branching pipes design

78
ÇUKUROVA UNIVERSITY ENGINEERING AND ARCHITECTURE FACUTY DEPARTMENT OF MECHANICAL ENGINEERING ENGINEERING PROJECT - I BRANCHING PIPES DESIGN BY 2008595015 CANER KİBAR ADVISOR

Upload: caner-kibar

Post on 20-Apr-2017

220 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Branching Pipes Design

ÇUKUROVA UNIVERSITY

ENGINEERING AND ARCHITECTURE FACUTY

DEPARTMENT OF MECHANICAL ENGINEERING

ENGINEERING PROJECT - I

BRANCHING PIPES DESIGN

BY

2008595015 CANER KİBAR

ADVISOR

ASSOC. PROF. AHMET PINARBAŞI

JANUARY 2012

ADANA

Page 2: Branching Pipes Design

ACKNOWLEDGEMENT

I would like to thank to my advisor Assoc. Prof. Ahmet PINARBAŞI for his guidance,

patience and supports during whole study process. Without his leadership, this book couldn’t

be written.

Also I want to thank to Prof. Dr. Beşir ŞAHİN and Prof. Dr. Hüseyin AKILLI for their

great expressions at Fluid Mechanics lectures. I used basic fluid mechanics principles many

times in this book and I owe this opportunity to them.

II

Page 3: Branching Pipes Design

CONTENTS

ACKNOWLEDGEMENT.......................................................................................................IILIST OF FIGURES................................................................................................................IVLIST OF TABLES...................................................................................................................VCHAPTER 1..............................................................................................................................61.INTRODUCTION..................................................................................................................6CHAPTER 2..............................................................................................................................72. FUNDAMENTAL PRINCIPLES........................................................................................7

2.1.Energy And Hydraulic Grade Lines.............................................................................72.2.Laminar or Turbulent Flow...........................................................................................92.3.Losses And Calculations...............................................................................................11

2.3.1.Head Loss................................................................................................................112.3.2. Minor Loss.............................................................................................................182.3.3. Other Local Losses................................................................................................29

CHAPTER 3............................................................................................................................323. THE ANALYSIS OF PIPING SYSTEM AND PUMP SELECTION...........................32

3.1. Pipe Design...................................................................................................................323.1.1. Pipe Materials........................................................................................................323.1.2. Pressure Class Guidelines....................................................................................323.1.3. Limiting Velocities................................................................................................33

3.2. Pipe Systems.................................................................................................................343.2.1. Single Pipes............................................................................................................343.2.2. Multiple Pipe Systems...........................................................................................35

3.3. Pump Characteristics and Selection...........................................................................393.3.1.Pump Characteristics............................................................................................393.3.2. Pump Selection......................................................................................................403.3.3. Revision of Energy Equation...............................................................................44

CHAPTER 4............................................................................................................................454. BRANCHING PIPES DESIGN.........................................................................................45

4.1. Problem Statement.......................................................................................................454.2. Calculation of Project..................................................................................................47

REFERANCES.......................................................................................................................53

III

Page 4: Branching Pipes Design

LIST OF FIGURES

Figure 2.1 – Representation of the energy line and hydraulic grade line. ................................. 8 Figure 2.2 - Experiment to illustrate type of flow. ..................................................................... 9 Figure 2.3 - Time dependence of fluid velocity at a point. ...................................................... 10 Figure 2.4 – The Moody Chart. ................................................................................................ 16 Figure 2.5 - Entrance flow conditions and loss coefficient ...................................................... 20 Figure 2.6 - Flow pattern and pressure distribution for a sharp-edged entrance. .................... 21 Figure 2.7 - Exit flow conditions and loss coefficient. ............................................................ 22 Figure 2.8 – Loss coefficient for a sudden contraction ............................................................ 22 Figure 2.9 – Loss coefficient for a sudden expansion .............................................................. 23 Figure 2.10 - Control volume used to calculate the loss coefficient for a sudden expansion. . 23 Figure 2.11 - Loss coefficient for a typical conical diffuser. ................................................... 25 Figure 2.12 - Character of the flow in a 900 bend and the associated loss coefficient ............. 26 Figure 2.13 - Character of the flow in a 900 mitered bend ....................................................... 27 Figure 2.14 – Boundary layer seperation ................................................................................. 29 Figure 2.15 – Local losses in pipe flow ................................................................................... 30 Figure 3.1 – Series and paralel pipe systems ........................................................................... 36 Figure 3.2 – Multiple pipe loop system. .................................................................................. 38 Figure 3.3 – Pump selection for a single pump ........................................................................ 41 Figure 3.4 – Selection of paralel pumps .................................................................................. 43 Figure 4.1 – Shematic Representation of Branching Pipes Water Supply System .................. 45 Figure 4.2 – Feasibility of Pump with Checking Power-Flow Rate Comparison .................... 52 Figure 4.3 - Feasibility of Pump with Checking Head Loss-Flow Rate Comparison ............. 52

IV

Page 5: Branching Pipes Design

LIST OF TABLES

Table 2.1 – Equivalent roughness for new pipes. .................................................................... 15 Table 2.2 - Loss coefficients for pipe components. ................................................................. 28 Table 2.3 - KL values for practical calculations ....................................................................... 30 Table 3.1 – Three most common types of problems ................................................................ 35 Table 4.1. – Necessary Informations of Branching Pipes System ........................................... 46 Table 4.2. – Flow Rate Distribution and Other Informations of Pipes .................................... 49

V

Page 6: Branching Pipes Design

CHAPTER 1

1.INTRODUCTION

At the start of this book “the branching” term should be well understood before investigate the

branching pipes system and due to achieve to aim of project. In a branching system a number

of pipes are connected to the main to form the topology of a tree. Assuming that the flow is

from the main into the smaller laterals it is possible to calculate the flow rate in any pipe as

the sum of the downstream consumptions or demands. If the laterals supply water to the main,

the same might be done. In either case by proceeding from the outermost branches toward the

main or “root of the tree” the flow rate can be calculated, and from the flow rate in each pipe

the head loss can be determined using the Darcy-Weisbach or Hazen-Williams equation. In

analyzing a pipe network containing a branching system, only the main is included with the

total flow rate specified by summing from the smaller pipes. Upon completing the analysis the

pressure head in the main will be known. By substracting individual head losses from this

known head, the heads (or pressures) at any point throughout the branching system can be

determined.

As they are mentioned above, before the analysis of design problem some basic principles

will be explained and examined at next chapters such as energy and hydraulic grade lines,

type of losses and their formulas, pump theory and characteristics, flow characteristics which

are foundations of the material in this book.

6

Page 7: Branching Pipes Design

CHAPTER 2

2. FUNDAMENTAL PRINCIPLES

2.1.Energy And Hydraulic Grade Lines

The Hydraulic Grade Line and Energy Line are simply graphical forms of the Bernoulli

Equation. The Bernoulli equation is actually an energy equation representing

the partitioning of energy for an inviscid, incompressible, steady flow. The sum of

the various energies of the fluid remains constant as the fluid flows from one section to

another.

A useful interpretation of the Bernoulli equation can be obtained through the use of

the concepts of the hydraulic grade line (HGL) and the energy line (EL). These ideas

represent a geometrical interpretation of a flow and can often be effectively used to better

grasp the fundamental processes involved.

The Bernoulli equation can be obtained by integrating “F=m.a” along a stream line and it is:

(2.1)

For steady, inviscid, incompressible flow the total energy remains constant along a

streamline. The concept of “head” was introduced by dividing each term in Eq. 2.1 by the

specific weight, “γ=ρ.g”, to give the Bernoulli equation in the following form:

(2.2)

Each of the terms in this equation has the units of length (feet or meters) and represents a

certain type of head. The Bernoulli equation states that the sum of the pressure head, the

velocity head, and the elevation head is constant along a streamline. This constant is called the

total head, H.

The energy line is a line that represents the total head available to the fluid. As shown

in Fig. 2.1, the elevation of the energy line can be obtained by measuring the stagnation

pressure with a Pitot tube. (A Pitot tube is the portion of a Pitot-static tube that measures the

stagnation pressure.) The stagnation point at the end of the Pitot tube provides

a measurement of the total head (or energy) of the flow. The static pressure tap connected

to the piezometer tube shown, on the other hand, measures the sum of the pressure

head and the elevation head, “p/γ + z”, This sum is often called the piezometric head. The

static pressure tap does not measure the velocity head.

