You are on page 1of 137

CE 376

HYDROMECHANICS

1
1. INTRODUCTION

Scope

In many water systems, transportation of water from


one location to another is main concern.
Two main modes of transportation are:
1. Closed conduits with pressurized flow inside

2. Open conduits with free surface flow inside

2
• The main objective in this course is to study the flow
in closed conduits (mainly pipes) and in open channels

3
4
5
6
INTERNAL FLOW

Fluid flow is classified as external if it is forced to flow over a


surface
OR
internal, if it is forced to flow in a conduit.

In internal flow where the conduit is completely filled with the


fluid, and the flow is driven primarily by a pressure difference.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
7
Fundamentals and Applications, McGraw Hill, 2017.
Liquid or gas flow through pipes or ducts is is usually forced to flow
by a fan or pump through a flow section. FRICTION must be paid a
particular attention here, as it is directly related to the pressure
drop and head loss during flow through pipes and ducts. The
pressure drop is then used to determine the pumping power
requirement.
A typical piping system involves pipes of different diameters
connected to each other by various fittings or elbows to route the
fluid, valves to control the flow rate, and pumps to pressurize the
fluid.
Most fluids, especially liquids, are transported in circular pipes. This
is because pipes with a circular cross section can withstand large
pressure differences between the inside and the outside
without undergoing significant distortion.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
8
Fundamentals and Applications, McGraw Hill, 2017.
Noncircular pipes are usually used in applications such as the
heating and cooling systems of buildings where the pressure
difference is relatively small.

The theory of fluid flow is reasonably well understood. However,


theoretical solutions are obtained only for a few simple cases such
as fully developed laminar flow in a circular pipe.
Experimental results and empirical relations for most fluid flow
problems are much preferred than analytical solutions. However
experimental results are not exact as well.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
9
Fundamentals and Applications, McGraw Hill, 2017.
The fluid velocity in a pipe changes from zero at the wall because of the
no-slip condition to a maximum at the pipe center. In fluid flow, it is
convenient to work with an average velocity Vavg, which remains constant in
incompressible flow when the cross-sectional area of the pipe is constant

The average velocity in heating and cooling applications may


change somewhat because of changes in density with temperature. But, in
practice, we evaluate the fluid properties at some average temperature and
treat them as constants. The convenience of working with constant
properties usually more than justifies the slight loss in accuracy.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
10
Fundamentals and Applications, McGraw Hill, 2017.
The value of the average velocity Vavg at some streamwise cross
section is determined from the requirement that the conservation of
mass principle be satisfied. That is,

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
11
Fundamentals and Applications, McGraw Hill, 2017.
12
13
THE ENTRANCE REGION
Consider a fluid entering a circular pipe at a uniform velocity. Because of the no-slip
condition, the fluid particles in the layer in contact with the Wall of the pipe come to a
complete stop. This layer also causes the fluid particles in the adjacent layers to slow
down gradually as a result of friction. To make up for this velocity reduction, the velocity of
the fluid at the midsection of the pipe has to increase to keep the mass flow rate through
the pipe constant. As a result, a velocity gradient develops along the pipe.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
14
Fundamentals and Applications, McGraw Hill, 2017.
The region from the pipe inlet to the point at which the velocity profile is fully developed is
called the hydrodynamic entrance region, and the length of this region is called the
hydrodynamic entry length Lh. Flow in the entrance region is called hydrodynamically
developing flow since this is the region where the velocity profile develops. The region
beyond the entrance region in which the velocity profile is fully developed and remains
unchanged is called the hydrodynamically fully developed region. The flow is said to be
fully developed when the normalized temperature profile remains unchanged as well.
Hydrodynamically fully developed flow is equivalent to fully developed flow when the fluid
temperature remains constant. The velocity profile in the fully developed region is
parabolic in laminar flow and is much flatter (or fuller) in turbulent flow due to eddy motion.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
15
Fundamentals and Applications, McGraw Hill, 2017.
Entry Lengths
The hydrodynamic entry length is usually taken to be the distance from
the pipe entrance to where the wall shear stress (and thus the friction
factor) reaches within about 2 percent of the fully developed value.

