Vous êtes sur la page 1sur 31

4

Flow in pipes and closed


conduits

4 .1

Introduction

The flow of water, oil and gas in pipes is of immense practica l significance
in civil engineering. Water is conveyed from its source, normally in pressure
pipelines (Fig. 4 .1), to water treatment plants where it enters the distribution
system and finally arrives at the consumer. Surface water drainage and
sewerage is conveyed by closed condu its, which do not usually operate under
pressure, to sewage treatment plants, from where it is usually discharged
to a river or the sea. Oil and gas are often transferred fro m theif source
by pressure pipeli nes to refi neries (oi l) or into 3 distribution ncrwork fo r
supply (gas) .
Surprising as it may seem, a comprehensive theory of rhe flow of fluids in
pipes was not developed until the late 1930s, and pracrica l design methods
for the eva luation of discharges, pressures and head losses d id not appear
until 1958. Until these design tools were ava ilable, the efficient design of
pipeline systems was not possi ble.
This chapter describes the theories of pipe fl ow, beginning with a review
of the historica l context and ending with the practical applications.

4.2

The historical context

Table 4.1 lists the names of the main contributors, and their contributions,
to pipe flow theories in chronologica l order.
The Colebrook-Wh ite transition fo rmu la represents the culmination of
all the previous work, and ~an be .applied to any flu id in any pipe operating under turbu lent fl ow conditions. The later contributions of M oody,
Ackers and Barr are mainl y concerned with the practica l application o f the
Colebrook- White equation.

92

FLOW IN PIPES AND CLOSED CONDUITS

Figure 4. t

The synthetic hydrological cycle.

There a re three major concepts described in the table. These are:


1. the distinction between laminar and turbulent flow;
2. the distinction between rough and smooth pipes;
3. the distinction between aftificially roughened pipes and commercial
pipes.
To undersrand these concepts, the best starting point is the co ntribution of
Reynolds, followed by [he laminar flow equations, before proceedi ng to the
morc complex turbulent flow equations.

TI-lE HISTORICAL cmnurr


Tabl~

4.1

Th~

chronological

d~vclopm~nt

of pipe flow

theori~s.

Contribution

Nam~

Date

93

1839-41
1850
1884

Hag~n and Poiseuille


Darcy and W~isbach
Reynolds

1913

Blasius

1914

Stanton and Pannell

1930

Nikuradse

19305

Prandtl and

laminar flow equation


turbulent flow equation
distinction betwn laminar and
turbulent flow - Reynolds' Number
friction factor equation for smooth
pi~

, 1937-39

\' 011

Karm:s'n

Colebrook and White

1944

Moody

1958

Ackers

1975

Sarr

experimental values of the friction


factor for smooth pi~
experimental values of the friction
factor (or artificially rough pipes
equations for rough and smooth
friction factors
experimental values of the friction
factor for commercial pi~ and the
transition formula
the Moody diagram for commercial
pipes
the Hrdraulics Research Station
Charts and Tables for the design of
pipes and channels
dire<:t solution of the ColehrookWhite equation

Laminar and tllrbillent flow

Reynolds' experiments demonstrated that there were twO kinds of flow laminar and turbulent - as described in Chapter 3. He found that transition
from laminar to turbulent flow occurred at a critical velocity for a given pipe
and flujd. Expressing his results in terms of the dimensionless parameter
Re = pDV/f.L, he found that for Re less than about 2000 the flow was always
laminar, and that for Re greater than about 4000 the flow was always
turbulent. For Re between 2000 and 4000, he found that the flow could be
either laminar or turbulent, and termed this the transition region.
In a further set of experiments, he found thaI for laminar flow the frictional head Joss in a pipe was proportional to the velocity, and that for
turbulent flow the head loss was proportional to the square of the velocity.
These two results had been previously determined by Hagen and Poiseuille
(hl oo,", and Darcy and Weisbach (lJlooV1l, but it was Reynolds who put
these equations in [he context of laminar and turbulent flow.

94

4.3

FLOW IN PIPES AND CLOSED CONDUITS

Fundamental concepts of pipe flow

The momentum equation


Before proceeding to derive the laminar and rurbulent flow equations, it is
instructive to consider the momentum (or dynamic) equation of flow and
the influence of the bou ndary layer.
Referring to Figure 4.2, showing an elementa l annulus of fluid, thickness
8r,Iength 8/, in a pipe of radius R, the forces acting are the pressure forces,
the shear forces and the weight of the fluid. The sum of the forces acting is
equal to the change of momentum. In this case momentum change is zero,
since the flow is steady and uniform. I-(ence
p2-rrr8r -

(p+~~ 8/) 2ttr8r+T2ttr81 - (T + :; 8r) 2tt{r +81') 81

+ pg21l'r8/8rsin a = 0
Sening sine = - dz/d/a nd divid ing by 2-rrr8r81 gives
dp dT T
d:z
- - - - - - - pg - = O
dl dr r
dl

('t +

1ir)21l(r + &-)&'

(p + .2&)2'111Sr

d/

p2v&-

Figure 4.2

Derivalion of Ihe momentum equation.