7

Page 8: Branching Pipes Design

Figure 2.1 – Representation of the energy line and hydraulic grade line.

According to Eq. 2.2, the total head remains constant along the streamline. Thus, a pitot tube

at any other location in the flow will measure the same total head, as is shown in the figure.

The elevation head, velocity head, and pressure head may vary along the streamline, however.

The locus of elevations provided by a series of pitot tubes is termed the energy line,

EL. The level provided by a series of piezometer taps is termed the hydraulic grade line,

HGL. Under the assumptions of the Bernoulli equation, the energy line is horizontal. If the

fluid velocity changes along the streamline, the hydraulic grade line will not be horizontal.

If viscous effects are important, the total head does not remain constant due to a loss in energy

as the fluid flows along its streamline. This means that the energy line is no longer horizontal.

8

Page 9: Branching Pipes Design

2.2.Laminar or Turbulent Flow

Figure 2.2 - (a) Experiment to illustrate type of flow. (b) Typical dye streaks.

The flow of a fluid in a pipe may be laminar flow or it may be turbulent flow. Osborne

Reynolds (1842–1912), a British scientist and mathematician, was the first to distinguish the

difference between these two classifications of flow by using a simple apparatus as shown

in Fig. 8.3a. If water runs through a pipe of diameter D with an average velocity V, the

following characteristics are observed by injecting neutrally buoyant dye as shown. For

“small enough flowrates” the dye streak (a streakline) will remain as a well-defined line as it

flows along, with only slight blurring due to molecular diffusion of the dye into the

surrounding water. For a somewhat larger “intermediate flowrate” the dye streak fluctuates in

time and space, and intermittent bursts of irregular behavior appear along the streak. On the

other hand, for “large enough flowrates” the dye streak almost immediately becomes blurred

and spreads across the entire pipe in a random fashion. These three characteristics, denoted as

laminar, transitional, and turbulent flow, respectively, are illustrated in Fig. 2.2b.

The curves shown in Fig. 2.3 represent the x component of the velocity as a function

of time at a point A in the flow. The random fluctuations of the turbulent flow (with the

associated particle mixing) are what disperse the dye throughout the pipe and cause the

blurred appearance illustrated in Fig. 2.2b. For laminar flow in a pipe there is only one

component of velocity, V=uî. For turbulent flow the predominant component of velocity is

also along the pipe, but it is unsteady and accompanied by random components normal to the

9

Page 10: Branching Pipes Design

pipe axis, V=uî+vj+wk. Such motion in a typical flow occurs too fast for our eyes to follow.

Slow motion pictures of the flow can more clearly reveal the irregular, random, turbulent

nature of the flow.

Figure 2.3 - Time dependence of fluid velocity at a point.

Dimensional quantities should not be label as being “large” or “small,” such as “small enough

flowrates” in the preceding paragraphs. Rather, the appropriate dimensionless quantity should

be identified and the “small” or “large” character attached to it. A quantity is “large” or

“small” only relative to a reference quantity. The ratio of those quantities results in a

dimensionless quantity. For pipe flow the most important dimensionless parameter is the

Reynolds number, Re—the ratio of the inertia to viscous effects in the flow. Re= ρVD/ μ

where V is the average velocity in the pipe. That is, the flow in a pipe is laminar, transitional,

or turbulent provided the Reynolds number is “small enough,” “intermediate,” or “large

enough.” It is not only the fluid velocity that determines the character of the flow—its density,

viscosity, and the pipe size are of equal importance. These parameters combine to produce the

Reynolds number. The distinction between laminar and turbulent pipe flow and its

dependence on an appropriate dimensionless quantity was first pointed out by Osborne

Reynolds in 1883.

The Reynolds number ranges for which laminar, transitional, or turbulent pipe flows

are obtained cannot be precisely given. The actual transition from laminar to turbulent flow

may take place at various Reynolds numbers, depending on how much the flow is disturbed

by vibrations of the pipe, roughness of the entrance region, and the like. For general

engineering purposes (i.e. without undue precautions to eliminate such disturbances), the

following values are appropriate: The flow in a round pipe is laminar if the Reynolds number

10

Page 11: Branching Pipes Design

is less than approximately 2100. The flow in a round pipe is turbulent if the Reynolds number

is greater than approximately 4000. For Reynolds numbers between these two limits, the

flow may switch between laminar and turbulent conditions in an apparently random fashion

(transitional flow).

2.3.Losses And Calculations

2.3.1.Head Loss

The head loss in a pipe is a result of the viscous shear stress on the wall. Because of viscosity,

there is friction within the fluid as well as friction of the fluid against the piping or ducting

walls. This friction converts into heat some of the pressure energy of the flowing fluid and

raises the temperature of the fluid and piping. This phenomenon can be critical in the

operation of some quipment.

The energy equation for incompressible, steady flow between two locations can be written as:

(2.3)

Recall that the kinetic energy coefficients, α1 and α2, compensate for the fact that the velocity

profile across the pipe is not uniform. For uniform velocity profiles α=1, whereas for any

nonuniform profile, α >1.

The head loss term, hL, accounts for any energy loss associated with the flow. This loss is a

direct consequence of the viscous dissipation that occurs throughout the fluid in the pipe. For

the ideal (inviscid) cases α1 = α2 = 1, hL = 0 and the energy equation reduces to the familiar

bernoulli equation, Eq. 2.1.

Even though the velocity profile in viscous pipe flow is not uniform, for fully developed

flow it does not change from section (1) to section (2) so that α1 = α2. Thus, the kinetic

energy is the same at any section (α1.V12/2 = α2. V2

2/2) and the energy equation becomes:

(2.4)

Consider a volume element of fluid which is flowing in a circular pipe, with coordinate x in

the flow direction and r radially, lenght, L and diameter, D. With the definitions of wall shear

11

Page 12: Branching Pipes Design

stress (τw), net pressure force in element is (P1- P2= γ. hL), net shear force in the element is

(τw.π.D), force balance for equilibrium yields:

(2.5)

Substituting the equations 2.4 and 2.5, frictional head loss can be written as:

(2.6)

It is the shear stress at the wall (which is directly related to the viscosity and the shear stress

throughout the fluid) that is responsible for the head loss. A closer consideration of the

assumptions involved in the derivation of Eq. 2.6 will show that it is valid for both laminar

and turbulent flow.

As is discussed above, the pressure drop and head loss in a pipe are dependent on the wall

shear stres, τw, between the fluid and pipe surface. A fundamentaldifference between laminar

and turbulent flow is that the shear stress for turbulent flow is a function of the density of the

fluid, ρ. For laminar flow, the shear stress is independent of the density, leaving the viscosity,

μ, as the only important fluid property. Thus, the pressure drop, Δp, for steady,

incompressible turbulent flow in a horizontal round pipe of diameter D can be written in

functional form as Δp=f(V,D,L,ε,μ,ρ) where V is the average velocity, L is the pipe length,

and ε is a measure of the roughness of the pipe wall. It is clear that Δp should be a function of

V, D, and L. The dependence of Δp on the fluid properties μ and ρ is expected because of the

dependence of τ on these parameters.

Although the pressure drop for laminar pipe flow is found to be independent of the roughness

of the pipe, it is necessary to include this parameter when considering turbulent flow. For

turbulent flow there is a relatively thin viscous sublayer formed in the fluid near the pipe wall.

In many instances this layer is very thin; δs/D<<1, where δs is the sublayer thickness. If a

typical wall roughness element protrudes sufficiently far into (or even through) this layer, the

structure and properties of the viscous sublayer (along with Δp and τw) will be different than if

the wall were smooth. Thus, for turbulent flow the pressure drop is expected to be a function

of the wall roughness. For laminar flow there is no thin viscous layer thus, relatively small

roughness elements have completely negligible effects on laminar pipe flow. Of course, for

12

Page 13: Branching Pipes Design

pipes with very large wall roughness, (ε/D>0.1), the flowrate may be a function of the

roughness.