The pipes used in practice are


hydrodynamic usually several times the
entry length for length of the entrance region,
laminar flow and thus the flow through the
pipes is often assumed to be
hydrodynamic
fully developed for the entire
entry length for
length of the pipe.
turbulent flow
This simplistic approach gives
hydrodynamic entry reasonable results for long
length for turbulent pipes but sometimes poor
flow, an approximation results for short ones since it
underpredicts the wall shear
stress and thus the friction
factor. 16
LAMINAR AND TURBULENT FLOW

The transition from laminar to turbulent flow does


not occur suddenly; rather, it occurs over some
region in which the flow fluctuates between
laminar and turbulent flows before it becomes
fully turbulent.
Most flows encountered in practice are
turbulent. Laminar flow is encountered when
highly viscous fluids such as oils flow in small
pipes or narrow

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
17
Fundamentals and Applications, McGraw Hill, 2017.
LAMINAR AND TURBULENT FLOW

The existence of these laminar, transitional,


and turbulent flow regimes can be verified by
injecting some dye streaks into the flow in a
glass pipe, as the British engineer Osborne
Reynolds (1842–1912) did over a century
ago. We observe that the dye streak forms a
straight and smooth line at low velocities
when the flow is laminar (we may see some
blurring because of molecular diffusion), has
bursts of fluctuations in the transitional
regime, and zigzags rapidly and disorderly
when the flow becomes fully turbulent.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
18
Fundamentals and Applications, McGraw Hill, 2017.
These zigzags and the dispersion of
the dye are indicative of the
fluctuations in the main flow and the
rapid mixing of fluid particles from
adjacent layers.

The intense mixing of the fluid in


turbulent flow as a result of rapid
fluctuations enhances momentum
transfer between fluid particles, which
increases the friction force on the
pipe wall and thus the required
pumping power. The friction factor
reaches a maximum when the flow
becomes fully turbulent.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
19
Fundamentals and Applications, McGraw Hill, 2017.
20
REYNOLDS NUMBER

The Reynolds number can be viewed as the ratio of inertial forces


to viscous forces acting on a fluid element.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
21
Fundamentals and Applications, McGraw Hill, 2017.
The transition from laminar to turbulent flow depends on the
geometry, surface roughness, flow velocity, surface
temperature, and type of fluid, among other things.

After exhaustive experiments in the 1880s, Osborne Reynolds


discovered that the flow regime depends mainly on the ratio of
inertial forces to viscous forces in the fluid.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
22
Fundamentals and Applications, McGraw Hill, 2017.
Vavg = average flow velocity (m/s),
D = characteristic length of the geometry (diameter in this case, in m)
= / = kinematic viscosity of the fluid (m2/s)

Note that the Reynolds number is a dimensionless quantity.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
23
Fundamentals and Applications, McGraw Hill, 2017.
The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
24
Fundamentals and Applications, McGraw Hill, 2017.
FULLY DEVELOPED LAMINAR FLOW

25
From F=ma

26
27
28
29
30
31
32
33
Exp:

34
35
From Dimensional Analysis

36
37
38
In practice, it is convenient to express the pressure loss for all types of fully
developed internal flows (laminar or turbulent flows, circular or
noncircular pipes, smooth or rough surfaces, horizontal or inclined pipes)
as:

It is also called the Darcy–Weisbach friction factor, named after the


Frenchman Henry Darcy (1803–1858) and the German Julius Weisbach
(1806–1871), the two engineers who provided the greatest contribution to its
development.
The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
39
Fundamentals and Applications, McGraw Hill, 2017.
The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
40
Fundamentals and Applications, McGraw Hill, 2017.
Head Loss (Energy Considerations)

41
In the analysis of piping systems, pressure losses are commonly expressed
in terms of the equivalent fluid column height, called the head loss hL.
Noting from fluid statics that ∆P = gh and thus a pressure difference of
∆P corresponds to a fluid height of h = ∆P/ :

