0% found this document useful (0 votes)
256 views

Chapter 4 Pipe Flow

This chapter of the National Engineering Handbook covers pipe flow and hydraulics. It includes sections on common pipe flow equations, local losses, pumps, pipe networks, modeling software, appurtenances, transients, and applications to culverts and irrigation systems.

Uploaded by

trabajo
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
256 views

Chapter 4 Pipe Flow

This chapter of the National Engineering Handbook covers pipe flow and hydraulics. It includes sections on common pipe flow equations, local losses, pumps, pipe networks, modeling software, appurtenances, transients, and applications to culverts and irrigation systems.

Uploaded by

trabajo
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 76

Title 210 – National Engineering Handbook

Part 634 Hydraulics


National Engineering Handbook

Chapter 4 Pipe Flow

(210-634-H, 1st Ed., Feb 2022)


Title 210 – National Engineering Handbook

Issued: Feb 2022_

In accordance with Federal civil rights law and U.S. Department of


Agriculture (USDA) civil rights regulations and policies, the USDA, its
Agencies, offices, and employees, and institutions participating in or
administering USDA programs are prohibited from discriminating based on
race, color, national origin, religion, sex, gender identity (including gender
expression), sexual orientation, disability, age, marital status,
family/parental status, income derived from a public assistance program,
political beliefs, or reprisal or retaliation for prior civil rights activity, in any
program or activity conducted or funded by USDA (not all bases apply to all
programs). Remedies and complaint filing deadlines vary by program or
incident.

Persons with disabilities who require alternative means of communication


for program information (e.g., Braille, large print, audiotape, American Sign
Language, etc.) should contact the responsible Agency or USDA's TARGET
Center at (202) 720-2600 (voice and TTY) or contact USDA through the
Federal Relay Service at (800) 877-8339. Additionally, program information
may be made available in languages other than English.

To file a program discrimination complaint, complete the USDA Program


Discrimination Complaint Form, AD-3027, found online at How to File a
Program Discrimination Complaint and at any USDA office or write a letter
addressed to USDA and provide in the letter all of the information requested
in the form. To request a copy of the complaint form, call (866) 632-9992.
Submit your completed form or letter to USDA by: (1) mail: U.S.
Department of Agriculture, Office of the Assistant Secretary for Civil
Rights, 1400 Independence Avenue, SW, Washington, D.C. 20250-9410; (2)
fax: (202) 690-7442; or (3) email: [email protected].

USDA is an equal opportunity provider, employer, and lender.

Disclaimer of Endorsement: Reference to any specific product, trademark,


or manufacturer this handbook does not constitute an implied or expressed
recommendation or endorsement or favoring by NRCS.

(210-634-H, 1st Ed., Feb 2022)


634-4.i
Title 210 – National Engineering Handbook

Chapter 4 Pipe Flow

Table of Contents
634.0400 General .............................................................................................................................1
634.0401 Friction Loss Methods .....................................................................................................2
634.0402 Manning’s Equation for Pipelines ...................................................................................2
634.0403 Darcy-Weisbach Equation and Friction Factor ...............................................................6
634.0404 Pipe Flow Solutions Using Hazen-Williams Formula ..................................................10
634.0405 Local Losses in Pipelines ..............................................................................................14
634.0406 Pumps in Pipelines ........................................................................................................20
634.0407 Pipelines and Networks .................................................................................................25
634.0407 EPANET .......................................................................................................................33
634.0408 WaterCAD.....................................................................................................................37
634.0409 Appurtenances in Pipelines and Networks ....................................................................38
634.0410 Hydraulic Transients (Water Hammer) .........................................................................51
634.0411 Cavitation ......................................................................................................................54
634.0412 Culverts .........................................................................................................................57
634.0413 Sprinkler Irrigation ........................................................................................................68
634.0414 Microirrigation ..............................................................................................................69
634.0415 References .....................................................................................................................71