The list of parameters given in Δp=f(V,D,L,ε,μ,ρ) is apparently a complete one. experiments

have shown that other parameters (such as surface tension, vapor pressure, etc.) do not affect

the pressure drop for the conditions stated (steady, incompressible flow; round, horizontal

pipe). Since there are seven variables (k=7) which can be written in terms of the three

reference dimensions MLT (r=3), Δp=f(V,D,L,ε,μ,ρ) can be written in dimensionless form in

terms of k-r=4 dimensionless groups. One such representation is:

This result differs from that used for laminar flow in two ways. First, it have been chosen to

make the pressure dimensionless by dividing by the dynamic pressure, ρV2/2, rather than a

characteristic viscous shear stress, μV/D. This convention was chosen in recognition of the

fact that the shear stress for turbulent flow is normally dominated by τturb, which is a stronger

function of the density than it is of viscosity. Second, it have been introduced two additional

dimensionless parameters, the Reynolds number, Re= ρVD/ μ, and the relative roughness, ε/D

which are not present in the laminar formulation because the two parameters, ρ and ε are not

important in fully developed laminar pipe flow.

As was done for laminar flow, the functional representation can be simplified by imposing

the reasonable assumption that the pressure drop should be proportional to the pipe length.

(Such a step is not within the realm of dimensional analysis. It is merely a logical assumption

supported by experiments.) The only way that this can be true is if the L/D dependence is

factored as:

the quantity ΔpD/(LρV2/2) is termed the friction factor, f. Thus, for a horizontal pipe:

(2.7) where

13

Page 14: Branching Pipes Design

For laminar fully developed flow, the value of f is simply f=64/Re, independent of ε/D. For

turbulent flow, the functional dependence of the friction factor on the Reynolds number and

relative toughness, f=ϕ(Re, ε/D), is a rather complex one that cannot, as yet, be obtained from

a theoretical analysis. The results are obtained from an exhaustive set of experiments and

usually presented in terms of a curve-fitting formula or the equivalent graphical form.

The energy equation for steady incompressible flow is:

where hL is the head loss between sections (1) and (2). With the assumption of a constant

diameter ( D1=D2 so that V1=V2), horizontal (z1=z2) pipe with fully developed flow (α1= α2),

this becomes Δp=p1-p2=γ hL, which can be combined with Eq. 2.7 to give

(2.8)

Equation 2.8, called the Darcy-Weisbach equation, is valid for any fully developed, steady,

incompressible pipe flow-whether the pipe is horizontal or on a hill-. On the other hand, Eq.

2.7 is valid only for horizontal pipes. In general, with V1=V2 the energy equation gives

Part of the pressure change is due to the elevation change and part is due to the head loss

associated with frictional effects, which are given in terms of the friction factor, f.

It is sometimes useful to write the Darcy equation in terms of discharge Q, (using Q = AV)

or with a %1 error

It is not easy to determine the functional dependence of the friction factor on the Reynolds

number and relative roughness. Much of this information is a result of experiments conducted

by J. Nikuradse in 1933 and amplified by many others since then. One difficulty lies in the

determination of the roughness of the pipe. Nikuradse used artificially roughened pipes

produced by gluing sand grains of known size onto pipe walls to produce pipes with

sandpaper-type surfaces. The pressure drop needed to produce a desired flowrate was

14

Page 15: Branching Pipes Design

measured and the data were converted into the friction factor for the corresponding Reynolds

number and relative roughness. The tests were repeated numerous times for a wide range of

Re and ε/D to determine the f=ϕ(Re, ε/D) dependence.

In commercially available pipes the roughness is not as uniform and well defined as in the

artificially roughened pipes used by Nikuradse. However, it is possible to obtain a measure of

the effective relative roughness of typical pipes and thus to obtain the friction factor. Typical

roughness values for various pipe surfaces are given in Table 2.1. Figure 2.4 shows the

functional dependence of f on Re and ε/D and is called the Moody chart in honor of L. F.

Moody, who, along with C. F. Colebrook, correlated the original data of Nikuradse in terms

of the relative roughness of commercially available pipe materials. It should be noted that the

values of ε/D do not necessarily correspond to the actual values obtained by a microscopic

determination of the average height of the roughness of the surface. They do, however,

provide the correct correlation for f=ϕ(Re, ε/D)

Table 2.1 – Equivalent roughness for new pipes.

It is important to observe that the values of relative roughness given pertain to new, clean

pipes. After considerable use, most pipes (because of a buildup of corrosion or scale) may

have a relative roughness that is considerably larger (perhaps by an order of magnitude) than

that given. Very old pipes may have enough scale buildup to not only alter the value of ε but

also to change their effective diameter by a considerable amount.

The following characteristics are observed from the data of Fig. 2.4. For laminar flow,

f=64/Re, which is independent of relative roughness. For very large Reynolds numbers,

f=ϕ(Re, ε/D), which is independent of the Reynolds number. For such flows, commonly

15

Page 16: Branching Pipes Design

termed completely turbulent flow (or wholly turbulent flow), the laminar sublayer is so thin

(its thickness decreases with increasing Re) that the surface roughness completely dominates

the character of the flow near the wall. Hence, the pressure drop required is a result of an

inertia-dominated turbulent shear stress rather than the viscosity-dominated laminar shear

Figure 2.4 – The friction factor as a function of Reynolds number and relative roughness for

round pipes. – The Moody Chart.

16

Page 17: Branching Pipes Design

stress normally found in the viscous sublayer. For flows with moderate values of Re, the

friction factor is indeed dependent on both the Reynolds number and relative roughness

f=ϕ(Re, ε/D). The gap in the figure for which no values of f are given ( 2100<Re<4000 range)

is a result of the fact that the flow in this transition range may be laminar or turbulent (or an

unsteady mix of both) depending on the specific circumstances involved.

Note that even for smooth pipes (ε=0) the friction factor is not zero. That is, there is a head

loss in any pipe, no matter how smooth the surface is made. This is a result of the no-slip

boundary condition that requires any fluid to stick to any solid surface it flows over. There is

always some microscopic surface roughness that produces the no-slip behavior (and thus f 0)

on the molecular level, even when the roughness is considerably less than the viscous

sublayer thickness. Such pipes are called hydraulically smooth.

Various investigators have attempted to obtain an analytical expression for f=ϕ(Re, ε/D). Note

that the Moody chart covers an extremely wide range in flow parameters. The nonlaminar

region covers more than four orders of magnitude in Reynolds number – from Re=4.103 to

Re=108. Obviously, for a given pipe and fluid, typical values of the average velocity do not

cover this range. However, because of the large variety in pipes (D), fluids (ρ and μ) and

velocities (V), such a wide range in Re is needed to accommodate nearly all applications of

pipe flow. In many cases the particular pipe flow of interest is confined to a relatively small

region of the Moody chart, and simple semiempirical expressions can be developed for those

conditions. The Moody chart is universally valid for all steady, fully developed,

incompressible pipe flows. The following equation from Colebrook is valid for the entire

nonlaminar range of the Moody chart

(2.9)

In fact, the Moody chart is a graphical representation of this equation, which is an empirical

fit of the pipe flow pressure drop data. Equation 2.9 is called the Colebrook formula. The

turbulent portion of the Moody chart is represented by the Colebrook formula. A difficulty

with its use is that it is implicit in the dependence of f. That is, for given conditions (Re and

ε/D) it is not possible to solve for f without some sort of iterative scheme. With the use of

modern computers and calculators, such calculations are not difficult. A word of caution is in

order concerning the use of the Moody chart or the equivalent Colebrook formula. Because of

17

Page 18: Branching Pipes Design

various inherent inaccuracies involved (uncertainty in the relative roughness, uncertainty in

the experimental data used to produce the Moody chart, etc.), the use of several place

accuracy in pipe flow problems is usually not justified. As a rule of thumb, a 10% accuracy is

the best expected.

Note: The f value shown above is different to that used in American practice. Their

relationship is f American = 4f. Sometimes the f is replaced by the Greek letter λ, where

λ= f American = 4f. Consequently great care must be taken when choosing the value of f with

attention taken to the source of that value.

2.3.2. Minor Loss

Losses occur in straight pipes (major losses) and pipe system components (minor losses). the

head loss in long, straight sections of pipe can be calculated by use of the friction factor

obtained from either the Moody chart or the Colebrook equation. Most pipe systems,

however, consist of considerably more than straight pipes. These additional components

(valves, bends, tees, and the like) add to the overall head loss of the system. Such losses are

generally termed minor losses, with the apparent implication being that the majority of the

system loss is associated with the friction in the straight portions of the pipes, the major

losses. In many cases this is true. In other cases the minor losses are greater than the major

losses. In this section, how to determine the various minor losses that commonly occur in pipe

systems will be indicated.