The head loss hL represents the additional height that the fluid needs to be
raised by a pump in order to overcome the frictional losses in the pipe. The
head loss is caused by viscosity, and it is directly related to the wall shear
stress. Above equation is valid for both laminar and turbulent flows in
both circular and noncircular pipes.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
42
Fundamentals and Applications, McGraw Hill, 2017.
TURBULENT FLOW
Most flows encountered in engineering practice are turbulent. Turbulent flow is
characterized by disorderly and rapid fluctuations of swirling regions of
fluid, called eddies, throughout the flow. These fluctuations provide an
additional mechanism for momentum and energy transfer. In laminar flow,
fluid particles flow in an orderly manner along pathlines, and momentum and
energy are transferred across streamlines by molecular diffusion.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
Fundamentals and Applications, McGraw Hill, 2017. 1
2
In turbulent flow, the swirling eddies transport mass, momentum, and
energy to other regions of flow much more rapidly than molecular diffusion,
greatly enhancing mass, momentum, and heat transfer. As a result, turbulent
flow is associated with much higher values of friction, heat transfer, and
mass transfer coefficients.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
3
Fundamentals and Applications, McGraw Hill, 2017.
Even when the average flow is steady, the eddy motion in turbulent flow causes significant
fluctuations in the values of velocity, temperature, pressure, and even density (in
compressible flow). The figure shows the variation of the instantaneous velocity component u
with time at a specified location, as can be measured with a hot-wire anemometer probe or other
sensitive device. We observe that the instantaneous values of the velocity fluctuate about an
average value, which suggests that the velocity can be expressed as the sum of an average value
u– and a fluctuating component u′.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
4
Fundamentals and Applications, McGraw Hill, 2017.
5
6
Total Shear Stress in Turbulent Flow

7
Turbulent Shear Stress
turbulent shear stress

Turbulent shear
stress

eddy viscosity or turbulent viscosity:


accounts for momentum transport by
turbulent eddies

Total shear
stress

kinematic eddy viscosity or


kinematic turbulent viscosity (also
called the eddy diffusivity of
momentum).
8
mixing length lm: related to the average
size of the eddies that are primarily
responsible for mixing

Molecular diffusivity of
momentum v (as well as
µ) is a fluid property, and
its value is listed in fluid
handbooks.
Eddy diffusivity vt (as well
as µt), however, is not a
fluid property, and its
The velocity gradients at the value depends on flow
wall, and thus the wall shear conditions.
stress, are much larger for Eddy diffusivity µt
turbulent flow than they are decreases toward the wall,
for laminar flow, even though becoming zero at the wall.
the turbulent boundary layer Its value ranges from zero
is thicker than the laminar at the wall to several
one for the same value of thousand times the value
free-stream velocity. of the molecular diffusivity
in the core region. 9
lm is the mixing length.

L. Prandtl proposed that turbulent process could be viewed as the


random transport of bumdles of fluid particles over a certain length,
lm.

10
11
Turbulent Velocity Profile The very thin layer next to the wall where
viscous effects are dominant is the viscous (or
laminar or linear or wall) sublayer.
The velocity profile in this layer is very nearly
linear, and the flow is streamlined.
Next to the viscous sublayer is the buffer
layer, in which turbulent effects are becoming
significant, but the flow is still dominated by
viscous effects.
Above the buffer layer is the overlap (or
transition) layer, also called the inertial
sublayer, in which the turbulent effects are
much more significant, but still not dominant.
Above that is the outer (or turbulent) layer in
the remaining part of the flow in which
turbulent effects dominate over molecular
diffusion (viscous) effects.

The velocity profile in fully developed pipe flow is parabolic in laminar flow,
but much fuller in turbulent flow. Note that u(r) in the turbulent case is the
time-averaged velocity component in the axial direction (the overbar on u
12
has been dropped for simplicity).
friction velocity
law of the wall

Law of the wall is valid very near the smooth wall for 0≤yu*/ʋ≤5
In the viscous sublayer, the fluid viscosity is the important parameter. The
thickness of the viscous sublayer is proportional to the kinematic viscosity and
inversely proportional to the average flow velocity.

for 5≤yu*/ʋ≤70

13
Velocity
defect law
The deviation of velocity from the centerline value
umax - u is called the velocity defect.

The value n = 7 generally


approximates many flows in
practice, giving rise to the
term one-seventh power-law
velocity profile.