Table of Figures
Figure 4-1: Energy Heads in Pipe Flow .........................................................................................1
Figure 4-2: Pipe Flow calculations with the USDA-NRCS Hydraulics Formula ..........................4
Figure 4-3: Values of Manning’s resistance coefficient for pipe ...................................................5
Figure 4-4: Determining Velocity in a Pipeline Draining a Reservoir...........................................6
Figure 4-5: Absolute roughness values for pipe materials ..............................................................7
Figure 4-6: Iterative Solution for Q ...............................................................................................9
Figure 4-7: Determining Flow in a Pipeline Draining a Reservoir using the Darcy-Weisbach
Equation ...................................................................................................................................10
Figure 4-8: Values of the Hazen-Williams coefficient. ...............................................................12
Figure 4-9: Determining Discharge and Velocity in a Pipeline Draining a Reservoir using the
Hazen-Williams Formula.........................................................................................................13
Figure 4-10. Discharge from a pipe into a reservoir. ....................................................................14

(210-634-H, 1st Ed., Feb 2022)


634-4.ii
Title 210 – National Engineering Handbook

Figure 4-11. Entrance loss coefficients for typical pipe entrance shapes ......................................15
Figure 4-12. Pipe entrance loss coefficients. .................................................................................16
Figure 4-13: Solution for pipe flow with entrance loss using USDA-NRCS Hydraulics Formula.16
Figure 4-14: Local loss coefficients for selected pipe fittings. .....................................................17
Figure 4-15: Schematic of pipe flow between two reservoirs. .....................................................17
Figure 4-16: Solution for pipe flow including all local losses using USDA-NRCS Hydraulics
Formula ....................................................................................................................................20
Figure 4-17: Discharge versus Head from Pump Tests................................................................21
Figure 4-18: Pump discharge-head graph ....................................................................................21
Figure 4-19: Manufacturer-provided Pump Curves and Pump Sizes...........................................22
Figure 4-20: Pump-pipeline System for Analysis. .......................................................................22
Figure 4-21: System Curve (Tabular) ..........................................................................................23
Figure 4-22: Pump System Curve (Graphical) .............................................................................23
Figure 4-23: System Schematic showing Motor M, the Electric Supply E, and Pump P ............24
Figure 4-24: Three Pipelines in Series Connecting Two Reservoirs ...........................................25
Figure 4-25 Local Head Loss Coefficients for a Sudden Pipe Contraction ..................................26
Figure 4-26: Effect of neglecting local losses on an example pipeline system ............................29
Figure 4-27: Three pipelines in parallel connecting two reservoirs. ............................................30
Figure 4-28: Three pipelines converging to a single delivery point.............................................31
Figure 4-29 Hydraulic example for EPANET ...............................................................................34
Figure 4-30: Pump Curve editor screen shot. Only one point is entered for this example and the
curve is calculated from the program. .....................................................................................35
Figure 4-31: Screen shot of EPANET output...............................................................................36
Figure 4-32: Output table of Junction information for EPANET example ..................................36
Figure 4-33: Output table of links (Pipelines) for EPANET example. ........................................36
Figure 4-34: Continuous acting air valve .....................................................................................39
Figure 4-35: Location and dimensions of air vent chamber near a water intake. ........................45
Figure 4-36: Spring operated pressure relief valve ......................................................................46
Figure 4-37: Schematic of pump-pipeline system protected with a surge chamber.....................49
Figure 4-38: Check valve installed on a pumping plant...............................................................51
Figure 4-39 Siphon conduit connecting two reservoirs. ................................................................54
Figure 4-40: Flow regimes for submerged inlet flow in culverts ..................................................57
Figure 4-41: Culvert discharge calculation ..................................................................................60
Figure 4-42: Culvert Discharge Convergence of Values .............................................................61
Figure 4-43: Culvert discharge calculation using USDA-NRCS Hydraulics Formula Tool ......61
Figure 4-44: Nomogram for headwater depth for concrete pipe culverts with inlet control. ........63
(210-634-H, 1st Ed., Feb 2022)
634-4.iii
Title 210 – National Engineering Handbook