The head loss associated with flow through a valve is a common minor loss. The purpose of a

valve is to provide a means to regulate the flowrate. This is accomplished by changing the

geometry of the system (i.e. closing or opening the valve alters the flow pattern

through the valve), which in turn alters the losses associated with the flow through the valve.

The flow resistance or head loss through the valve may be a significant portion of the

resistance in the system. In fact, with the valve closed, the resistance to the flow is infinite –

the fluid cannot flow. Such minor losses may be very important indeed. With the valve wide

open the extra resistance due to the presence of the valve may or may not be negligible.

It is not difficult to realize that a theoretical analysis to predict the details of such flows to

obtain the head loss is not, as yet, possible. Thus, the head loss information for essentially all

components is given in dimensionless form and based on experimental data. The most

common method used to determine these head losses or pressure drops is to specify the loss

coefficient, KL, which is defined as

18

Page 19: Branching Pipes Design

so that or

(2.10)

Losses due to pipe system components are given in terms of loss coefficients. The pressure

drop across a component that has a loss coefficient of KL=1 is equal to the dynamic pressure,

ρV2/2.

The actual value of is strongly dependent on the geometry of the component considered. It

may also be dependent on the fluid properties. It is, KL=ϕ(geometry, Re) where Re=ρVD/ μ is

the pipe Reynolds number. For many practical applications the Reynolds number is large

enough so that the flow through the component is dominated by inertia effects, with viscous

effects being of secondary importance. This is true because of the relatively large

accelerations and decelerations experienced by the fluid as it flows along a rather curved,

variable-area (perhaps even torturous) path through the component. In a flow that is

dominated by inertia effects rather than viscous effects, it is usually found that pressure drops

and head losses correlate directly with the dynamic pressure. This is the reason why the

friction factor for very large Reynolds number, fully developed pipe flow is independent of

the Reynolds number. The same condition is found to be true for flow through pipe

components. Thus, in most cases of practical interest the loss coefficients for components are

a function of geometry only, KL=ϕ(geometry, Re).

Minor losses are sometimes given in terms of an equivalent length, Leq. In this terminology,

the head loss through a component is given in terms of the equivalent length of pipe that

would produce the same head loss as the component. That is,

or

where D and f are based on the pipe containing the component. The head loss of the pipe

system is the same as that produced in a straight pipe whose length is equal to the pipes of the

original system plus the sum of the additional equivalent lengths of all of the components of

19

Page 20: Branching Pipes Design

the system. Most pipe flow analyses, including those in this book, use the loss coefficient

method rather than the equivalent length method to determine the minor losses.

Many pipe systems contain various transition sections in which the pipe diameter changes

from one size to another. Such changes may occur abruptly or rather smoothly through some

type of area change section. Any change in flow area contributes losses that are not accounted

for in the fully developed head loss calculation (the friction factor). The extreme cases involve

flow into a pipe from a reservoir (an entrance) or out of a pipe into a reservoir (an exit).

Figure 2.5 - Entrance flow conditions and loss coefficient (a) Reentrant, KL =0.8, (b) sharp-

edged, KL =0.5, (c) slightly rounded, KL =0.2 (d) well-rounded, KL =0.04

A fluid may flow from a reservoir into a pipe through any number of different shaped

entrance regions as are sketched in Fig. 2.5. Each geometry has an associated loss coefficient.

A typical flow pattern for flow entering a pipe through a square-edged entrance is sketched in

Fig. 2.6. A vena contracta region may result because the fluid cannot turn a sharp right-angle

corner. The flow is said to separate from the sharp corner. The maximum velocity at section

(2) is greater than that in the pipe at section (3), and the pressure there is lower. If this high-

speed fluid could slow down efficiently, the kinetic energy could be converted into pressure

(the Bernoulli effect), and the ideal pressure distribution indicated in Fig. 2.6 would result.

The head loss for the entrance would be essentially zero.

20

Page 21: Branching Pipes Design

Figure 2.6 - Flow pattern and pressure distribution for a sharp-edged entrance.

Minor head losses are often a result of the dissipation of kinetic energy. Although a fluid may

be accelerated very efficiently, it is very difficult to slow down (decelerate) a fluid efficiently.

Thus, the extra kinetic energy of the fluid at section (2) is partially lost because of viscous

dissipation, so that the pressure does not return to the ideal value. An entrance head loss

1pressure drop2 is produced as is indicated in Fig. 2.6. The majority of this loss is due to

inertia effects that are eventually dissipated by the shear stresses within the fluid. Only a small

portion of the loss is due to the wall shear stress within the entrance region. The net effect is

that the loss coefficient for a square-edged entrance is approximately KL =0.50. One-half of a

velocity head is lost as the fluid enters the pipe. If the pipe protrudes into the tank (a reentrant

entrance) as is shown in Fig. 2.5a, the losses are even greater.

An obvious way to reduce the entrance loss is to round the entrance region as is shown in Fig.

2.5c, thereby reducing or eliminating the vena contracta effect. A significant reduction in KL

can be obtained with only slight rounding.

21

Page 22: Branching Pipes Design

Figure 2.7 - Exit flow conditions and loss coefficient. (a) Reentrant, KL =1.0 (b) sharp-edged,

KL=1.0, (c) slightly rounded, KL=1.0, (d) well-rounded, KL=1.0.

A head loss (the exit loss) is also produced when a fluid flows from a pipe into a tank as is

shown in Fig. 2.7. In these cases the entire kinetic energy of the exiting fluid (velocity V1) is

dissipated through viscous effects as the stream of fluid mixes with the fluid in the tank and

eventually comes to rest( V2=0). The exit loss from points (1) and (2) is therefore equivalent

to one velocity head, or KL=1.

Figure 2.8 – Loss coefficient for a sudden contraction

22

Page 23: Branching Pipes Design

Figure 2.9 – Loss coefficient for a sudden expansion

Losses also occur because of a change in pipe diameter as is shown in Figs. 2.8 and 2.9. The

sharp-edged entrance and exit flows discussed in the previous paragraphs are limiting cases of

this type of flow with either A1/A2= , or A1/A2=0, respectively. The loss coefficient for a

sudden contraction, KL=hL/(V2/2g), is a function of the area ratio, A1/A2 , as is shown in Fig.

2.8. The value of KL changes gradually from one extreme of a sharp-edged entrance (A1/A2=0

with KL=0.50) to the other extreme of no area change (A1/A2=1 with KL=0).

In many ways, the flow in a sudden expansion is similar to exit flow. As is indicated in Fig.

2.10, the fluid leaves the smaller pipe and initially forms a jet-type structure as it enters the

larger pipe. Within a few diameters downstream of the expansion, the jet becomes dispersed

across the pipe, and fully developed flow becomes established again. In this process [between

sections (2) and (3)] a portion of the kinetic energy of the fluid is dissipated as a result of

viscous effects. A square-edged exit is the limiting case with A1/A2=0.

Figure 2.10 - Control volume used to calculate the loss coefficient for a sudden expansion.

23

Page 24: Branching Pipes Design

A sudden expansion is one of the few components (perhaps the only one) for which the loss

coefficient can be obtained by means of a simple analysis. To do this the continuity and

momentum equations for the control volume shown in Fig. 2.10 are considered and the energy

equation applied between (2) and (3). It’s assumed that the flow is uniform at sections (1), (2),

and (3) and the pressure is constant across the left-hand side of the control volume

(pa=pb=pc=p1). The resulting three governing equations (mass, momentum, and energy) are

and

These can be rearranged to give the loss coefficient, KL=hL/(V2/2g), as

This result, plotted in Fig. 2.9, is in good agreement with experimental data. As with so many

minor loss situations, it is not the viscous effects directly (i.e. the wall shear stres) that cause

the loss. Rather, it is the dissipation of kinetic energy (another type of viscous effect) as the

fluid decelerates inefficiently.

