many engineers refer to 4f as the friction factor and denote it by (or,
unfortunately, f). Be very wary of the definition. You should be able to distinguish it by the
expression for f in laminar flow: 16/Re in our notation; 64/Re with the alternative definition.

4
Henri Darcy (1803-1858), French engineer; conducted experiments on pipe flow.
5
Julius Weisbach, German professor who, in 1850, published the first modern textbook on hydrodynamics.

Hydraulics 2 T2-9 David Apsley
1.5 Expressions for the Friction Factor f

Smooth Pipes
Laminar flow (theory)
Re
16
= f
Turbulent flow (Blasius
6
)
25 . 0
Re
079 . 0
= f


Rough Pipes

Nikuradse (1933) used sand grains to roughen pipe surfaces. He defined a relative roughness
k
s
/D, where k
s
is the sand grain size and D the diameter of the pipe. His experimental curves
for friction factor showed 5 regions:
1. laminar flow (Re < Re
crit
2000; roughness irrelevant)
2. transition (2000 < Re < 4000)
3. smooth-wall turbulence (k
s
/D small; f is a function of Reynolds number only)
4. rough-wall turbulence (k
s
/D large; f is a function of relative roughness only)
5. transitional turbulence (f a function of both Re and k
s
/D)

In the limiting cases of totally smooth or roughness-dominated, Prandtl
7
and Von Krmn
8

gave, respectively:
Smooth-wall turbulence:
26 . 1
Re
log 0 . 4
1
10
f
f
=
Rough-wall turbulence:
s
k
D
f
71 . 3
log 0 . 4
1
10
=

Unfortunately, in practice, many pipes of interest lie in the 5
th
regime both roughness and
Reynolds number are important so that the friction factor is not constant for any particular
pipe. Colebrook and White (1937) combined smooth- and rough-wall turbulence laws into a
single formula, the Colebrook-White transition formula.


Colebrook-White Transition Formula


|
|
.
|

\
|
+ =
f
D
k
f
s
Re
26 . 1
71 . 3
log 0 . 4
1
10
(14)


This remains the principal formula for the friction factor. The main difficulty is that it is
implicit (f appears on both sides of the equation) and an iterative solution is usually required.

6
P.R.H. Blasius (1883-1970); first postgraduate student of Prandtl; also derived a famous equation for the
velocity profile in a laminar, flat-plate boundary layer.
7
Ludwig Prandtl (1875-1953); German engineer; introduced boundary-layer theory.
8
Theodore von Krmn (1881-1963); Hungarian mathematician and aeronautical engineer; gave his name to the
double row of vortices shed from a 2-d bluff body and now known as a Krmn vortex street

Hydraulics 2 T2-10 David Apsley
Various workers have produced explicit approximations to (14), accurate to within a few
percent for realistic ranges of Reynolds number see the references in Massey and Whites
textbooks.

Graphical solutions of (14) exist. The most well known are:
the Moody chart ( f versus Re for various values of k
s
/D)
Hydraulic Research Station Charts for the Hydraulic Design of Channels and Pipes
(a set of charts one for each value of k
s
giving hydraulic quantities D, V, Q, S
f
on
the same diagram).

Equivalent Sand Roughness

For commercial pipes the pattern of surface roughness may be very different to that in the
artificially-roughened surfaces of Nikuradse. Colebrook (1939) and Moody (1944) gathered
data to establish effective roughness for typical pipe materials. Typical values of k
s
are given
in the Appendix.


1.6 Other Losses

Pipeline systems are subject to two sorts of losses:
wall-friction losses, contributing a slow change of head over a large distance;
minor losses due to abrupt changes in geometry; e.g. pipe junctions, valves, etc.

Each type of loss can be represented by a change in head and are quantified in terms of a loss
coefficient K, the ratio of head change ( H) to dynamic head (V
2
/2g):

g
V
K H
2
2
= (15)


Thus, for pipe friction:

D
L
f K 4 = (16)

Typical values of K are given in the tables below (from Massey, 1998). For pictures of the
various valve types see Whites textbook.