Figure 4-45: Nomogram for headwater depth for corrugated metal (CM) pipe culverts with inlet
control. .....................................................................................................................................65
Figure 4-46: Nomogram for concrete pipe culverts flowing full with outlet control, Manning’s n =
0.012. .......................................................................................................................................66
Figure 4-47: Nomogram for head for corrugated metal (CM) pipe culverts flowing full with outlet
control, n = 0.024.....................................................................................................................67
Figure 4-48: Central pivot irrigation system near Grace, Idaho...................................................68
Figure 4- 49: Schematic of a sprinkler system layout ..................................................................69
Figure 4-50: A manifold with drip tape on a corn field ...............................................................70

(210-634-H, 1st Ed., Feb 2022)


634-4.iv
Title 210 – National Engineering Handbook

Part 634 – Hydraulics


Chapter 4 – Pipe Flow

634.0400 General

A. Pipe flow exists when a closed conduit of any form is flowing f'u1l of water. In pipe flow
the cross-sectional area of flow is fixed by the cross section of the conduit and the water
surface is not exposed to the atmosphere. The internal pressure in a pipe may be equal to,
greater than, or less than the local atmospheric pressure.
B. The principles of pipe flow apply to the hydraulics of such structures as some culverts,
drop inlets, siphons, and various types of pipelines.
C. Consider figure 4-1 which shows the hydraulic grade line, energy line and the energy
heads at two points of the flow. Most pipe flow occurs through constant-diameter pipe, thus,
making the velocity heads at sections (1) and (2) the same (V1 = V2 = V). This in turn means
that the energy line (E.L.) and hydraulic grade line (H.G.L.) are parallel. The energy equation
for constant-diameter pipe is written as:
𝑧 𝑧 ℎ (eq. 4-1)

Where hf represents the energy head losses between sections (1) and (2) in figure 4-1.
Because the energy head losses are due to the friction between the moving fluid mass and the
walls of the pipe, hf is also referred to as the friction losses. Friction losses may be calculated
by measuring the piezometric heads,
ℎ 𝑧 and ℎ 𝑧
and taking the difference
ℎ ℎ ℎ

Figure 4-1: Energy Heads in Pipe Flow

(210-634-H, 1st Ed., Feb 2022)


634-4.1
Title 210 – National Engineering Handbook

D. The energy slope, or slope of the energy line, is the ratio of the friction losses, hf, to the
length of the pipe, L (x in figure 4-1), and may be expressed as:

𝑆 𝑧 𝑧 (eq. 4-2)

E. Flow in pipelines occurs in two different regimes, laminar and turbulent. Laminar flow
occurs at relatively low velocities and is characterized by fluid particles moving in stable
flow layers and with a parabolic velocity distribution. Increasing the flow velocity
sufficiently causes instabilities of the flow layers which, in turn, cause the flow to become
turbulent. In turbulent flow the fluid particles no longer move in stable layers, but rather
clump together to form different-sized eddies. Turbulent flow is characterized by a more
uniform velocity distribution across the pipe typically described mathematically by
logarithmic or power-law functions.

634.0401 Friction Loss Methods

Calculation of friction losses is an important step Solving the energy equation, in analyzing
flow in pipes, and in the selection of pipe sizes for specific applications. The easiest method
of calculating friction losses in pipe can be performed using empirical equation. Several of
the more common equation are Manning’s equation, the Darcy-Weisbach equation, or the
Hazen-Williams formula.