FUNDAMENTAL CONCEPTS OF PIPE FLOW

9S

(ignoring second-order terms), or


dp' '1 d'1
------= 0
dl
r dr

where V(= p +pgz) is the piezometric pressure measured from the datum
4=0. As
1 d
I (dT
dT T
--('11')=r-+'1) =-+r dr
r dr
dr r

then
dp 1 d
-----(n)=O
di
r dr

Rearranging,

Integrating both sides with respect to r,


'1r =

dp "
-&"2
+constant

At the centreline r = 0, and therdore constant = O. Hence


dp' r
'1 = - - dl 2

(4.1)

Equation (4.1) is the momentum equation for steady unifo rm flow in


a pipe. It is equally applicable to laminar or turbulent flow, and relates
the shear stress '1 at radius r to the rate of head loss with distance along
the pipe. If an expression for the shear force can be found in terms of the
velocity at radius r, then the momentum equation may be used to relate the
velocity (and hence discharge) to head loss.
In the case of laminar flow, this is a simple matter. However, for the case
of turbulent flow it is more complicated, as will be seen in the following
sections.
The dcvelopmcm of boulldary layers

Figure 4.3(a) shows the development of laminar flow in a pipe. At entry to


the pipe, a laminar boundary layer begi ns to grow. However, the growth

96

FLOW IN PIPES AND CLOSED CONDUITS


boundary layer

/
"

la' Laminar now


laminar

.~t, ,.y"

lurbulenl bounda ry layer

laminar bounda ry layer


(bj Turbuknl flow

Figure 4.3

Boundary layers and velocity distributions.

of the bounda ry layer is halted when it reaches the pipe centreline. and
thereafter the fl ow consists entirely of a boundary layer of thickness r. The
resulting velocity distribution is as shown in Figure 4.3(a).
For Ihe case of turbulent flow shown Ul Figure 4.3(b}, the growth of the
boundary layer is not suppressed umi l it becomes a turbulent bou nda ry
layer with the accompa nying laminar sublayer. The resulting velocity
profile therefore differs considerably (rom the laminar case. The existence
of the laminar sub-layer is of prime importance in explaini ng the difference
between smooth and rough pipes.
Expressions relating shear stress to velocity have been developed in
Chapter J, and these will be used in explaining the pipe flow equations in
the following sections.

LAMINAR FLOW

4.4

97

Laminar flow

For the case of laminar flow , Newton's law of viscosity may be used to
evaluate the shear stress (T) in terms of velocity (u):

du

dll,

T =~ d)' =-~dr

Substituting into the momentum equation (4.1),

dlf,

dV r
dl 2

T= - ~ -=---

d,

0'

dtl,
1 dp'
-=--,
dr
2~ dl
Integrating,
II,

At the pipe boundary,

II,

1 dp'
= 4~ dT,-l + conStant

= 0 and r = R, hence
constant =

1 dp'

---R'
4" dl

and

(4.2)
Equation (4.2) represents a parabolic velocity distribution, as shown in
Figure 4.3(a). The discharge (Q) may be determined from (4.2). Returning
to Figure 4.1 and considering the elemental discharge (SQ) through the
an nulus, then

5Q = 21Tr5m,
Integrating

Q=21T 10 r lI , dr
and substituting for

tI,

from (4.2) gives

Q=_ 21Tdp' ( Rr(R 2 _r 2)dr


4jJ. dl 10

98

FLOW IN PIPES AND CLOSED CONDUITS

0'

dp'
8" dt
'IT

Q~---- R'

(4.3)

Also the mcan velocity (V) may be obtained directly from Q;

1Tdp' 4 '
8"" df
'lTR2

V~ - ~----R

0'

I dp-

V=---R'
8" dt

(4.4)

In practice, it is usual to express (4.4 ) in terms of frictional head loss by


maki ng the substitution
dp'

hl~-

pg

Equation (4.4) then becomes

0'

b = 32""LV
I
pgO'

(4.5)

This is the Hagen-Poiscuillc equation, named aftcr the two peoplc who first
carried out (independen tly) the experimental work leading to it.
The wall shear stress (TO) may be related to the mean velocity (V) by
eliminating dfJ' ( dl from (4.1) and (4.4) to give
(4.6)
As T= TO when r= R, then
(4.7)

Equation (4.6) shows that (for a given V) the shear stress is proportiona l
to r, and is zero at the pipe centreline, with a maximum value (TO) at the
pipe boundary.

LAMINAR FLOW

Example 4.1

99

Lamil1ar pipe flow

Oil flows through a 25 mm diameter pipe with a mean vclocil)' of 0.3 mls. Given
that JL = 4.8 X 10- 2 kg/m sand p = 800ks/m}, calculate (a) the pressure drop in
a 45 m length and (b) the maximum vclocil)', and the velocity 5 mm from the pipe
wall.

SO/lItio"
First check thaI flow is laminar, i.e. Re < 2000.
Re = pDV/JL = 800 x 0.025 x 0.3/4.8 x 10- 2
= 125
(a) To find thc pressure drop, apply (4.5):

hf = 32JLLV/pgOZ
= (32 x 4.8 x 10-2 x 45 x 0.3)/(800 x 9.81 x 0.025 2)
= 4.228m(of oil)
or Ap = -pgh, = -33. 18kN/m1 . (Note: the negative sign indicates that pressure
reduces in Ihe direction of flow.)
(b) To find the "elocilies, apply (4.2):