Commercial pipe fittings (approximate)

Fitting K
Globe valve 10
Gate valve wide open 0.2
Gate valve open 5.6
Pump foot valve 1.5
90 elbow 0.9
45 elbow 0.4
side outlet of T-junction 1.8



Hydraulics 2 T2-11 David Apsley
Entry/exit losses

Configuration K
Bell-mouthed entry 0
Abrupt entry 0.5
Protruding entry 1.0
Bell-mouthed exit 0.2
Abrupt enlargement 0.5
Exit to atmosphere
*
1.0

* Exit to Atmosphere

Applying Bernoullis equation with losses:
H
g
V
z
g
p
g
V
z
g
p
)
2
( )
2
(
2
1
1
1
2
2
2
2
= + + + +

For a reservoir, V
1
is 0. Strictly the exit velocity V
2
is part of the
total head at section 2. In practice, it is convenient to transfer it to
the RHS as a minor loss in piezometric head:

loss exit
g
V
H z
g
p
2
) (
2
= +
The extra term on the RHS is equivalent to a minor loss with loss coefficient 1.0.


Minor losses are a one-off loss, occurring at a single point. Frictional losses are
proportional to the length of pipe L and, in the grand scheme of things, usually dominate. For
long pipelines minor losses are often ignored.


1.7 Pipeline Calculations

The main objective of pipeline calculations is to establish the relationship between the
available head and the quantity of flow (head-discharge relation) for the system.

The basic method is an energy one: to balance the difference in head at the two ends (usually
known from their heights) against the sum of all the head changes along the pipeline:

H(end) H(start) = H(pumps) + H(turbines) + H(losses)


H for pumps is positive. H for turbines and losses is negative.

Since each type of loss is proportional to the dynamic head ( H V
2
/2g) this leads to a
relationship between head changes and average velocity V (and, thence, the discharge Q).

The examples in this section focus on cases where the main loss is frictional and minor losses
can be neglected. This is the case for many large-scale reservoir and distribution systems. If
significant, minor losses can be readily incorporated into the head-loss equation via loss


Hydraulics 2 T2-12 David Apsley
coefficients K.

Basic pipe parameters are illustrated below. Note that, although we draw a reservoir at each
end of the pipe, this is simply a diagrammatic way of saying a point at which the total head
is known.


h
L
D
Q


Typical pipeline problems are: given two parameters from:
head difference h
quantity of flow Q
diameter D
find the third.

Other parameters are the pipe length L, wall roughness k
s
and the kinematic viscosity of the
fluid .

Such problems are solved by the simultaneous solution of:
(1) a head-loss equation
e.g. for pipe friction:
)
2
( 4
2
g
V
D
L
f h = (13)
(2) an expression for the friction factor
e.g. the Colebrook-White equation

|
|
.
|

\
|
+ =
f
D
k
f
s
Re
26 . 1
71 . 3
log 0 . 4
1
10
(14)


In most problems it is necessary to solve (14) iteratively. The exception is the calculation of
quantity of flow Q when head loss h and diameter D are known, because in this special case
(Type 1 below) the Reynolds-number-dependent part can be expanded to give:

|
|
.
|

\
|
+ =
2
10
26 . 1
71 . 3
log 0 . 4
1
fV D
D
k
f
s
(17)
Since the combination fV
2
can be found from (13), f can be found directly. Knowledge of
both fV
2
and f allows one to find V and hence Q.



Hydraulics 2 T2-13 David Apsley
Type 1: Diameter D and head difference h known; find the quantity of flow Q.

Example. A pipeline 10 km long, 300 mm diameter and with roughness 0.03 mm, conveys
water from a reservoir (top water level 850 m above datum) to a water treatment plant (inlet
water level 700 m above datum). Assuming that the reservoir remains full, estimate the
quantity of flow. Take = 1.110
6
m
2
s
1
.


Solution.
List known parameters:
L = 10000 m
D = 0.3 m
h = 150 m
k
s
= 310
5
m
= 1.110
6
m
2
s
1


Since D and h are known, the head-loss equation enables us to find fV
2
:
2 2 2
2
s m 02207 . 0
10000 2
150 3 . 0 81 . 9
2
)
2
( 4

=


= = =
L
gDh
fV
g
V
D
L
f h

Hence, rewriting the Colebrook-White equation,
94 . 16
02207 . 0 3 . 0
10 1 . 1 26 . 1
3 . 0 71 . 3
10 3
log 0 . 4
26 . 1
71 . 3
log 0 . 4
Re
26 . 1
71 . 3
log 0 . 4
1
6 5
10
2
10
10
=
|
|
.
|

\
|

+

=
|
|
.
|

\
|
+ =
|
|
.
|

\
|
+ =

fV D
D
k
f
D
k
f
s
s

Hence,
003485 . 0
94 . 16
1
2
= = f
Knowledge of both fV
2
and f gives
1
2
s m 517 . 2
003485 . 0
02207 . 0

= = =
f
fV
V

Finally, the quantity of flow may be computed as velocity area:
1 3
2 2
s m 0.1779
4
3 . 0
517 . 2 )
4
(

=

= =
D
V Q

Answer: Quantity of flow = 0.18 m
3
s
-1
.