The losses may be quite different if the contraction or expansion is gradual. Typical results for

a conical diffuser with a given area ratio, A1/A2, are shown in Fig. 2.11. (A diffuser is a

device shaped to decelerate a fluid.) Clearly the included angle of the diffuser, θ, is a very

important parameter. For very small angles, the diffuser is excessively long and most of the

head loss is due to the wall shear stress as in fully developed flow. For moderate or large

angles, the flow separates from the walls and the losses are due mainly to a dissipation of the

kinetic energy of the jet leaving the smaller diameter pipe. In fact, for moderate or large

values of θ (i.e. θ>350 for the case shown in Fig. 2.11), the conical diffuser is, perhaps

unexpectedly, less efficient than a sharp-edged expansion which has KL=1. There is an

optimum angle (θ 80 for the case illustrated) for which the loss coefficient is a minimum. The

24

Page 25: Branching Pipes Design

relatively small value of θ for the minimum KL results in a long diffuser and is an indication

of the fact that it is difficult to efficiently decelerate a fluid.

It must be noted that the conditions indicated in Fig. 8.29 represent typical results only. Flow

through a diffuser is very complicated and may be strongly dependent on the area ratio A1/A2,

specific details of the geometry, and the Reynolds number. The data are often presented in

terms of a pressure recovery coefficient, Cp=(p2-p1)/ (ρV21/2), which is the ratio of the static

pressure rise across the diffuser to the inlet dynamic pressure.

Figure 2.11 - Loss coefficient for a typical conical diffuser.

Flow in a conical contraction (a nozzle; reverse the flow direction shown in Fig. 2.11) is less

complex than that in a conical expansion. Typical loss coefficients based on the downstream

(high-speed) velocity can be quite small, ranging from KL=0.02 for θ=300, to KL=0.07 for

θ=600 for example. It is relatively easy to accelerate a fluid efficiently.

Bends in pipes produce a greater head loss than if the pipe were straight. The losses are due to

the separated region of flow near the inside of the bend (especially if the bend is sharp) and

the swirling secondary flow that occurs because of the imbalance of centripetal forces as a

result of the curvature of the pipe centerline. These effects and the associated values of KL for

large Reynolds number flows through a 900 bend are shown in Fig. 2.12.

25

Page 26: Branching Pipes Design

Figure 2.12 - Character of the flow in a 900 bend and the associated loss coefficient

For situations in which space is limited, a flow direction change is often accomplished by use

of miter bends, as is shown in Fig. 2.13, rather than smooth bends. The considerable losses in

such bends can be reduced by the use of carefully designed guide vanes that help direct the

flow with less unwanted swirl and disturbances.

Another important category of pipe system components is that of commercially available pipe

fittings such as elbows, tees, reducers, valves, and filters. The values of KL for such

components depend strongly on the shape of the component and only very weakly on the

Reynolds number for typical large Re flows. Thus, the loss coefficient for a 900 elbow

depends on whether the pipe joints are threaded or flanged but is, within the accuracy of the

data, fairly independent of the pipe diameter, flow rate, or fluid properties (the Reynolds

number effect). Typical values of KL for such components are given in Table 2.2. These

typical components are designed more for ease of manufacturing and costs than for reduction

of the head losses that they produce. The flowrate from a faucet in a typical house is sufficient

whether the value of for an elbow is the typical KL=1.5, or it is reduced to KL=0.2 by use of a

more expensive long-radius, gradual bend (Fig. 2.12).

26

Page 27: Branching Pipes Design

Figure 2.13 - Character of the flow in a 900 mitered bend and the associated loss coefficient:

(a) without guide vanes, (b) with guide vanes

A valve is a variable resistance element in a pipe circuit. Valves control the flowrate by

providing a means to adjust the overall system loss coefficient to the desired value. When the

valve is closed, the value of KL is infinite and no fluid flows. Opening of the valve reduces

producing the desired flowrate. Loss coefficients for typical valves are given in Table 2.2. As

with many system components, the head loss in valves is mainly a result of the dissipation of

kinetic energy of a high-speed portion of the flow.

27

Page 28: Branching Pipes Design

Table 2.2 - Loss coefficients for pipe components.

28

Page 29: Branching Pipes Design

2.3.3. Other Local Losses

Large losses in energy usually occur only where flow expands. The mechanism at work in

these situations is that as velocity decreases (by continuity) so pressure must increase (by

Bernoulli). When the pressure increases in the direction of fluid outside the boundary layer

has enough momentum to overcome this pressure that is trying to push it backwards. The

fluid within the boundary layer has so little momentum that it will very quickly be brought to

rest, and possibly reversed in direction. If this reversal occurs it lifts the boundary layer away

from the surface as shown in Figure 8. This phenomenon is known as boundary layer

separation.

Figure 2.14 – Boundary layer seperation

At the edge of the separated boundary layer, where the velocities change direction, a line of

vortices occur (known as a vortex sheet). This happens because fluid to either side is moving

in the opposite direction. This boundary layer separation and increase in the turbulence

because of the vortices results in very large energy losses in the flow. These separating/

divergent flows are inherently unstable and far more energy is lost than in parallel or

convergent flow.

Some common situation where significant head losses occur in pipe are shown in Fig. 2.15.

29

Page 30: Branching Pipes Design

Figure 2.15 – Local losses in pipe flow

The values of KL for these common situations are shown in Table 2.3. It gives value that are

used in practice.

Table 2.3 - KL values for practical calculations

30

Page 31: Branching Pipes Design

The summation of all friction and local losses in a pipe system can be expressed as:

(2.11)

(2.12)

(2.13)

It is important to use the correct pipe diameter for each pipe section and local loss. In the past

some have expressed the local losses as an equivalent pipe length: L/d = KL /f. It simply

represents the length of pipe that produces the same head loss as the local or minor loss. This

is a simple, but not a completely accurate method of including local losses. The problem with

this approach is that since the friction coefficient varies from pipe to pipe, the equivalent

length will not have a unique value. When local losses are truly minor, this problem becomes

academic because the error only influences losses which make up a small percentage of the

total. For cases where accurate evaluation of all losses is important, it is recommended that

the minor loss coefficients KL be used rather than an equivalent length.

The challenging part of making minor loss calculations is obtaining reliable values of KL. The

final results cannot be any more accurate than the input data. If the pipe is long, the friction

losses may be large compared with the minor losses and approximate values of KL will be

sufficient. However, for short systems with many pipe fittings, the local losses can represent a

significant portion of the total system losses, and they should be accurately determined.

Numerous factors influence KL. For example, for elbows, KL is influenced by the shape of the

conduit (rectangular vs. circular), by the radius of the bend, the bend angle, the Reynolds

number, and the length of the outlet pipe. For dividing or combining tees or Y-branches, the

percent division of ßow and the change in pipe diameter must also be included when

estimating KL. One factor which is important for systems where local losses are significant is

the interaction between components placed close together. Depending on the type, orientation,

and spacing of the components, the total loss coefficient may be greater or less than the

simple sum of the individual KL values.

31

Page 32: Branching Pipes Design

CHAPTER 3

3. THE ANALYSIS OF PIPING SYSTEM AND PUMP SELECTION

3.1. Pipe Design

3.1.1. Pipe Materials

Materials commonly used for pressure pipe transporting liquids are ductile iron, concrete,

steel, fiberglass, PVC, and polyolefin. Specifications have been developed by national

committees for each of these pipe materials. The specifications discuss external loads, internal

design pressure, available sizes, quality of materials, installation practices, and information

regarding linings. Standards are available from the following organizations:

American Water Works Association (AWWA)

American Society for Testing and Materials (ASTM)

American National Standards Institute (ANSI)

Federal Specifications (FED)

Plastic Pipe Institute (PPI)

Turkish Institute of Standarts (TSE)

In addition, manuals and other standards have been published by various manufacturers and

manufacturer‘s associations. All of these specifications and standards should be used to guide

the selection of pipe material. ANSI contains a description of each of these pipe materials and

a list of the specifications for the various organizations which apply to each material. It also

discusses the various pipe-lining materials available for corrosion protection.

For air- and low-pressure liquid applications one can use unreinforced concrete, corrugated

steel, smooth sheet metal, spiral rib (sheet metal), and HDPE (high-density polyethylene)

pipe. The choice of a material for a given application depends on pipe size, pressure

requirements, resistance to collapse from internal vacuums, external loads, resistance to

internal and external corrosion, ease of handling and installing, useful life, and economics.