Idp" 2 2
" =---(R -r)
,
4JL dl
The maximum velocity (U....~) occurs at the pipe centreline, i.e. when r = 0, hence
1
33. 18x10J
2
U..... x =- 4 x.)(
48 ) 02
x(0.025/2)
45
=0.6m/s
(Nnte: fln .. ~ _2x mean velocity (colllpare (4.2) and (4.4).))
To find the velocity 5 111m from the pipe wall (Us), usc (4.2) with r = (0.025/2)-

0.005, i.e. r = 0.0075:


=_

1/

I
)(_33.18XIO)(0.01252_0.00752)
4x4.8 )( 10-1
45

=0.384m/s

100

FLOW IN PIPES AND CLOSED CONOUITS

4.5 T urbulent fl ow
For rurbulent fl ow, Newton's viscosity law d~s not apply and, as described
in Chapter 3, sem i-empirica l relationships for TO were derived by Pra ndtl.
Also, Reynolds' experiments, and the ea rl ier ones of Darcy and Weisbach,
indicated that head loss waS proportiona l to mea n velocity squared. Using
the momentum equation (4. 1J. then
dp' R
'To = - - dl 2

,nd
dp = hfpg
dl
L
hence

Assuming h, = KV 2 , based on the experimental results cited above, then

0'

(1o, h, ~ KV').
Returning [0 the momentum equation and making the substitution
TO = K. V.t, then

hence

0'

TIJRBULENT FLOW

101

Making the substitution X= 8Kdp, then


(4.8 )
This is the Darcy-Weisbach equation, in which" is ca lled the pipe friction
factor and is sometimes referred to as f (America n practice) or 4f (ea rly
British practice). In current practice, A is the normal usage and is found, for
instance, in the Hydraulics Research Stalion charts and tables. It should be
noted that A is dimensionless, and may be used with any system of units.
The original investigators presumed that the friction factor was constant.
This was subsequently found to be incorrect (as described in secrion 3.6).
Equations relating A to both the Reynolds Number and the pipe roughness
were developed later.

Smooth pipes and the Blasius equation


Experimenta l investigations by Blasius and others early in the 20th century
led (0 [he equation
X= 0.316/Re.2J

(4.9)

The later experiments of Stanton and Pannel, using drawn brass tubes,
confirmed the validity of the Blasius equation for Reynolds' Numbers up to
105 However, at higher values of Re (he Blasius equation underestimated
X for these pipes. Before further progress could be made, the distinction
between 'smooth' and 'rough' pipes had to be established.

Artificially rough pipes mId Nikllradse's experimental results


Nik uradse made a major contribution ( 0 the theory of pipe flow by objec
tively differentiati ng between smooth a nd rough turbulence in pipes. He
carried out a painstaking series of experiments to determine both the fric
tion factor :lTld the velocity distributions at various Reynolds' Numbers
up to 3 x 106 In these experiments, pipes were artificially roughened by
sticking uniform sand grains on to smooth pipes. He defined the relative
roughness (k,/D) as the ratio of the sand grain size to the pipe diameter.
By using pipes of different diameter and sand grains of different size, he
produced a set of experimenta l resu lts of ~ and Re for a range of relative
roughness of 1130 [0 1/1014.

He plotted his results as log ~ against log Re for each value of K,/D, as
shown in Figure 4.4. This figure shows that there are five regions of flow,
as follows:
(a) Laminar flow. The region in which the relative roughness has no influ ence on lhe friction factor. This was assumed in deriving the I-IagenPoiseuillc equation (4.5). Equating this to the Darcy-Weisbach equation
(4.8) gives
32~VL

pglY

~LVl

2gD

0'

~=64~=64
pDV
Re

(4.10)

Hence, the Darcy-Weisbach equation may also be used for laminar


flow, provided that ~ is eva luated by (4.10).
(b) Transition from laminar to turbulent fl ow. An unstable region
between Re = 2000 and 4000. Fortunately, pipe flow normally lies
outside this region.
(c) Smooth turbulence. TIle limiting line of turbu lent flo w, approached
by all values of relative roughness as Re decreases.
(d) Transitional turbulence. The region in which}., varies with both Re
and k,/D. The limit of this region varies wit h k,/D. In practice, most of
pipe flow lies within th is region.

0. 116

"

\~

J1

transitional
turbulence

,,
,,
,,

0.1ll2

00t6

""

00'

"

rough

"

"'~/" 11J,6t,

"-

I.....

00'

Re (log loCale)
Figure 4.4

"
'"
'"
'"
'"
2$2

,,
00'

DIlks

IUrbule~e

Nikuradsc's experimental results.

00'

TIJRBULENT FLOW

103

(e) Rough turbulence. The region in which A. remai ns constant for a given
k,/D, and is independent of Re.
An explanation of why these five regions exist has already been given in
section J.6. It may be summarized as follows:
Laminar flow.
miSSion.

Surface roughness has no influence on shear stress trans-

Transitional turbulence. The presence of the laminar sub-layer 'smooths'


the effect of surface roughness.
Rough turbulence. The surface roughness is large enough to break up the
laminar sub-layer giving turbulence right across the pipe.
The rOllgh aud smooth laws of tlO"

Komia" and Pra"dtl

The publication of Nikuradse's experimental results (particularly his


VelOCity distribution measurements) was used by von Karman and Prandtl
to supplement their own work on rurbulcnt boundary layers. By combining
their theories of turbulent boundar}' layer flows with the experimental
results, they derived the semi-empirical rough and smooth laws. These were:
for smooth pipes
-=210g--

1
J'i:

ReJ'i:
2.51

(4. 11 )

_1_ = 2 log 3.7D


J'i:
k,

(4. 12)

for rough pipes

The smooth law is a bener fit to the experimental data than the Blasius
equation .
The Colebrook- White tra"s;t;ol1 fomlllia

The experimenta l work of Nikuradse and the theoretica l work of von


Karman and Prandtl provided the framework for a theory of pipe fric tion. However, these results were not of direct use to engineers because
they applied only to artificia ll y roughened pipes. Commercial pipes have
roughness which is uneven both in size and spacing, and do not, therefore,
necessari ly correspond to Ihe pipes used in Nikuradse's experiments.