Hydraulics 2 T2-14 David Apsley
Type 2: Diameter D and quantity of flow Q known; find the head difference h.

Example. The outflow from a pipeline is 30 L s
1
. The pipe diameter is 150 mm, length
500 m and roughness estimated at 0.06 mm. Find the head loss along the pipe.

Solution.
List known parameters:
Q = 0.03 m
3
s
1

L = 500 m
D = 0.15 m
k
s
= 610
5
m
= 1.110
6
m
2
s
1


Inspect the head-loss equation:
)
2
( 4
2
g
V
D
L
f h =
We can get V from Q and D, but to find h we will require the friction factor.

First V:
1
2 2
s m 698 . 1
4 / 0.15
0.03
4 /

= =
D
Q
V

Inspect the Colebrook-White equation:
|
|
.
|

\
|
+ =
f
D
k
f
s
Re
26 . 1
71 . 3
log 0 . 4
1
10

To use this we require the Reynolds number:
231500
10 1 . 1
15 . 0 698 . 1
Re
6
=

= =

VD


Substituting values for k
s
, D and Re in the Colebrook-White equation and rearranging for f:
2
6
4
10
)
10 443 . 5
10 078 . 1 ( log 0 . 16
1
(
(


+
=

f
f

Iterating from an initial guess, with successive values substituted into the RHS:
Initial guess: f = 0.01
First iteration f = 0.004351
Second iteration f = 0.004515
Third iteration f = 0.004507
Fourth iteration f = 0.004507

f can then be substituted in the head-loss equation to derive h:
m 831 . 8
81 . 9 2
698 . 1
15 . 0
500
004507 . 0 4 )
2
( 4
2 2
=

= =
g
V
D
L
f h

Answer: Head loss = 8.8 m.

Hydraulics 2 T2-15 David Apsley
Type 3 (Sizing problem): Quantity of flow Q and available head h known; find the required
diameter D.

Example. A flow of 0.4 m
3
s
1
is to be conveyed from a headworks at 1050 m above datum to
a treatment plant at 1000 m above datum. The length of the pipeline is 5 km. Estimate the
required diameter, assuming that k
s
= 0.03 mm.

Solution.
List known parameters:
Q = 0.4 m
3
s
1

h = 50 m
L = 5000 m
k
s
= 310
5
m
= 1.110
6
m
2
s
1


Before iterating, try to write D in terms of f. From the head-loss equation:

5 2
2
2
2
2
32
4 /
2
)
2
( 4
gD
fL Q
D
Q
gD
fL
g
V
D
L
f h = |
.
|

\
|
= =

5 / 1
5 / 1
2
2
32
f
gh
L Q
D
|
|
.
|

\
|
=
Substituting values of Q, L and h gives a working expression (with D in metres):

5 / 1
395 . 1 f D = (*)

The Colebrook-White equation for f is:

|
|
.
|

\
|
+ =
f
D
k
f
s
Re
26 . 1
71 . 3
log 0 . 4
1
10

The Reynolds number can be written in terms of the diameter D:

D D
Q D
D
Q VD
5
2
10 630 . 4 1 4
4 /
Re

=
|
.
|

\
|
=
|
.
|

\
|
= =
Substituting this expression for Re we obtain an iterative formula for f:

2
6 6
10
)
10 721 . 2 10 086 . 8
( log 0 . 16
1
(
(

=

f
D
D
f (**)

Iterate (*) and (**) in turn, until convergence.
Guess: f = 0.01 D = 0.5554 m

Iteration 1: f = 0.003049 D = 0.4379 m

Iteration 2: f = 0.003232 D = 0.4431 m

Iteration 3: f = 0.003223 D = 0.4428 m
Iteration 4: f = 0.003223 D = 0.4428 m

Answer: Required diameter = 0.44 m.
In practice (as anyone who does DIY knows) commercial pipes are only made with certain
standard diameters and the next available larger diameter should be chosen.

Hydraulics 2 T2-16 David Apsley
1.8 Energy and Hydraulic Grade Lines

Energy grade lines and hydraulic grade lines are graphical means of portraying the energy
changes in, for example, reservoir/pipeline systems.

Three elevations may be drawn:
pipe centreline: z geometric height

hydraulic grade line (HGL): z
g
p
+ piezometric head
energy grade line (EGL):
g
V
z
g
p
2
2
+ + total head

p is the gauge pressure (i.e. pressure relative to atmospheric).