Power-law velocity profiles


for fully developed
turbulent flow in a pipe for
different exponents, and its
comparison with the
14
laminar velocity profile.
Comparison of Laminar & Turbulent Flow

15
Comparison of Laminar & Turbulent Flow
Laminar Turbulent
Can solve exactly Cannot solve exactly (too complex)
Flow is unsteady, but it is steady in the
Velocity profile is parabolic
mean
Pipe roughness does not affect
Mean velocity profile is fuller
the flow Pipe roughness is very important
Vavg 85% of Umax (and depends on Re )
No analytical solution, but there are
some good semi-empirical expressions
that approximate the velocity profile
shape.

Instantaneous
profiles

16
Major Losses (Turbulent Flow)

17
18
19
20
Darcy Weisbach Equation
21
Turbulent Flow (Re>4000)

δs = thickness of the
Velocity profile, viscous sublayer
R=D/2
y u = u f(y) ε = the roughness of
δs pipe wall
x

δs ε ε δs
ε

(a) Smooth wall (b) Transitional flow (c) Rough wall


22
Turbulent Flow (Re>4000)

δs ε ε δs
ε

(a) Smooth wall (b) Transitional flow (c) Rough wall

For hydraulically smooth pipe f=f(Re) only


For frictionally transition zone f=f(Re, ε/D)
For fully rough pipe f=f(ε/D) only.

23
24
THE MOODY CHART
The friction factor in fully developed turbulent pipe flow depends on the
Reynolds number and the relative roughness /D, which is the ratio of the
mean height of roughness of the pipe to the pipe diameter.

The functional form of this dependence cannot be obtained from a theoretical


analysis, and all available results are obtained from experiments using
artificially roughened surfaces (usually by gluing sand grains of a known size
on the inner surfaces of the pipes).

In 1939, Cyril F. Colebrook (1910–1997) combined the available data for


transition and turbulent flow in smooth as well as rough pipes into the
following implicit relation known as the Colebrook equation:
The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics:
25
Fundamentals and Applications, McGraw Hill, 2017.
26
27
28
We make the following observations from the Moody chart:

•For laminar flow, the friction factor decreases with increasing Reynolds number,
and it is independent of surface roughness.

•The friction factor is a minimum for a smooth pipe (but still not zero because of
the no-slip condition) and increases with roughness. The Colebrook equation in
this case ( = 0) reduces to the Prandtl equation expressed as 1/√f = 2.0
log(Re√f ) − 0.8.

•The transition region from the laminar to turbulent regime (2300 < Re < 4000) is
indicated by the shaded area in the Moody chart. The flow in this region may be
laminar or turbulent, depending on flow disturbances, or it may alternate between
laminar and turbulent, and thus the friction factor may also alternate between the
values for laminar and turbulent flow. The data in this range are the least reliable.
At small relative roughnesses, the friction factor increases in the transition region
and approaches the value for smooth pipes.

29
• At very large Reynolds numbers (to the right of the dashed line on the
Moody chart) the friction factor curves corresponding to specified relative
roughness curves are nearly horizontal, and thus the friction factors are
independent of the Reynolds number.

• The flow in that region is called fully rough turbulent flow or just fully rough
flow because the thickness of the viscous sublayer decreases with increasing
Reynolds number, and it becomes so thin that it is negligibly small compared
to the surface roughness height.