104

FLOW IN PIPES AND CLOSED CONDUITS

Colebrook and White made nyo major contributions to the development


and application of pipe friction theory to engineering design. Initially, they
carried out experiments to determine the effect of nonuniform roughness as
found in commercial pipes. They discovered that in the turbulent transition
region the A-Re curves exhibited a gradual change from smooth to rough
turbulence in contrast to Niku radse's 'S'-shaped cu rves fo r uniform roughness, size and spacing. Colebrook then went on to determine the 'effective
roughness' size of many commercial pipes. He achieved [his by studying
published results of frictional head loss and discharge for commercia l pipes,
ranging in size from 4 inches (10 1.6mm) to 61 inches (l549.4mm), and
for materials, including drawn brass, galvanized, cast and wrought iron,
bitumen-lined pipes and concrete-lined pipes. By comparing the friction
factor of these pipes with Nikuradse's results for uniform roughness size in
the rough turbulent zone, he was able to determine an 'effective roughness'
size for the commercial pipes equivalent to Nikuradse's results. He was thus
able to publish a list of k. values applicable to commercial pipes.
A second contribution of Colebrook and White stemmed from their
experimental results on non-uniform roughness. They combined the von
Kamlan-Prandtl rough and smooth laws in the form

\
\ (k,
2.5\ )
.J). = -2 og 3.7D + Rc.J).

(4.13)

This gave predicted results very dose to the observed transitional behaviour
of commercial pipes, and is known as the Colebrook-White transition
formula. It is applicable to the whole of the turbulent region for commercial
pipes using an effective roughness val ue determined experimentally for each
type of pipe.

The practical application of the Colebrook-Wlhite tra1lsition formula


Equation (4. 13) was not at first used very widely by engineers, mainly
because it was not expressed directly in terms of the standard engineering
variables of diameter, discharge and hydraulic gradient. In addition, the
equation is implicit and requ ires a rrial-and-error solution. In the 19405,
sl ide-rules and logarithm tables were the main computational aids of t he
engineer, since pocket calculators and computers were not then available.
So these objections to the use of the Colebrook-White equation were not
unreasonable.
The first attempt to make engineering calculations easier was made by
Moody. He produced a h-Re plot based all (4.13) for commercial pipes, as
shown in Figure 4.5 which is now known as the Moody diagram. He also
presented an explicit formula for h:

h = 0.0055 [1 + (200~Ok.

+ ~:) 13]

(4.14)

Iriln"'llOnallUrbulencc

'~ MIRe
II.m":I~';c,
.01<
~~
tun 2

,_

IH)64

\ ;

O.INS

11.1

" k_1r.msitional

;,

.,
.,
"

\1

"

\1

0,(136

0,10 2

Re

). 0.02S

0.05

0.03

~ r-:.---t-

0.02
0.015

"

--

~---

" ",

~ I::::-

'.

0 .002

""

~ ---

o III 2

smOOlh ----pipes

001 0

U.fJOK

0.01
0.008

~~

\
IWI

"-

IUl2 0

t---..-

eT "

IUlZ 4

rough turbulence

0.04

lurbulence

"

,
10

2W'}

-"-

I '"

10'

I I I I" I

10

Re

Figure 4.5

The Moody diagram.

"'

0 .0004

"-

I I " ' " 10

.001
0 .0008
.0006

0 .0002

"

O.<MXI 001

;("""",, 0.000 005

10'

0 .0001

"-

0 .00005

"

0.lXIOO1

10'

106

FLOW IN PIPES AND CLOSED CONDUITS

which gives A correct to 5% for 4 x 103 < Re < "I x 107 and for k./D <
0.01.
In a more recent publication, Barr (1975) gives another explicit formulation for X:

(k, 5.1286)

1
21
.j).=- og 3.7D+ Reo.n

(4.15)

In this fo rmula the smooth law component (2.5 lIRe A) has been replaced
by a n approximation (5. 1286/ Rco.89 ). For Re > l Os this provides a solution
for S,{hf/L) to an accuracy better than I %.
However, the basic engineering objections to the use of the ColebrookWhite equation were not overcome until rhe publication of Charts for the
Hydralt!;c Desigll o(ClJalllleis and Pipes in 1958 by the Hydraulics Resea rch
Station. In this publication, the three dependent engineering variables (Q,
D and Sf) were presented in the form of a series of charts for various k,
values, as shown in Figure 4.6. Additional information regarding suitable
design values for k, and other matters was also included. Table 4.2 lists
rypical values for various materials.
These chans are based on the combination of the Colebrook-White equation (4.13) with the Darcy-Weisbach formula (4.8 ), to give

(k

2.51.)