Illustrations




Pipe friction only







Pipe friction with minor
losses (exaggerated),
including change in pipe
diameter.






Pumped system




energy grade line
hydraulic grade line
pipeline
p/ g
V /2g
2 reservoir
reservoir

E
G
L
H
G
L
pipeline
entry loss
exit loss

E
G
L
H
G
L
pipeline
pump


Hydraulics 2 T2-17 David Apsley
Energy Grade Line

Shows the change in total head along the pipeline.
starts at level of water in supply reservoir;
small discontinuities correspond to entry loss, exit loss or another minor loss;
steady downward slope reflects pipe friction (slope change if pipe radius changes);
large discontinuities correspond to turbines (loss of head) or pumps (gain of head).
Non-frictional losses can often be ignored.

The EGL always lies a distance V
2
/2g above the HGL. For uniform pipes, the two are
parallel.

The EGL represents, at any station, the maximum height to which water may be
delivered.


Hydraulic Grade Line

Shows the change in piezometric head along the pipeline.

For pipe flow the HGL lies a distance p/ g above the pipe centreline. Thus, the
difference between pipe elevation and hydraulic grade line gives the static pressure. If
the HGL drops below pipe elevation this means negative gauge pressures (i.e. less
than atmospheric). This is generally undesirable since:
extraneous matter may be sucked into the pipe through any leaks;
for large negative gauge pressures, dissolved gases may come out of solution and
cause water hammer.
A hydraulic grade line more than p
atm
/ g (about 10 m of water) below the pipeline is
impossible.

The HGL is the height to which the liquid would rise in a piezometer tube.

For open-channel flows, pressure is atmospheric (i.e. p = 0) at the surface; the HGL is
then the height of the free surface.





Hydraulics 2 T2-18 David Apsley
Example. Two reservoirs, the water levels in which are at elevations 180 m and 150 m
respectively, are connected by a pipe 3000 m long, 600 mm diameter and friction factor
0.00625. The elevation of the ground along the line of the pipeline is given in the table below.




Assuming a rounded inlet and an abrupt outlet calculate the quantity of flow. Find the
maximum depth of the pipeline below ground if the absolute pressure therein is not to fall
below 3 m of water.

Draw to scale on graph paper, the ground level, pipeline level and energy and hydraulic grade
lines for a suitable pipeline. Take atmospheric pressure to be 10 m of water.


Solution.
The losses are as follows (average velocity V in m s
1
):
Friction:
2
2 2
371 . 6
81 . 9 2 6 . 0
3000
00625 . 0 4
2
) ( 4 V
V
g
V
D
L
f h
f
=

= =
Entry loss: 0 ) entry ( =
L
h
Exit loss:
2
2
025 . 0 )
2
( 5 . 0 ) exit ( V
g
V
h
L
= =

2 2
396 . 6 ) 025 . 0 371 . 6 ( loss head Total V V = + =

Total head loss = difference in water levels of reservoirs. Hence
30 396 . 6
2
= V

1
s m 166 . 2
396 . 6
30

= = V
The quantity of flow is therefore
1 3
2 2
s m 6124 . 0
4
6 . 0
166 . 2
4

= = =
D
V Q

It is useful at this stage to calculate also the dynamic head (V
2
/2g = 0.2391 m) and the exit
loss (V
2
/2g = 0.1196 m). Note that the dynamic head is very small compared to the total
head change.

The energy grade line can now be drawn (see below) it starts at the water level in the first
reservoir and descends with uniform slope to a height 150.1196 m (i.e. allowing for the exit
loss) at the second reservoir.

The hydraulic grade line can then be drawn a distance V
2
/2g = 0.2391 m below the energy
grade line. The start and end coordinates give it an equation (with lengths in metres):
x z
HGL
009960 . 0 7609 . 179 =

The ground elevation is now marked on the graph from the data given in the equation.

We are now asked to ensure that the absolute pressure does not fall below 3 m of water: in
Distance (m) 0 150 300 1800 3000
Elevation (m) 175 165 190 140 147


Hydraulics 2 T2-19 David Apsley
other words (since atmospheric pressure is equivalent to 10 m of water) that the pipeline is
not more than 7 m above the hydraulic grade line. The most significant problem occurs at
x = 300 m, where the maximum pipeline height can be
m 8 . 183 7 300 009960 . 0 7609 . 179 7 = + = + =
HGL pipeline
z z
Since the ground level here is 190 m, the pipeline depth below ground must be
m 2 . 6 8 . 183 190 =
Since unnecessary excavation is undesirable, the pipeline is laid at ground level, except
where it must be lowered to satisfy pressure constraints. A suitable pipeline is marked on the
diagram.