3.1.2. Pressure Class Guidelines

Procedures for selecting the pressure class of pipe vary with the type of pipe material.

Guidelines for different types of materials are available from AWWA, ASTM, ANSI, FED,

PPI, TSE and from the pipe manufacturers. These specifications should be obtained and

studied for the pipe materials being considered.

32

Page 33: Branching Pipes Design

The primary factors governing the selection of a pipe pressure class are (1) the maximum

steady state operating pressure, (2) surge and transient pressures, (3) external earth loads and

live loads, (4) variation of pipe properties with temperature or long-time loading effects, and

(5) damage that could result from handling, shipping, and installing or reduction in strength

due to chemical attack or other aging factors. The influence of the first three items can be

quantified, but the last two are very subjective and are generally accounted for with a safety

factor which is the ratio of the burst pressure to the rated pressure. There is no standard

procedure on how large the safety factor should be or on how the safety factor should be

applied. Some may feel that it is large enough to account for all of the uncertainties. Past

failures of pipelines designed using this assumption prove that it is not always a reliable

approach. The procedure recommended by the author is to select a pipe pressure class based

on the internal design pressure (IDP) defined as

in which Pmax is the maximum steady state operating pressure, Ps is the surge or water hammer

pressure, and SF is the safety factor applied to take care of the unknowns (items 3 to 5) just

enumerated. A safety factor between 3 and 4 is typical.

The maximum steady state operating pressure (Pmax) in a gravity flow system is usually the

difference between the maximum reservoir elevation and the lowest elevation of the pipe. For

a pumped system it is usually the pump shutoff head calculated based on the lowest elevation

of the pipe.

Surge and transient pressures depend on the specific pipe system design and operation.

Accurately determining Ps requires analyzing the system using modern computer techniques.

Selection of wall thickness for larger pipes is often more dependent on collapse pressure and

handling loads than it is on burst pressure. A thin-wall, large-diameter pipe may be adequate

for resisting relatively high internal pressures but may collapse under negative internal

pressure or, if the pipe is buried, the soil and groundwater pressure plus live loads may be

sufficient to cause collapse even if the pressure inside the pipe is positive.

3.1.3. Limiting Velocities

There are concerns about upper and lower velocity limits. If the velocity is too low, problems

may develop due to settling of suspended solids and air being trapped at high points and along

33

Page 34: Branching Pipes Design

the crown of the pipe. The safe lower velocity limit to avoid collecting air and sediment

depends on the amount and type of sediment and on the pipe diameter and pipe profile.

Velocities greater than about 1 m/sec are usually sufficient to move trapped air to air release

valves and keep the sediment in suspension. Problems associated with high velocities are (1)

erosion of the pipe wall or liner (especially if coarse suspended sediment is present), (2)

cavitation at control valves and other restrictions, (3) increased pumping costs, (4) removal of

air at air release valves, (5) increased operator size and concern about valve shaft failures due

to excessive flow torques, and (6) an increased risk of hydraulic transients. Each of these

should be considered before making the final pipe diameter selection. A typical upper velocity

for many applications if 6 m/sec. However, with proper pipe design and analysis (of the

preceding six conditions), plus proper valve selection, much higher velocities can be tolerated.

A typical upper velocity limit for standard pipes and valves is about 6 m/sec. However, with

proper design and analysis, much higher velocities can be tolerated.

3.2. Pipe Systems

Pipe systems may contain a single pipe with components or multiple interconnected pipes.

3.2.1. Single Pipes

The nature of the solution process for pipe flow problems can depend strongly on which of

the various parameters are independent parameters (“given”) and which is the dependent

parameter (“determine”). The three most common types of problems are shown in Table 3.1

in terms of the parameters involved. It is assumed that the pipe system is defined in terms of

the length of pipe sections used and the number of elbows, bends, and valves needed to

convey the fluid between the desired locations. In all instances it is assumed the fluid

properties are given.

In a Type I problem the desired flowrate or average velocity is specified and the necessary

pressure difference or head loss is determined.

In a Type II problem the applied driving pressure (or, alternatively, the head loss) is specified

and the flowrate is determined.

In a Type III problem the pressure drop and the flowrate is specified and the diameter of the

pipe needed is determined.

34

Page 35: Branching Pipes Design

Table 3.1 – Three most common types of problems

Pipe flow problems in which it is desired to determine the flowrate for a given set of

conditions (Type II problems) often require trial-and-error solution techniques. This is

because it is necessary to know the value of the friction factor to carry out the calculations,

but the friction factor is a function of the unknown velocity (flowrate) in terms of the

Reynolds number.

In pipe flow problems for which the diameter is the unknown (Type III), an iterative

technique is required. This is, again, because the friction factor is a function of the diameter -

through both the Reynolds number and the relative roughness. Thus, neither 4ρQ/πμD nor ε/D

are known unless D is known.

3.2.2. Multiple Pipe Systems

In many pipe systems there is more than one pipe involved. The complex system of tubes in

our lungs or the maze of pipes in a city’s water distribution system are typical of such

systems. The governing mechanisms for the flow in multiple pipe systems are the same as for

the single pipe systems discussed in this chapter. However, because of the numerous

unknowns involved, additional complexities may arise in solving for the flow in multiple pipe

systems. Some of these complexities are discussed under this title.

The simplest multiple pipe systems can be classified into series or parallel flows, as are shown

in Fig. 8.35. The nomenclature is similar to that used in electrical circuits. Indeed, an analogy

35

Page 36: Branching Pipes Design

between fluid and electrical circuits is often made as follows. In a simple electrical circuit,

there is a balance between the voltage (V), current (i), and resistance (R) as given by Ohm’s

law: V=i.R. In a fluid circuit there is a balance between the pressure drop (Δp), the flowrate or

velocity (Q or V), and the flow resistance as given in terms of the friction factor and minor

loss coefficients (f and KL). For simple flow [Δp=f.(L/D).(ρV2/2)], it follows that Δp=Q2.R,

where R, a measure of the resistance to flow, is proportional to f.

Figure 3.1 – (a) Series and (b) paralel pipe systems

The main differences between the solution methods used to solve electrical circuit problems

and those for fluid circuit problems lie in the fact that Ohm’s law is a linear equation

(doubling the voltage doubles the current), while the fluid equations are generally nonlinear

(doubling the pressure drop does not double the flowrate unless the flow is laminar). Thus,

although some of the standard electrical engineering methods can be carried over to help

solve fluid mechanics problems, others cannot.

One of the simplest multiple pipe systems is that containing pipes in series, as is shown in

Fig. 3.1a. Every fluid particle that passes through the system passes through each of the pipes.

Thus, the flowrate (but not the velocity) is the same in each pipe, and the head loss from point

A to point B is the sum of the head losses in each of the pipes. The governing equations can

be written as follows:

36

Page 37: Branching Pipes Design

and

where the subscripts refer to each of the pipes. In general, the friction factors will be different

for each pipe because the Reynolds numbers (Rei=ρViDi/μ) and the relative roughnesses

(εi/Di) will be different. If the flowrate is given, it is a straightforward calculation to determine

the head loss or pressure drop (Type I problem). If the pressure drop is given and the flowrate

is to be calculated (Type II problem), an iteration scheme is needed. In this situation none of

the friction factors, fi, are known, so the calculations may involve more trial-and-error

attempts than for corresponding single pipe systems. The same is true for problems in which

the pipe diameter (or diameters) is to be determined (Type III problems).

Another common multiple pipe system contains pipes in parallel, as is shown in Fig. 3.1b. In

this system a fluid particle traveling from A to B may take any of the paths available, with the

total flowrate equal to the sum of the flowrates in each pipe. However, by writing the energy

equation between points A and B it is found that the head loss experienced by any fluid

particle traveling between these locations is the same, independent of the path taken. Thus, the

governing equations for parallel pipes are

and

Again, the method of solution of these equations depends on what information is given and

what is to be calculated.

Another type of multiple pipe system called a loop is shown in Fig. 3.2. In this case the

flowrate through pipe (1) equals the sum of the flowrates through pipes (2) and (3), or

Q1=Q2+Q3. As can be seen by writing the energy equation between the surfaces of each

reservoir, the head loss for pipe (2) must equal that for pipe (3), even though the pipe sizes

and flowrates may be different for each. That is,

for a fluid particle traveling through pipes (1) and (2), while

for fluid that travels through pipes (1) and (3). These can be combined to give hL2=hL3.