V = - 2"r:;:;:;;;
2gDSf log - ' - + r>::n<
3.7D D,/2gDS,

(4.16)

where Sf = ",IL, the hydrau lic gradient. (Note; for further details
concerning the hydraulic gradient refer to Chapter 12. ) In this equation
the velocity (and hence discharge) can be computed directly for a known
diameter and frictional head loss.
More recently, the Hydraulics Research Station have also produced
Tables fOT the Hydraulic Design of Pipes.
In practice, any two of the three variables (Q, D and Sf) may be known,
and therefore the most appropriate solution technique depends on circumstances. For instance, in the case of an existing pipeline, the diameter and
available head are known and hence the discharge may be found direcrly
from (4. 16). For the case of a new insta llation, the available head and
required discharge are known and the requisite diameter must be found.
This witJ involve a trial-and-error procedure unless the HRS charts or
tables are used. Finally, in the case of analysis of pipe networks, the
required discharges a nd pipe diameters are known and the head loss must
be computed. This problem may be most easily solved using an explicit
formula fo r A or the I-IRS charts.

DiamelC:r (m)

Figurt 4.6

Hydraulics Research Station chart for k, = 0.03 rum.

Tablt 4.2 Typical k, valuts.


Pipt maltrial

k, (mm)

brass, copptr, glass, I'ersptx


ccrntnl

0.003

:I~ I>e$fos

wrought iron
galvanized iron
plastic
bitumen-lined ductile iron
spun concrele lined ductile iron
slimed concrtte sewer

O.oJ
0.06
0.15
0.03
0.03
0.03
6.0

108

FLOW IN PIPES AND CLOSED CONDUITS

Examples illustrating the applica tion of the various methods to the solution of :1 simple pipe friction problem now follow.
Example 4.2

EstimatiOlt of discharge givell diameter Qltd head loss

A pipeline IOkm long, 300mm in diameter and with roughness size 0.03 mm,
conveys water from a reservoir (top water level 850111 300ve datum) 10 a water
treatment plant (inlet water level 700 m above datum). Assuming that the re~rvoir
remains fuJI, rslimate the discharge, using the following methods:
(a) the Colebrook-Whitt formula;
(b) the Moody diagram;
(e) the HRS chartS.
Note: Assume v = 1.13 x 10-' ml/s.

So/lltion
(a) Using (4. 16),

D=O.3m

k.=O.OJmm

Sf = (850- 700)/10000 = 0.0 15

hence

v=

-2,flg)( 0.3 )( 0.01510g ( 0.03 x IO-J


3.7)( 0.3

+ ;;~2.~5F,1~X~I~.IFJFX~I~O~-TI" )
O.3J2g x 0.3 x 0.0 15

= 2.5 14 mls

=A=

2.5 14 )( 'II"xO.3

0 78

=. l m /s

(h) The same solution should be obtainable using the Moody diag ram; however,
it is less accurate since it involves interpolation from a graph. The solution method

is as follows:
(1) ca lculate ksl D
(2) guess a value for V

(3) calculate Re
(4) estimate A using the M oody d iagram
(5) calculate h(
(6) compare h, with the available head (H)
(7) if H". hr' then repeat from step 2.

109

11}RBUl.ENT Fl.OW

This is a tedious solution technique, but it shows why the HRS charts were
produced!

(I ) kslD = 0.03 x 10-)/0.3 = 0.0001.


(2) As the solution for V has already been found in part (a) take V = 2.5 m/s.
(3)

DV
0.3 x 2.5
Re = -,- - '1~.1~3~x::-';'I';O-:"

= 0.664 x 10 6
(4) Referring to Figure 4.5, Re = 0.664 x 106 and k,ID = 0.000 1 confirms that
the flow is in the transitional turbulent region. I:ollowing the k, ID curve until
it intersects with Re yields
A::=0.014

(Note: Interpolation is difficult due


(5) Using (4 .8),

10

the logarithmic scale.)

IJ =ALV1 =0.0 14xl0 x2.5 1

'2gD

2gxO.3

= 148.7m

(6) H-(850-700)-150,. 148.7.


(7) A bener guess for V is obtained by increasing V slightly. This will not
significa ntly alter A, but will increase hr. In this instance, convergence to the
solution is rapid because the corree! solution for V was assumed initially!
(c) If the HRS chart shown in Figure 4.6 is used, then the solution of the equation
lies at the interseetion of the h)'draulic gradient line (sloping downwards right to left)
with the diameter (venical), reading off the corresponding discha rge (line sloping
downwards left to right).

s, = 0.0 15

1ODS, = J.5
and D=300mm
giving Q = 180 lIs = 0. 18m 1 /5

Examp le 4.3

EstimatiOll of pipe diameter give" discharge al1d head

A discharge of 400Vs is 10 be conveyed from a headworks at 1050m above datum


to a treatment plant at 1000m above datum. The pipeline length is 5 km. Estimate
the requi red diameter. assuming that ks = 0.03 mOl.

1\0

FLOW IN PIPES AND CLOSED CONDUITS

Solutio"
This (('quires an iterative solution if methods (a) or (b) of Ihe previous example are
used. However, a direct solUlion can be obtained using the HRS charts.
S,_ SO/SOOO
100S,,,,, I
and Q = 400 lIs
giving D = 440 mm
In practice. the nearest (larger) available diameter would be used (450 mill in this
case).