175
180
165
190
140
147
150
exit loss = 0.12 m
velocity head = 0.239 m
E
G
L
H
G
L
0 150 300 1800 3000
p
i
p
e
l
i
n
e
7 m
excavation necessary
m





Hydraulics 2 T2-20 David Apsley
1.9 Pipe Networks

For all pipe combinations the following basic principles apply:
(1) mass conservation at junctions (total flow in = total flow out)
(2) there can be only one value of head at any one point
(3) each pipe must satisfy its individual head- discharge relation, H = kQ
2


The last of these comes from the proportionality between head loss and dynamic head, i.e.
g
V
H
2
2

The total head loss is the sum of individual components minor losses and friction. For
frictional head losses only:
2
2
2
4
2
4
)
2
( 4
|
.
|

\
|
= =
D
Q
gD
fL
g
V
D
L
f H
Hence, for friction,
2
kQ H = where
5 2
32
gD
fL
k =
In pipe-network calculations, the discharge coefficient k is often taken as a constant for each
pipe (although, in reality, it will vary slightly with Reynolds number and hence with Q).

There is a close analogy with electrical networks:
head H potential V
discharge Q current I
However, the hydraulic equivalent of Ohms law is usually non-linear:
head loss H Q
2
potential difference V I


Pipes in Series

Q
1
= Q
2
same flow in each pipe
H = H
1
+ H
2
total head loss is sum of individual losses


Pipes in Parallel

H
1
= H
2
same head loss across each pipe
Q = Q
1
+ Q
2
total flow is sum of individual flows


Branched Pipes Single Junction

This is a classic problem also known as the
three-reservoir problem.

Heads H
A
, H
B
and H
C
are known (from the
top-water levels in the reservoirs). The head
at J must be found (iteratively) to satisfy:

A
C
J
B
?

1
2

1
2


Hydraulics 2 T2-21 David Apsley
(a) the loss equation (H = kQ
2
) for each pipe; in this example:

2
AJ AJ J A
Q k H H =

2
BJ BJ J B
Q k H H =

2
JC JC C J
Q k H H =

(b) continuity at junction J:

JC BJ AJ
Q Q Q = +

Note that the direction of flow in pipe BJ cannot be established a priori but must emerge as
part of the solution, depending on whether the head at B is higher or lower than that at J.


Solution Procedure

(1) Guess H
J

(2) Calculate flow rates in all pipes (from head differences)
(3) Calculate flow into and flow out of J
(4) If necessary, adjust H
J
to reduce any flow imbalance and repeat from (2)


If the direction of flow in pipe BJ is not obvious then a good initial guess is to set H
J
= H
B
so
that there is initially no flow in this pipe. The first flow-rate calculation will then establish
whether H
J
should be lowered or raised.


Example. Reservoirs A, B and C have constant water levels of 150, 120 and 90 m
respectively. A 300 mm pipe, 1600 m long, leaves A and runs to J at elevation 137 m. Here it
divides into a 200 mm pipe, 1600 m long, leading to B and a 150 mm pipe, 2400 m long,
leading to C.

(a) Assuming f = 0.005 in all pipes, calculate the flow in each pipe.

(b) Calculate the reading of a Bourdon pressure gauge attached to the junction J.


A
B
C
J
L
=
1
6
0
0
m
D
=
0
.3
m
L
=
1
6
0
0

m
D
=
0
.
2

m
L
=
2
4
0
0

m
D
=
0
.
1
5

m
150 m
120 m
90 m



Hydraulics 2 T2-22 David Apsley
Solution.
The first step is to prepare head-discharge relations for each pipe:
)
2
( 4
2
g
V
D
L
f h
f
= where
4 /
2
D
Q
V =

2
5 2
32
Q
gD
fL
h
f
=

Substituting f, L and D for each pipe we obtain the head-discharge relationships:
Pipe AJ:
2
1088
AJ J A
Q H H = or
1088
150
J
AJ
H
Q

=
Pipe JB:
2
8265
JB B J
Q H H = or
8265
120
=
J
JB
H
Q
Pipe JC:
2
52243
JC C J
Q H H = or
52243
90
=
J
JC
H
Q

The flow in each pipe is calculated for a sequence of values of H
J
until the flow into junction
J equals the flow out of junction J.
If there is more flow into the junction than out of it, H
J
needs to be increased.
If there is more flow out of the junction than into it, H
J
needs to be decreased.
After the first two guesses at H
J
, subsequent iterations are guided by interpolation.