37

Page 38: Branching Pipes Design

Figure 3.2 – Multiple pipe loop system.

This is a statement of the fact that fluid particles that travel through pipe (2) and particles that

travel through pipe (3) all originate from common conditions at the junction (or node, N) of

the pipes and all end up at the same final conditions.

These informations have been introduced for a purpose of a pre-study to branching pipe

subject. Branching pipe will be explained at Chapter 4.

The analysis of piping networks, no matter how complex they are, is based on two simple

principles:

1- Conservation of mass throughout the system must be satisfied. This is done by

requiring the total flow into a junction to be equal to the total flow out of the juction

for all junctions in the system. Also, the flow rate must remain constant in pipes

connected series regardless of the changes in diameters.

2- Pressure drop (and thus head loss) between two junctions must be the same for all

paths between the two juctions. This is because pressure is a point of function and it

cannot have two values at specified point. In practice this rule is used by requiring that

the algebraic sum of head losses in a loop (for all loops) be equal to zero. (A head loss

is taken to be positive for flow in the clockwise direction and negative for flow in

counterclockwise direction.)

38

Page 39: Branching Pipes Design

3.3. Pump Characteristics and Selection

3.3.1.Pump Characteristics

The actual head rise, ha, gained by fluid flowing through a pump can be determined with an

experimental arrangement with using the energy equation with hp=hs-hL where hs is the shaft

work head and is identical to and hL is the pump head loss.

(3.1)

with sections (1) and (2) at the pump inlet and exit, respectively. The head, is the same as

used with the energy equation, where hp is interpreted to be the net head rise actually gained

by the fluid flowing through the pump, i.e. hp= hs-hL. Typically, the differences in elevations

and velocities are small so that:

(3.2)

The power, W, gained by the fluid is given by the equation

(3.3)

and this quantity, expressed in terms of horsepower is traditionally called the water

horsepower. Thus;

(3.4)

Note that if the pumped fluid is not water, the γ appearing in Eq. 3.4 must be the specific

weight of the fluid moving through the pump.

In addition to the head or power added to the fluid, the overall efficiency, η, is of interest,

where

(3.5)

39

Page 40: Branching Pipes Design

The overall pump efficiency is affected by the hydraulic losses in the pump, as previously

discussed, and in addition, by the mechanical losses in the bearings and seals. There may also

be some power loss due to leakage of the fluid between the back surface of the impeller hub

plate and the casing, or through other pump components. This leakage contribution to the

overall efficiency is called the volumetric loss. Thus, the overall efficiency arises from three

sources, the hydraulic efficiency, ηh, the mechanical efficiency, ηm, and the volumetric

efficiency, ηv, so that η= ηh. ηm. ηv

3.3.2. Pump Selection

Optimizing the life of a water supply system requires proper selection, operation, and

maintenance of the pumps. During the selection process, the designer must be concerned

about matching the pump performance to the system requirements and must anticipate

problems that will be encountered when the pumps are started or stopped and when the pipe is

filled and drained. The design should also consider the effect of variations in flow

requirements, and also anticipate problems that will be encountered due to increased future

demands and details of the installation of the pumps.

Selecting a pump for a particular service requires matching the system requirements to the

capabilities of the pump. The process consists of developing a system equation by applying

the energy equation to evaluate the pumping head required to overcome the elevation

difference, friction, and minor losses. For a pump supplying water between two reservoirs, the

pump head required to produce a given discharge can be expressed as

or

in which the constant C is defined by Eq. 2.13. Figure 3.3 shows a system curve for a pipe

having an 82-meter elevation lift and moderate friction losses. If the elevation of either

reservoir is a variable, then there is not a single curve but a family of curves corresponding to

differential reservoir elevations.

40

Page 41: Branching Pipes Design

Figure 3.3 – Pump selection for a single pump

The three pump curves shown in Figure 3.3 represent different impeller diameters. The

intersections of the system curve with the pump curves identify the ßow that each pump will

supply if installed in that system. For this example both A and B pumps would be a good

choice because they both operate at or near their best efficiency range. Figure 3.3 shows the

head and flow that the B pump will produce when operating in that system are 97 m and 450

L/m. The net positive suction head (NPSH) and brake horsepower (bhp) are obtained as

shown in the figure.

The selection process is more complex when the system demand varies, either due to

variations in the water surface elevation or to changing flow requirements. If the system must

operate over a range of reservoir elevations, the pump should be selected so that the system

curve, based on the mean (or the most frequently encountered) water level, intersects the

pump curve near the midpoint of the best efficiency range. If the water level variation is not

too great, the pump may not be able to operate efficiently over the complete flow range.

The problem of pump selection also becomes more difficult when planning for future

demands or if the pumps are required to supply a varying flow. If the flow range is large,

multiple pumps or a variablespeed drive may be needed. Recent developments in variable-

41

Page 42: Branching Pipes Design

frequency drives for pumps make them a viable alternative for systems with varying ßows.

Selection of multiple pumps and the decision about installing them in parallel or in series

depend on the amount of friction in the system. Parallel installations are most effective for

low-friction systems. Series pumps work best in high-friction systems. For parallel pump

operation the combined two pump curve is constructed by adding the flow of each pump.

Such a curve is shown in Figure 3.4 (labeled 2 pumps). The intersection of the two-pump

curve with the system curve identifies the combined flow for the two pumps. The pump

efficiency for each pump is determined by projecting horizontally to the left to intersect the

single-pump curve. For this example, a C pump, when operating by itself, will be have an

efficiency of 83%. With two pumps operating, the efficiency of each will be about 72%. For

the two pumps to operate in the most efficient way, the selection should be made so the

system curve intersects the single-pump curve to the right of its best efficiency point.

Starting a pump with the pipeline empty will result in filling at a very rapid rate because

initially there is little friction to build backpressure. As a result, the pump will operate at a

ßow well above the design flow. This may cause the pump to cavitate, but the more serious

problem is the possibility of high pressures generated by the rapid filling of the pipe.

Provisions should be made to control the rate of filling to a safe rate. Start-up transients are

often controlled by starting the pump against a partially open discharge valve located near the

pump and using a bypass line around the pump. This allows the system to be filled slowly and

safely. If the pipe remains full and no air is trapped, after the initial filling, subsequent start-up

of the pumps generally does not create any serious problem. Adequate air release valves

should be installed to release the air under low pressure.

42

Page 43: Branching Pipes Design

Figure 3.4 – Selection of paralel pumps

For some systems, stopping the pump, either intentionally or accidentally, can generate high

pressures that can damage the pipe and controls. If the design process does not consider these

potential problems, the system may not function trouble free. Downtime and maintenance

costs may be high. Not all systems will experience start-up and shutdown problems, but the

design should at least consider the possibility. The problem is more severe for pipelines that

have a large elevation change and multiple high points. The magnitude of the transient is

related to the length and profile of the pipeline, the pump characteristics, the magnitude of the

elevation change, and the type of check valve used. The downsurge caused by stopping the

pump can cause column separation and high pressures due to flow reversals and closure of

the check valves. Surge-protection equipment can be added to such systems to prevent

damage and excessive maintenance.

Selection of the proper pump will improve reliability, extend the economic life of the system,

and reduce maintenance. Failure to complete a transient analysis and include the required

controls will have the opposite effect. A system is only as good as it is designed to be.

43

Page 44: Branching Pipes Design

3.3.3. Revision of Energy Equation

When a piping system involves a pump, the steady-flow energy equation on a unit mass basic

can be expressed as:

where =Wpump/g is the useful pump head delivered to the fluid and hL is the total head

loss in piping (including minor losses if they are significant) between points (1) and (2). The

pump head is zero if the piping system does not involve a pump.

Many practical piping systems involve a pump to move a fluid from one reservoir to another.

Taking points (1) and (2) to be free surfaces of the reservoirs, the energy equation in this case

reduces for useful pump head:

since the velocities at free surfaces are negligible and the pressures are atmospheric pressure.

Therefore, the useful pump head is equal to the elevation difference between the two

reservoirs plus the head loss. If the head loss is negligible compared to , the useful

pump head is simply equal to elevation difference between two reservoirs.