Example 4.4

Estimation of head loss gillen discharge and diameter

The known outflow from a hranch of a disuiburion syStem is 30 lis. The pipe
diameter is ISO mm, length 500 III and roughness coefficient estimated at 0.06 mm.
Find the head loss in the pipe, uSlOg the explicit formulae of Moody and Barr.

Solutioll
Again, the H RS charts could be used directly. Howc\'er, if Ihe analysis is being
carried out by computer, solution is more efficient using an equ:ltion.

Q "", 0.03 mJ/s,

D _ O. 15m

V = L7m/ s
Re=0.15x 1.7/ 1.1 3 x 10-'
Re=0.226 x 10'
Using the Moody formula (4.14)

~=0.OO55 ( 1+ (
~

20oooXO.06 X 10-'
1 )''')
0.15
+0.226

"" 0.0182

Using the Barr formula (4.15)

1
(0.06 X 10- 1
21
./):.= - 08 3.7xO.15
~_0.0182

5.1286)

+ (0.226 x 10')0."

111

TURBULENT FLOW

The accuracy of Ihese formu lae may be compared by substituting in Ihe


Colebrook-While equarion (4.13) as follows:

0.0182

1/ ,ff;

2.51)
I
ks
-2log ( 3.7D + Re....'A =....'A

7.415

7.441

This confirms Ihat both formu lae are accurale in Ihis case.
The head loss may now be compulcd using the Darcy-Weisbach formula (4.8):

" = 0.0 182x500x 1.72


f
2gxO.15
= 8.94m

The Hazen- \Villiams formula


The emphasis here has been placed on the development and use of the
Colebrook-White transition formula. Using the charts or table~ it is simple
to apply to single pipelines. However, for pipes in series or parallel or for
the more general case of pipe networks it rapid ly becomes impossible to use
for hand calculations. For this reason, simpler empirical formulae are sti ll
in common use. Perhaps the most notable is the Hazen-Williams formula,
which takes the form

or, alternatively,

_ 6.78L

hf - D1.I65

(V)""
C

where C is a coefficient. The va lue of C varies from about 70 to 150,


dependi ng on pipe diameter, material and age.
This formula gives reasonably accurate results over the range of Re
commonly found in water distribution systems, and because the value of
C is assumed to be constant, it ca n be easily used for hand calculation. In
reality, C should change with Re, and caution shou ld be exercised in its use.
An interesting problem is to compare the predicted discharges as calculated
by the Colebrook-White equation and by the Hazen-Williams formula
over a large range of Re fo r a given pipe. The use of a microcomputer is
recommended fo r this exercise.

112

4.6

FLOW IN PIPES AND CLOSED CONDUITS

local hc=ad losses

Head losses, in addition to those due to friction, are always incurred at


pipe bends, junctions and valves, etc. These additional losses are due to
eddy formatio n generated in the fluid at Ihe fining, and, for completeness,
they must be taken into accou nt. in the case of long pipelim:s (e.g. severa l
kilometres) the local losses may be negligible, but for short pipelines, they
may be greater than the frictiona l losses.
A general theoretical treatment for local head losses is not available. It is
usual to assume rough turbulence si nce this leads to the si mple eq uation
(4. 17)
where hl is the local head loss and kl. is a constant for a particular fitting.
For the particular case of a sudden enlargement (for insta nce, exit from a
pipe to a rank) an expression may be derived for kL in terms of the area of
the pipe. This result may be extended to the case of a sudden contr::J.ction (for
instance, entry to a pipe from :\ tank). For all other cases (e.g. bends, valves,
junctions, bellmouths, etc.) val ues for kl must be derived experimenta lly
Figure 4.7(a) shows the case of a sudden enlargement. From position (1)
to (2) the velocity decreases and therefore the pressure increases. At position

I~

I~
I

CDta) AI tvckkn

-/f
I
CD

(I))

(II
~nla rttmt11 1

(II

AI. s\tdckft conlraction

Figure 4.7

Local head loss.

LOCAL H EAD LOSSES

113

(I ') turbulent eddies are formed, which gives rise to a loca l energy loss. As
the pressure ca nnOI change instantaneously at the sudden enlargement, it is
usua lly assumed that at position (1') the pressure is the same as at position
(1). Applying Ihe momentum equation between (I) and (2).

The continuity equation (Q = A2 V2) is now used to eliminate Q. so, with


some rearrangement,

(.)

The local head loss may now be found by applying the enerb'>' equation
f<om (I) ro (2),

or

(b)

If (a) and (b) are combined and rearranged,

The continuity equation may now be used again to express the result in
terms of the ~\l0 areas. Hence, substituting V.A. / A 2 for V2

0'

(4.18)

114

FLOW IN PIPES AND CLOSED CONDUITS

Equation (4.18 ) relates hl to the areas and rhe upstream velocity.


Comparing this equation with (4 . 17) yields

For the case of a pipe discharging into a tank, A2 is much grea ter tha n AI>

and hence kl = 1. In o ther words, fo r" sudden large expansion, the head
loss equals the velocity head before expansion.
Figure 4.7(b) shows the case of a s udden contraction. From posi tio n (1)
to ( 1') the flow contracts, forming a vena contracta. Experimenrs indicate

that the contraction of the fl ow a rea is generally about 40%. If the energy
loss from {I) to {1 ') is assumed to be negligible, then the rema in ing head

loss occurs in the expansion from (I') to (2). Si nce an expansion loss gave
rise to (4.18), that equation may now be applied here. As

then

0'

b L = 0.44 Vi /2g

(4. 19)

i.e. kL = 0.44 .
Typical kL va lues for other important local losses (bends, tees, bdlmouths
a nd valves) are given in Table 4.3.