The sequence is conveniently set out in a table.
H
J
(m)


Q
AJ
(m
3
s
1
)

1088
150
J
H
= )
Q
JB
(m
3
s
1
)

8265
120
=
J
H

Q
JC
(m
3
s
1
)

(
52243
90
=
J
H
)
flow out flow in
(m
3
s
1
)
(= Q
JB
+ Q
JC
Q
AJ
)
140 0.0959 0.0492 0.0309 0.0158
145 0.0678 0.0550 0.0324 0.0196
142.2 0.0847 0.0518 0.0316 0.0013
142.4 0.0836 0.0521 0.0317 0.0002

This is sufficient accuracy (0.0002/0.0836 or about 0.24%). The quantity of flow in each pipe
is given in the bottom row of the table.

(b) A Bourdon gauge measures absolute pressure. From the piezometric head at the junction:
z
g
p
H
J
+ =
where p is the gauge pressure. Hence,
Pa 53000
) 137 4 . 142 ( 81 . 9 1000 ) (
=
= = z H g p
J


Taking atmospheric pressure as 101000 Pa, the absolute pressure is then
101000 + 53000 = 154000 Pa

Answer: 1.54 bar.

Hydraulics 2 T2-23 David Apsley
Multiple Junctions

This is considerably more complex as more than one unknown head must be adjusted. These
sort of problems will not be covered in this course (although there is one such problem on the
example sheet). General algorithms the loop method and the nodal method are discussed
in Chadwick and Morfett (2004). They are based on making estimates for the small
increments Q
i
or H
i
respectively in pipes or junctions and coupling these increments by a
requirement to satisfy head-loss equations in each pipeline and continuity at each junction.
The result is a set of simultaneous equations for the Q
i
or H
i
.


Hydraulics 2 T2-24 David Apsley
2. OPEN-CHANNEL FLOW

Flow in open channels (e.g. rivers, canals, guttering, ...) and partially-full conduits (e.g.
sewers) is characterised by the presence of a free surface, at which pressure is atmospheric.

Since the (gauge) pressure is zero at the free surface, the hydraulic grade line (p/ g + z)
coincides with the free surface. The energy grade line lies a distance g V /2
2
above it.

Open-channel flows are driven by gravitational forces.


2.1 Normal Flow

The flow is uniform if the velocity profile does not change along the channel. The flow is
steady if it does not change with time.

Steady uniform flow is called normal flow and the depth of water is called the normal depth.

h
h
f
L
atm
p
=
p




In normal flow:
the component of weight down the slope exactly balances bed friction;
the loss of fluid head is exactly equal to the loss on elevation;
the bed slope, hydraulic grade line and energy grade line are all parallel; i.e.
geometric slope S
0
= friction slope S
f

L
h
x
z
f
b
=
d
d




Hydraulics 2 T2-25 David Apsley
2.2 Hydraulic Radius and the Drag Law

In both open channels and partially-full pipes, wall friction
only occurs along the wetted perimeter.

Let A be the cross-sectional area occupied by fluid and P the
wetted perimeter.


For steady, uniform flow, the component of weight
down the slope balances wall friction:
PL g AL
w
sin ) ( =
where
w
is the average wall friction. Hence,
sin ) (
P
A
g
w
=

The quantity

perimeter wetted
area
=
P
A
R
h
(18)

is called the hydraulic radius.

Hence, for normal flow,

S gR
h w
= (19)

where S (= drop length) is the slope; note that tan sin for small angles.


Examples.
(1) For a circular pipe running full,
D R
R
R
P
A
R
h
4
1
2
1
2
2
= = = = (20)
i.e. for a full circular pipe, the hydraulic radius is half the geometric radius. (Sorry
folks, this is just one of those things!). As a result, it is common to define a hydraulic
diameter D
h
by
h h
R D 4 =

(2) For a very wide channel of uniform depth h, side walls make negligible contribution
to the wetted perimeter and hence
h R
h
=
i.e. R
h
is equal to the depth of flow.

To progress we need an expression for the average wall stress
w
.