44

Page 45: Branching Pipes Design

CHAPTER 4

4. BRANCHING PIPES DESIGN

4.1. Problem Statement

Figure 4.1. – Shematic Representation of Branching Pipes Water Supply System

45

Page 46: Branching Pipes Design

As it shown in Figure 4.1 the basic structure of four branched system is being considered at

this chapter. This system can be a pilot project to industry applications such as water demand

of workers which are working on different elevations at big construction fields, AC plants and

cooling towers, river water pumping, agricultural farming etc. In objective case, they have the

same principle and similar aim which is selection of suitable pump.

In this system, there is water usage at each tank which is shown in informations as outer

volume flow rate.

When looked to the catalogs for select a suitable pump, it is clearly seen that two distinct

property is needed which are required volume flow rate and head loss of pump. Before

proceeding to calculation subject of these two properties, necessary informations and

explanations of abbreviations are given below.

WSElnormal : Normal Water Surface Elevation

WSElmin : Minimum Water Surface Elevation

Qout : Outer Volume Flow Rate

QP : Volume Flow Rate of Pump

L : Length of Pipe

D : Diameter of Pipe

ε : Roughness of Pipe

A : Cross-Section Area

f : Friction Factor

V : Volume

Tank A Tank B Tank C Tank D

WSElnormal 32 m 28 m 24 m 18 m

WSElmin 22 m 20 m 16 m 12 m

Diameter (d) 16 m 10 m 8 m 10 m

Qout 0,08 m3/s 0,06 m3/s 0.04 m3/s 0,06 m3/s

Table 4.1. – Necessary Informations of Branching Pipes System

46

Page 47: Branching Pipes Design

4.2. Calculation of Project

First of all it is known that QP and hp must be determined for selection of pump and so, as a

start, QP can be calculated as:

QP = QA + QB + QC

Determination of volume flow rate of Tank A:

Qout,A = 0,08 m3/s = 288 m3/h

Volume of water which is out of tank after eight hour:

Vout,A =

Vout,A = 2304 m3

Volume of water between normal elevation and minimum elevation:

VA =

VA =

VA = 2010,61 m3

Amount of water requirement from pump to Tank A during eight hour:

Vreq,A = Vout,A – VA

Vreq,A = 2304 – 2010,61 m3

Vreq,A = 293,39 m3

QA = Vreq,A / 8 hr

QA = 0,010 m3/s = 36,673 m3/hr

Determination of volume flow rate of Tank B:

Qout,B = 0,06 m3/s = 216 m3/h

Volume of water which is out of tank after eight hour:

Vout,B =

Vout,B = 1728 m3

Volume of water between normal elevation and minimum elevation:

VB =

47

Page 48: Branching Pipes Design

VB =

VB = 628,31 m3

Amount of water requirement from pump to Tank B during eight hour:

Vreq,B = Vout,B – VB

Vreq,B = 1728 – 628,31 m3

Vreq,B = 1099,69 m3

QB = Vreq,B / 8 hr

QB = 0,038 m3/s = 137,46 m3/hr

Determination of volume flow rate of Tank C:

Qout,C = 0,04 m3/s = 144 m3/h

Volume of water which is out of tank after eight hour:

Vout,C =

Vout,C = 1152 m3

Volume of water between normal elevation and minimum elevation:

VC =

VC =

VC = 402,12m3

Amount of water requirement from pump to Tank C during eight hour:

Vreq,C = Vout,C – VC

Vreq,C = 1152 – 402,12 m3

Vreq,C = 749,88 m3

QC = Vreq,C / 8 hr

QC = 0,026 m3/s = 93,73 m3/hr

After these determinations volume flow rate of pump can be easily calculated:

QP = QA + QB + QC

QP = 0,010 m3/s + 0,038 m3/s + 0,026 m3/s

QP = 0.074 m3/s = 266,4

According to previous page calculations, flow rate distribution on each pipe can be shown.

48

Page 49: Branching Pipes Design

Pipe 1 Pipe 2 Pipe 3 Pipe 4 Pipe 5

Length (L) 1000 m 1400 m 1800 m 2200 m 1600 m

Diameter (D) 0,35 m 0,35 m 0,2 m 0,3 m 0,3 m

Roughness (ε) 3,0 mm 3,0 mm 3,0 mm 3,0 mm 3,0 mm

Flow Rate (Q) 0,074 m3/s 0,074 m3/s 0,038 m3/s 0,010 m3/s 0,026 m3/s

Table 4.2. – Flow Rate Distribution and Other Informations of Pipes

Second step is calculation of head loss of pump and on this purpose, Bernoulli equation must

be written between Tank A which is at the highest elevation of the system and Tank D.

Where ρwater = 1000 kg/m3 , zD = WSElnormal,D and zA = WSElnormal,A.Also Noting that the fluid at

both surfaces of tanksa re open to atmosphere and thus PD = PA = Patm. And because of the

velocities at wide tanks can be assumed as zero, VD = VA = 0. So, Bernoulli equation yields:

hpump = (zA – zD) + hL,total

Here it is, elevation differences are known but not hL,total. Calculation of this loss is:

hL,total = hL,major + hL,minor

hL,major =

Before identify friction coefficient f, first look to the flow in pipes which is laminar or

turbulent with Reynold number. μ = 1,002x10-3 kg/m.s

V1 =

49

Page 50: Branching Pipes Design

V2 =

V3 =

V4 =

V5 =

As it can seen all of the Reynold number of pipes are higher than critical Re value and thus

flow is turbulent. Which means Moody diagram should be used to find friction factor, f.

50

Page 51: Branching Pipes Design

hL,major =

hL,major = 38,473 m

hL,minor =

Where KL = 1,5 from Table 2.3 for branch tol ine condition and so:

hL,minor = 0,045 + 0,045 + 0,111 + 0,001 + 0,01

hL,minor = 0,212

hL,total = hL,major + hL,minor = 38,473 + 0,212 = 38,685 m

hpump = (zA – zD) + hL,total

hpump = (32 – 18) + 38,685

hpump = 52,685 m

With the help of Figure 3.3 and Standard pump selection program, pump can be selected as:

SNM 100 – 200 (3000) Singlestage Centrifugal Pump

QP = 266,4 m3/hr

hmax = 70 m

ηpump = 80%

N = 3000 rpm

Wpump,shaft =

51

Page 52: Branching Pipes Design

Figure 4.2. – Feasibility of Pump with Checking Power-Flow Rate Comparison

Figure 4.3. - Feasibility of Pump with Checking Head Loss-Flow Rate Comparison

52

Page 53: Branching Pipes Design

REFERENCES

1. Munson, B.R. Young D.F. and Okiishi T.H. Fundamentals of Fluid Mechanics, 4th

Ed., John Wiley & Sons Inc., New York, 2002.

2. Hinze, J. O. Turbulence, 2nd Ed., McGraw-Hill, New York, 1975.

3. White, F. M. Fluid Mechanics, 4th Ed. McGraw-Hill, New York, 1979.

4. Moody, L. F. Friction Factors for Pipe Flow, Transactions of the ASME, Vol.66.1944

5. Streeter, V. L. and Wylie, E. B. Fluid Mechanics, 8th Ed. McGraw-Hill, New York,

1985.

6. Jeppson, R. W. Analysis of Flow in Pipe Networks, Ann Arbor Science Publishers,

Ann Arbor, Mich., 1976.

7. Hydraulic Institute, Engineering Data Book, 1st Ed., Cleveland Hydraulic Institute,

1979.

8. Karassick, I. J. Pump Handbook, 2nd Ed., McGraw-Hill, New York, 1985.

9. Nikuradse, J. Stomungsgesetz in Rauhen Rohren, VDIForschungsch, No. 361

10. Riley, W. F. and Sturges, L. D. Engineering Mechanics:Dynamics, 2nd Ed., Wiley,

New York, 1996.

11. Çengel, Y. A. and Cimbala, J. M. Fluid Mechanics, 4th Ed., McGraw-Hill, New York.

12. Thorley, A. R. D. Fluid Transients in Pipeline Systems, D&L, England, 1991.

13. Watters, G. Z. Analysis and Control of Unsteady Flow in Pipelines, 2nd Ed.,

Butterworths, Massachusetts, 1984.

14. Chaudry, M. H. Applied Hydraulic Transients, 2nd Ed., Van Nostrand Reinhold, New

York, 1987.

53