• Some authors call this zone completely (or fully) turbulent flow, but this is
misleading since the flow to the left of the dashed blue line in is also fully
turbulent.
30
Friction Losses in Pipes, hf
hf – Friction (Viscous, Major) loss

Determination of Friction Loss (hf):

1. Darcy-Weisbach Equation

8fL
LV 2
L 16 Q2
where K=
hf = f =f 5 2 = KQ 2 gπ2D5
D 2g D π 2g

31
2. Hazen-Williams Equation

6.8 L 1.85 10.6 L


hf = 1.85 1.165 V = 1.85 4.87 Q1.85 = KQ1.85
C D C D

D – pipe diameter (m)


V – average velocity (m/s)
g – gravitational acceleration (m/s2)
Q – flow rate (discharge) (m3/s)
L – pipe length (m)
f – Darcy – Weisbach friction factor (unitless)
C – Hazen-Williams Coefficient of Roughness (unitless)
32
Darcy-Weisbach Friction Factor in
Turbulent Flow
Smooth Pipe - 1  2.51 
= −2 log
 Re f

 f = func(Re )
Hydraulically Smooth Flow f  

Colebrook – 1  2.51 ε   ε 
= −2 log +  f = func Re, 
White Transitional Flow    D
f  Re f 3.7 D 

Rough Pipe - 1  ε  ε 
Hydraulically Rough Flow
= −2 log  f = func 
f  3. 7 D  D

All of these eqn.s are implicit (f on both RHS and LHS)



hard to solve SO INSTEAD use Swamee-Jain Eqn. 33
Roughness
Coefficients for
Hazen-Williams,
Manning and Darcy-
Weisbach Equations

6 .8 L
hf = 1.85 1.165
V1.85
C D

L V2
hf = f
D 2g

34
Swamee-Jain Formula (Explicit)
1.325
f = 2
  ε 5.74 
ln 3.7 D + Re 0.9 
  

for the range of 10 −6 < ε / D < 10 −2 and 5000 < Re < 108

35
Computation of Flow in Single Pipes
The flow computation in single pipes requires solution
of three equations simultaneously:

The energy equation: P1 V12 P2 V22


z1 + + α = z2 + +α + hf
γ 2g γ 2g

Equation of Contunity: Q =V1A1 =V2 A2

Darcy-Weisbach Equation: L V2
hf = f
D 2g

36
Computation of Flow in Single Pipes
• In general, there are 3 types (I, II and III) of problems
depending on the information given:

1. Find ″head loss″ problem Type I


Given : Q, L, D,ν , ε Find : h f
2. Find ″discharge ″ problem Type II
Given : h f , L, D,ν , ε Find : Q

3. Find ″diameter″ problem (design problem) Type III


Given : h f , L, Q,ν , ε Find : D
37
G- Given,
D-Determined

38
Determination of Head Loss (Type I)
Given : Q, L, D,ν , ε Find : h f
H1 = H 2 + hL and hL = hm + h f since hm = 0, hL = h f ⇒ h f = H1 − H 2
V2
p f LV 2 8f L 2
H = z+ + and h f = since V = Q / A ⇒ h f = Q
γ 2g D2 g gπ D
2 5

Q
1) Calculate velocity V=
π D2 / 4
2) Calculate Reynolds number VD
Re =
υ
3) Calculate relative roughness ε /D
4) Calculate the friction factor, f = f (Re, ε /from
D ) Moody Chart or
equations.

5) Calculate head loss from Darcy-Weissbach Eqn. 8f L


hf = Q 2

gπ 2 D 5
39
Determination of Velocity (Discharge)
(Type II) Given : h f , L, D,ν , ε Find : Q
H1 = H 2 + hL and hL = hm + h f since hm = 0, hL = h f ⇒ h f = H1 − H 2

f LV 2 f LV 2 D2g
hf = thus H1 − H 2 = ⇒V = (H 1 − H 2 ) (1)
D2g D2 g fL

D2 g
Equation (1) V = (H 1 − H 2 )
fL

f = f (Re, ε / D )

VD
Re =
There is V at both sides of Eqn. (1)! υ
40
Determination of Average Velocity
(Type II) Given : h , L, D,ν , ε f Find : Q