A
P

mg
L


Hydraulics 2 T2-26 David Apsley
2.3 Friction Laws Chzy and Mannings Formulae

A friction factor could be used to relate the (average) wall shear stress to the dynamic
pressure:
) (
2
2
1
V f
w
=
Hence, substituting the normal-flow relationship (19) for
w
:
S gR V f
h
) (
2
2
1
=
or

) 4 ( 2
4
2
) ( slope
2 2
h h
f
R g
V
f
gR
fV
L
h
S = = (21)
This is just the Darcy-Weisbach equation with D replaced by D
h
= 4R
h
. The friction factor
may be obtained from the Colebrook-White equation, but with Reynolds number
4
Re
V R V D
h h
= =
The procedure for solving problems is similar to that for full pipes, but with the added
problem of having to determine the level of fluid in the conduit as part of the solution.


Friction factors based on the Colebrook-White equation are unsatisfactory for many open
conduits (in particular, natural water courses such as rivers) because the shear stress is not
constant around the wetted perimeter. For practical problems, engineers tend to use simpler
empirical formulae due to Chzy
9
and Manning
10
.

The formula for the slope (21) can be rearranged as one for average velocity:

Chzys Formula
S R C V
h
= (22)

C ( f g/ 2 = ) is Chzys coefficient. It varies with channel roughness and hydraulic radius.

The most popular correlation for C is that of Manning who proposed n R C
h
/
6 / 1
= , in terms of
a roughness-dependent coefficient n. Combined with Chzys formula (22), this yields:

Mannings Formula


2 / 1 3 / 2
1
S R
n
V
h
= (23)


Very important: both Chzys C and Mannings n are dimensional and depend on the units
used. Typical values of n in metre-second units are given in the Appendix. Typical figures for
artificially-lined channels and natural water courses are about 0.015 and 0.03 respectively.

9
Antoine Chzy (1718-1798); French engineer who carried out experiments on the Seine and on the Courpalet
Canal in 1769.
10
Robert Manning (1816-1897); Irish engineer.

Hydraulics 2 T2-27 David Apsley
2.4 Uniform-Flow Calculations

Assuming that the channel slope, shape and lining material are known, there are two main
classes of problem:

(Type A) Given the depth (h) determine the quantity of flow (Q)

(1) Calculate the area A and wetted perimeter P from geometry.
(2) Calculate the hydraulic radius
P
A
R
h
= .
(3) Calculate the average velocity from Mannings formula:
2 / 1 3 / 2
1
S R
n
V
h
= .
(4) Calculate the quantity of flow as velocity area: Q = VA.


(Type B) Given the quantity of flow (Q) determine the depth (h)

(1) Follow the steps for Type A above to write algebraic expressions for,
successively, A, P, R
h
, V and Q in terms of depth h.

(2) Invert the Q vs h relationship graphically or numerically.


Example. A smooth concrete-lined channel has trapezoidal cross-section with base width 6m
and sides of slope 1V:2H. If the bed slope is 1 in 500 and the normal depth is 2 m calculate
the quantity of flow.


Solution.
We are given slope S = 0.002. From the Appendix, Mannings n is 0.012.

6 m
2 m
4 m


Break the trapezoidal section into rectangular and triangular elements to obtain, successively:
Area:
2
2
1
m 20 2 4 2 2 6 = + = A
Wetted perimeter: m 94 . 14 2 4 2 6
2 2
= + + = P
Hydraulic radius: m 339 . 1
94 . 14
20
= = =
P
A
R
h

Average velocity:
1 3 / 2 2 / 1 3 / 2
s m 528 . 4 727 . 3
1

= = =
h h
R S R
n
V
Quantity of flow:
1 3
s m 56 . 90 20 528 . 4

= = = VA Q

Answer: 91 m
3
s
1
.


Hydraulics 2 T2-28 David Apsley

Example. For the channel above, if the quantity of flow is 40 m
3
s
1
, what is the normal
depth?


Solution. This time we need to leave all quantities as functions of height h.
Area:
2
2 6 h h A + =
Wetted perimeter: h P 5 2 6 + =
Hydraulic radius:
P
A
R
h
=
Average velocity:
3 / 2
727 . 3
h
R V =
Quantity of flow: VA Q =

We can now try a few values of h.

h (m) A (m s
1
) P (m) R
h
(m) V (m s
1
) Q (m
3
s
1
)
2 20.00 14.94 1.339 4.528 90.56
1 8.00 10.47 0.7641 3.115 24.92
1.23 10.41 11.50 0.9052 3.488 36.31
1.28 10.96 11.72 0.9352 3.564 39.06
1.30 11.18 11.81 0.9467 3.593 40.17

After the first two guesses, subsequent choices of h home in on the solution by
interpolating/extrapolating from previous results.