Table 4.3

Local head loss coefficients.

kL value
Item

Theoretical

Design practice

Comments

0.05
0.2
0.4

0.10
0.5
0.5

v = velocity in pipe

exil
90 bend
90" " ..
in-line flow
branch to lint
gate valve (open)

0.35
1.20
0.12

0.4

(for equa l diameters)

1.5

(for equal diameters)

bdlmoUlh entrance

0.25

LOCAL HEAD LOSSES

11 5

Example 4.5 Discharge caiclilation (or a simple pipe system ifreluding


local losses
Solve Example 4.2 allowing fo r local head losses incurred by the following items:
20 90 bends
2
gate valves
1
bdlmouth entry
bdlmouth exit

SO/lllio"
The available static head ( 150m) is dissipated bOlh by friction and local head losses.
Hence

Using Table 4.3,


III = ((20 x 0.5)+ (2)( 0.25)+0.1 +0.5]yl j2g
= 11.1 ylj2g

Using the Colebrook-White formula (as in Example 4.2) now requires an iterative
solution, since h is initially unknown. A solution procedure is as follows:
( I) assume h, ~ 1-1 (i.e. ignore
(2) calculate Y

lid

(3) calculate hl using V


(4) calculate h(+hl
(5) if h(+hl =F 1-1, set h( '" H - III and return to (2)
Using Example 4.2, an initial solution fo r Y has already been found, i.e.
V=2.5 14mjs
Hence,

Adjust II,
11(=

150-3.58_ J46.42m

Sf = 146.42j10 000 = 0.0 1464

116

FLOW IN PIPES AND CLOSED CONDUITS

Substitute in (4.1 6),

V = - 2/2g x 0.3 x 0.0146410g (

0.03 x 10-1
2.51 x 1.13 x 10- ' )
+ ;;-;=:;;~~E~~'"
3.7xO. 15 O.3j2gxO.3xO.O I4 64

=2.386mj s
Recalculate hL'

Check
hL +hf = 146.42+3.22 = 149.64 :::: 150
This is sufficiently acc urate to be acceptable.
Hence,

Note; Ignoring "L gives Q = 0. 18 m1 / s.

4.7

Partially full pipes

Pipe systems for surface water drainage and sewerage are normally designed
to flow full, but not under pressure. This contrasts with water mains, which
are normally full and under pressure. The Colebrook-White equation ma y
be used for drainage pipes by noting that, because the pipe flow is not
pressurized, the water su rface is parallel to the pi pe invert, so the hydraulic
gradient equals the pipe gradien t:

where So is the pipe gradient.


Additionally, an estimate of t he discharge and velocity for the partially
fu ll condition is required. This enables the engineer to check if self-cleansing
velocities are maintained at the minimum discharge. Selfcleansing velocities are of crucial importance in the design of surface water drainage and
sewerage networks, where the flow may conta in a considerable suspended
solids load.
A free surface flow has one more variable t han full pipe flow, namely the
height of the free surface. This can introduce considerable complexity (refer
to Chapter 5). However, for the case of circular conduits, the ColebrookWhite equation may be modified to provide a solution.

11 7

PARTIAU Y FULl PIPES

Starting from the assumption that the frictio n facto r for the partially full
condition behaves similarly to that for the fu ll condi tion, it remains to fi nd
a parameter for the partially fu ll pipe which is equiva lent to the dia meter
for the full pipe case. The hydraulic radi us R is such a parameter:

where A is the water cross-sectiona l area and P is the wetted peri meter. For
a pipe flowing full,

A/ P = 1I"D2/ 411"D = D/4


0'

Hence the Colebrook-White transition law applied to partiall y full pipes


becomes

(k,

I
21
fi. = - og 3.7 x 4R

2.51 )

(4.20)

+ Refi.

where Re = 4RV/ II.


Figure 4.8 shows a pipe with partia lly full fl ow (al a depth d). Sta rting
from the Darcy-Weisbach equation (4.8) and replaci ng h(/ I.. by So gives

V' = 2g5,DI"
Hence, for a given pipe with partiall y fu ll flow,

V = (2g5, 4RI")'"

Figure 4.8

Pipe running parlially full.

118

FLOW IN PIPES AND CLOSED CONDUITS

0'

v=

constant R 1/2 IX 1/2

Forming the ratio V,,/ YD = Vp gives


V, ,_}..
1/2 RI/l/ X'il
- D
,.
II

(4.21)

where the subscripts p, D and d refer, respectivdy, to the proportional

value, the full depth (D) and the partially full depth (eI). Simi larl y,
R1 / Z/}..I/z
Q, ._}..I/lA
- D
p,.
Ii

(4.22)

For a circular pipe,


Ad =

('I> - ;;n <I> ) D'

Pd = ",0/2

R, =

(I _S;;"') ~

and hence

A,= ("'-2:""')

(4.23)

S
R,=(I _ ;;"')

(4.24)

Substitution of (4.23) and (4.24) into (4.2 1) and (4.22) allows calculation

of the proportional velocity and discharge for any proportional depth (dID).
The expression for}.. (equation (4.20)) is, however, rather awkwa rd to

manipulate. Consider first the case of rough turbulence. Then,


_1_ =2 10g (3.70)
,f'A
ks
Hence.