1) Calculate relative rougness, ε/D


2) Make an initial guess for friction factor, f
assume completely rough turbulent flow for which f=f(ε/D) only
& f(i)=f(0). h D2 g f
V=
3) Calculate velocity fL
VD
4) Calculate Re Re =
ν
5) Determine f(i+1) using ε/D from Step 1 and Re from Step 4
using Moody chart or Equations
6) Check if f(i+1)= f(i). If no, go to Step 3 with f(i+1).
If yes,
πD 2
7) calculate Q Q =V
4 41
Iteration Table for Type II
Given : h f , L, D,ν , ε Find : Q

h f D2 g VD
f(i) V= Re = f(i+1)*
fL ν
f(0) assumed calculated calculated f(1)

f(1) calculated calculated f(2)

f(2) calculated calculated f(3)


. . . .
. .
iteration .
stops when .
f(i) = f(i+1)
. . . .

f(i) f(i+1)

* obtained from Moody Chart [f=f(ε/D, Re)] , or determined using equations.


42
Determination of Pipe Diameter (Type III)
Given : h f , L, Q,ν , ε Find : D
1) Make an initial guess for friction factor, f
assume that f(i), i=0 f(0)=0.015 or 0.020, which is a middle value
in Moody chart. 1/ 5
8 LQ 2
 8LQ 2 
2) Calculate pipe diameter D=5 f =  2 
 f 1/ 5
hlπ g  hlπ g 
2
VD 4Q
3) Calculate Re Re = =
ν πDν
4) Calculate relative rougness ε/D
5) Determine f(i+1) using ε/D from Step 4 and Re from Step 3
using Moody chart or Equations
6) Check if f(i+1)= f(i). If no, go to Step 2 with f(i+1).
If yes, stop
7) Diameter calculated at Step 2 is the result
8) Select the next larger commercially available diameter 43
Iteration Table for Type III
Given : h f , L, Q,ν , ε Find : D

1
 8LQ  1 VD
ε /D
2 5
f(i) D =   f 5 Re = f(i+1)*
2  ν
 hlπ g 
f(0) assumed calculated calculated calculated f(1)

f(1) calculated calculated calculated f(2)

f(2) calculated calculated calculated f(3)


. . . . .
. .
iteration .
stops when f(i) = f.(i+1) .
. . . . .

f(i) f(i+1)

* obtained from Moody Chart [f=f(ε/D, Re)] , or determined using equations.


44
MINOR LOSSES
Head Losses in Pipes
Total Head Loss, hl

h l = h f + hm

hf – Friction (Viscous, Major) loss


hm– Local (Minor) loss

2
Local (Minor) Losses
For any pipe system, in addition to the friction loss
computed for the length of pipe, there are additional so-
called minor losses due to
Pipe entrance or exit
Sudden expansion or contraction
Bends, elbows, tees, and other fittings
Valves, open or partially closed
Gradual expansions or contractions

The losses may not be so minor; e.g., a partially closed


valve can cause a greater pressure drop than a long pipe.
The minor loss coefficient is usually given as a ratio
of the head loss hm through the device to the velocity
head V2/(2g) of the associated piping system.
hm
Km = 2
V
2g
2
V
hm = K m
2g
Values of K have been determined experimentally for
various fittings and geometry changes.
One exception for this is the sudden expansion from
an area A1 to A2.
Sudden Expansion

C.V.

Continuity Eqn: Q = V1A1 = V2 A 2


Momentum Eqn. in x-direction:
( p1 − p2 ) A2 = ρQ(V2 − V1 )
( p1 − p2 ) 1 2
= (V2 − V1V2 )
γ g
Energy Eqn. btw (1) and (2)
2 2
p V p V
z1 + 1 + 1 = z 2 + 2 + 2 + hm
γ 2g γ 2g

2 2
p −p V − V2
hm = 1 2 + 1
γ 2g

Using above equations,


2
(V2 − V1 ) 2 V1 A1 V1 D1
hm = , V2 = = 2
2g A2 D2
2
 D  2
 V2
hm =  1  − 1 1
 D2   2 g

Km
Sudden Contraction
The flow separates at the corners, and the flow is constricted into the
vena contracta region formed in the midsection of the pipe.