Answer: h = 1.30 m

Note. Inverting Q(h) is an ideal application for Microsoft Excel and particularly for the
Solver Tool in that program.

Hydraulics 2 T2-29 David Apsley
2.5 Optimal Shape of Cross-Section

The most hydraulically-efficient shape of channel is the one which can pass the greatest
quantity of flow for any given area or, equivalently, the smallest area for a given quantity of
flow. From Mannings formula and the corresponding expression for quantity of flow we see
that this occurs for the minimum hydraulic radius or, equivalently, for the minimum wetted
perimeter.

A semi-circle is the most hydraulically-efficient of all channel cross-sections. However,
hydraulic efficiency is not the only consideration and one must also consider, for example,
fabrication costs, excavation and, for loose granular linings, the maximum slope of the sides.
Many applications favour trapezoidal channels.

Expressions for A, P and R
h
for important channel shapes are given in the table below.

rectangle


h
b


trapezoid


b
h


circle

R
h

area A bh
tan
2
h
bh +
) 2 sin (
2
1
2
R
wetted perimeter P h b 2 +
sin
2h
b +
2R
hydraulic radius R
h

b h
h
/ 2 1+

sin / 2
tan /
h b
h b
h
+
+

|
.
|

\
|

2
2 sin
1
2
R



Trapezoidal Channels

For a trapezoidal channel we have, from the table in the
previous section:
Area:
tan
2
h
bh A + =
Wetted perimeter:
sin
2h
b P + =

What depth of flow and what angle of side should we choose for maximum hydraulic
efficiency?

To minimise the wetted perimeter for maximum hydraulic efficiency, we substitute for b in
terms of the fixed area A:
)
tan
1
sin
2
(
sin
2
)
tan
(
sin
2
+ = + = + = h
h
A h h
h
A h
b P (24)
b
h



Hydraulics 2 T2-30 David Apsley

To minimise P with respect to water depth we set
0 )
tan
1
sin
2
(
2
= +

h
A
h
P

and, on substituting the bracketed term into the expression (24) for P, we obtain
h
A
P
2
=
The hydraulic radius is then
2
h
P
A
R
h
=
In other words, for maximum hydraulic efficiency, a trapezoidal channel should be so
proportioned that its hydraulic radius is half the depth of flow.


Similarly, to minimise P with respect to the angle of slope of the sides, , we set
0 ) cos 2 1 (
sin
) sec
tan
1
cos
sin
2
(
2
2
2 2
= = +

h
h
P

This occurs when cos = . i.e. the most efficient side angle for a trapezoidal channel is 60.

Substituting these results for h and into the general expression for R
h
one obtains
2 / 3 / = b h ; i.e. the most hydraulically-efficient trapezoidal channel shape is half a regular
hexagon.


Circular Ducts

In similar fashion it can be shown that the maximum quantity of flow for a circular duct
actually occurs when the duct is not full in fact for a depth about 94% of the diameter
(Exercise. Prove it; then try to explain in words why you might expect this).



Hydraulics 2 T2-31 David Apsley
Appendix

Material k
s
(mm)
Riveted steel 0.9 9.0
Concrete 0.3 3.0
Wood stave 0.18 0.9
Cast iron 0.26
Galvanised iron 0.15
Asphalted cast iron 0.12
Commercial steel or wrought iron 0.046
Drawn tubing 0.0015
Glass 0 (smooth)

Table 1. Typical roughness for commercial pipes (from White, 2002).


n (metre-second units) k
s
(mm)
Artificial lined channels:
Glass 0.01 0.3
Brass 0.011 0.6
Steel, smooth 0.012 1.0
painted 0.014 2.4
riveted 0.015 3.7
Cast iron 0.013 1.6
Concrete, finished 0.012 1.0
unfinished 0.014 2.4
Planed wood 0.012 1.0
Clay tile 0.014 2.4
Brickwork 0.015 3.7
Asphalt 0.016 5.4
Corrugated metal 0.022 37
Rubble masonry 0.025 80
Excavated earth channels:
Clean 0.022 37
Gravelly 0.025 80
Weedy 0.03 240
Stony, cobbles 0.035 500
Natural channels:
Clean and straight 0.03 240
Sluggish, deep pools 0.04 900
Major rivers 0.035 500
Floodplains:
Pasture, farmland 0.035 500
Light brush 0.05 2000
Heavy brush 0.075 5000
Trees 0.15 ?

Table 2. Typical values of Mannings n and equivalent sand roughness (from White, 2002).

You might also like