Fo

210g(3.7 x 4R,/ks)
,;;;; = 210g(3.70/k s)

This may be expressed by its equ iva lent:

Fo
10gR,
-';;;;-, = I + ;lo-g(~3."'7""0"'/k'"'s)

(4.25)

11 9

PARTIALLY FULL PIPES

as

1+

10gR,
log(3.7D/ks)+logR,
=
log(3 .7D/ ks}
log(3.7 D/ks}

log[ (3. 7 D/ ksllR,/ Ro) I


log(3.7D/ ks}

log[3.7 x 4R,/ksl
log(3.7D/ ks}

Equation (4.25 ) may be substituted into (4.21) and (4.22) to yield


10gR,
) 'I'
V, = ( 1+ log(3.7 D/ ks ) R,
and
10gR,

Qp = ( 1+ log(3.7D/k,}

' f2
ApRp

The equivalent ex pressions for the transition region (as derived in


Hydraulics Research Paper No. 2, published in 1959) are
_ (

V,_

10gR, ) l{l
1+ 10g3.70 R,

(4.26)

and

Q = ( 1+ 10gR)
p
ARI/l
P
log 3.70
p p

(4.27)

( r

(4.28)

where

a . . . . ks +
-

3600DS~/l

These resuhs for 0 = 1000 are ploned in Figure 4.9. Tabulated values for
various 0 may be found in Hydraulics Resea rch Ltd (1983a). Neither Vp
nor Q, are very sensitive to O.
Figure 4.9 shows that the discharge in a pa rtially full pipe may be greater
than the discharge for a fu ll pipe. This is because the wetted perimeter
reduces rapidly immediately the pipe ceases to be full whereas the area
does not, with a consequent increase in velocity. However, this condition
is usually ignored for design purposes because, if the pipe runs full at any

120

flOW IN "IPES AND CLOSED CONDUITS


0.0

"0.8
"0.'

Q,

e O~
~

0.'

v,

0.3
02

0. 1

01

01

0.1

0-4

OS

06

0.7

0.8

09

10

1I

12

Qd/Of) and VIIIV D

Figure 4.9

I'roporlionai discharge and velocity for pipes flowing partially full


(with a"" 1000).

section (e.g. due to wave action or unsteady conditions), Ihen the discharge
will rapidly reduce to the full pipe condition and cause a 'backing up' of
the flow upstream.

Example 4.6 Hydraulic desigl1 of a selver


A sewerage pipe is to be laid at a gradient of I in 300. The design maximum
discharge is 75 Us and the design minimum now is estimated to be IOUs. Determine
the required pipe diameu~ r 10 bOlh ca rry the maximum discharge and maimain a
sel fcleansing velocity of 0.75 mls at the minimum discharge.

Soilltioll
The easiest way 10 solve this problem is to use the HRS design ch:ms or tables. For
a sewer, ks = 6.00mm (Table 4.2). Howt:ver, to illustratt: tht: solution, Figurt: 4.6
is uSt:d (for which ks = 0.03mm):

Q= 75 1/5
IOOh, / L

= 100/300 =0.333

Using Figure 4.6

D=300mm and

V _ I .06m/s

REFERENCES AND FURTH ER READING

121

Next check the velocity for Q = 10 l/s

Q,= 10/75=0.133
Using Figure 4.9 (neglecting the effect of 0),

dID =O.25 for Q, = 0.133


H ence V, _ 0.72 and
" ,,=0.72 )( 1.06 = 0.76m/s
This val ue excttds the selfcleansing velocity, and hence Ihe solution is D= 300mm.
In cases where Ihe self-cleansing \'c!ocity is nOI maimaincd, it is necessary 10 increase
Ihe diameter or the pipe gradiem.
Note: The solution using ks = 6mm and accounting for 0 gives the following
values:
D "" 375mm for Q= 811/s and V _ O.73m/s
0 = 45
Q, = 10/8 1 = 0.123
d/D=O.024
V, = O.67m/s
V" =0.49m/ s
Hence It would be nessary to increase D or
be p~ferable.

So.

In this case, increasing So would

Rcfcre nces ~ nd further reading


Ackers, P. (J958) ResistallU of Fillids Flowing ill Cllmmels and Pipes. H yd raulics
Research )laper No. I, HMSO, London.
Barr, D. I. H. (1975) Two additional methods of dirt solution of Ihe ColebrookWhite function. Proc. Instn Giv. Engrs, 59, 827.
Colebrook, C. F. (1939) Turbulent nows in pipes, with particular reference to the
tra nsition region between the smooth and rough pipe laws. J. Instn Gill. Engrs,
t I , 133.
Colebrook, C. F. and White, C. M. (1937) Experiments with nuid friction in rough.
ened pipes. Proc. Roy. Soc., A16 1, 367.
Hydraulics Research Limited ( 1983a) Tables for 'he Hydraulic Design of Pipes,
4th edn, Thomas Telford, London.
Hyd raulics Research Limited (1983b) CI,arts fo r the Hydraulic Design of Gllallne/s
and Pipes, 51h edn, Thomas Telford, London.
Moody, L. F. (1944) Friction factors for pipe flows. Trans. Am. Soc. Mew. Engrs.,
66,671.
Webber, N. B. (1971) Fluid Mechallic:s for Civil Engineers, Chapman and Hall,
London.

Vous aimerez peut-être aussi