The above text and figures are taken from: Çengel, Y. A. & Cimbala, J.M., Fluid Mechanics: Fundamentals
and Applications, McGraw Hill, 2017.
Minor loss coefficients for some transitions and fittings (1)
Minor loss coefficients for some transitions and fittings (2)
Ball valve
PIPELINE SYSTEMS
Pipes in Series and Equivalent Pipes

Pipes A and B are connected in series, pipe C is the


equivalent of them.
hLA + hLB = hLC
Q A = QB = Q C = Q
8Q2 fcLc 8Q2 fAL A 8Q2 fBLB
2 5
= 2 5
+ 2 5
gπ Dc gπ D A gπ DB

fc L c f A L A fBL B fc, Lc, Dc are the equivalent friction


5
= 5 + 5
Dc DA DB factor, equivalent pipe length, and
equivalent diameter, respectively.
if fA=fB=fc=, then

LC L A LB
5
= 5 + 5
DC DA DB

Generalizing for n pipes in series:


L eq n Li
5
=∑ 5
Deq i=1Di

Minor losses may be expressed as equivalent lengths


and added to the actual lengths of pipe:
Leq V 2 V2 KmD
f = Km Leq =
D 2g 2g f
Pipes in Parallel
Consider the parallel piping arrangement shown below
Pipes in Parallel and Equivalent Pipes

Parallel pipes A and B Pipe C is the


equivalent
hL1− 2 = hLA = hLB

Q = Q A + QB
5 1/ 2 5 1/ 2
 hLA gπ2
DA   hLBgπ
2
DB 
Q A =  
 and → QB =  
 8 f AL A  
 8fBLB 
5 1/ 2 5 1/ 2 5 1/ 2
 hLCgπ DC 
2
 hLA gπ 2
DA   hLBgπ
2
DB 
  =   +  
 8f L   
 C C   8f A L A   8fBLB 

1/ 2 1/ 2 1/ 2
 D5c  D5A DB5 
  =   +  
f L  f L  f L 
 C C  A A  B B

if fA=fB=fC, then

5 1/ 2 5 1/ 2 5 1/ 2
 Dc   DA   DB 
  =   + 
L  L  L 
 C  A   B 
5 1/ 2 5 1/ 2
 Deq  n  Di 
Generalizing for n pipes in parallel:    =∑   
L  i =1  Li 
 eq 
Exp
Consider the three-pipe series system as shown
below. The total pressure drop is pA-pB = 150 kPa,
and the elevation drop is zA-zB = 5 m. The pipe data
are Pipe L (m) D (cm) ε (mm)
1 100 8 0.24
2 150 6 0.12
3 80 4 0.20

The fluid is water (ρ= 1000 kg/m3, ν= 1.02x10-6 m2/s).


Calculate the flow rate Q (m3/hr). Neglect minor
losses.
Exp
Assume that the same three pipes of previous
example are now in parallel with the same total loss
of 20.3 m. Compute the total rate Q(m3/hr),
neglecting the minor losses.

L1= 100 m

L2= 150 m

L3= 80 m
Branching pipes, junctions
A branching network is illustrated below.
Determination of the discharges and flow directions:
All flows are considered as
negative towards the junction!
QA + QB + QC = 0
This implies that one or two of the
J flows must be outgoing from
junction!

The pressure must change through each pipe to provide the


same piezometric head at the junction. In other words, let the
HGL at the junction has the elevation:

 pJ 
hJ =  + z J 
 γ 
Solution Procedure
1.Guess a value for hJ.

2.Determine f for each pipe.

J
3.Solve the equations for VA,
VB & VC and hence for QA, QB
& QC.
LA VA2
hf A = fA = HA − HJ
DA 2 g 4.Iterate computations until
LB VB2 flow rate balance at the
hf B = fB = HB − HJ
DB 2 g junction QA+QB+QC=0

If QA+QB+QC >0, hJ is too


high, reduce hJ and vice versa.
Exp
Find the flow rates in each pipe, neglecting minor
losses (ν = 1.02 x 10-6 m2/s).

You might also like