634.0402 Manning’s Equation for Pipelines

A. Manning’s equation for open-channel flow was presented as 634.03 equation 3-42 in
terms of the flow velocity V, the hydraulic radius R, the channel bed slope So , the Manning’s
resistance coefficient n, and a constant Cu that depends on the system of units used (Cu = 1.0
for the International System, and Cu = 1.486 for the English System). When applied to
pipelines, the bed slope So is replaced by the energy slope Sf = hf /L, while the hydraulic
radius is calculated in terms of the area, A (634.03 equation 3-2 repeated here), and wetted
perimeter, P, of a full circular cross-section of diameter D, using:
𝜋𝐷
𝐴
4
𝑃 𝜋𝐷 (eq. 4-3)
which results in:

𝑅 (eq. 4-4)

B. With these substitutions, the Manning’s equation for a pipeline results in:

𝑉 (eq. 4-5)

C. Typically, the equation is rewritten by isolating hf:

ℎ 𝑉 (eq. 4-6)

(210-634-H, 1st Ed., Feb 2022)


634-4.2
Title 210 – National Engineering Handbook

D. Equations based on the above equation (velocity) for both the International System (SI)
and English System (ES) follow:
(1) International System (SI): hf (m), Cu = 1.0, L(m), D(m), V(m/s)

ℎ 6.3496 𝑉 (eq. 4-7)

(2) English System (ES): hf (ft), Cu = 1.486, L(ft), D(ft), V (ft/s, fps)

ℎ 2.8755 𝑉 (eq. 4-8)

E. Sometimes it is preferred to work with the discharge Q, instead of the flow velocity V. For
circular pipelines of diameter D, the equation of continuity is written as either 634.03
equation 3-3 or equation 3-4, which are reproduced below:
𝜋𝐷
𝑄 𝑉𝐴 𝑉
4
4𝑄
𝑉
𝜋𝐷
F. Substituting equation 3-4 into equation 4-5, and solving for Q:

𝑄 (eq. 4-9)

or, solving for hf.

ℎ 𝑄 (eq. 4-10)

G. Equations based on the above equation (discharge) for both the International System (SI)
and English System (ES) follow:
(1) International System (SI): hf(m), Cu = 1.0, L(m), D(m), Q(m3/s)

ℎ 10.2936 𝑄 (eq. 4-11)

(2) English System (ES): hf (ft), Cu = 1.486, L (ft), D (ft), Q (ft3/s, cfs)

ℎ 4.6615 𝑄 (eq. 4-12)

H. A demonstration of how to calculate a pipeline’s discharge using Manning’s equation can


be found in Example 4-1
(1) Example 4-1 – To determine the discharge Q that can be conveyed by a 3.0-in-
diameter corrugated-plastic pipeline if a head hf = 10 ft, is to be dissipated in a length
L = 100 ft use equation 4-9 to calculate the discharge. The data to use are the
following: D = 3.0 in = 3.0/12 = 0.25 ft, hf = 10 ft, L = 100 ft, Cu = 1.486, n = 0.015
(from fig 4-2). The resulting discharge is:

𝜋𝐶 𝐷 ℎ 3.1416 1.486 0.25 10


𝑄 0.24 𝑐𝑓𝑠
4 𝑛 𝐿 4 0.015 100

(210-634-H, 1st Ed., Feb 2022)


634-4.3
Title 210 – National Engineering Handbook

I. The software USDA-NRCS Hydraulics Formula provides for the calculation of pipe flow
using Manning’s equation. To activate this solution, select the Pipe Flow tab, which produces
the entry form shown in figure 4-2. The formula shown in the entry form is equivalent to
equation 4-09, but with the diameter D in inches and with local loss coefficients equal to zero
(i.e., Ke = 0, Kb = 0 reference 634.0405).
(1) Example 4-2 – Pipeline discharge calculation using the USDA-NRCS Hydraulics
Formula software Using the Pipe Flow tab in the USDA-NRCS Hydraulics Formula
software, with Ke = 0, Kb = 0, determine the discharge Q conveyed by a 3-in-
diameter corrugated-plastic pipeline if a head hf = 10 ft, is to be dissipated in a length
L = 100 ft. These data are the same as for the previous example 4-1. The resulting
discharge is 0.24 cfs. The equation for pipe flow shown in the figure 4-2 will be
derived after introducing the concept of local losses in pipe (see section 634.0405).

Figure 4-2: Pipe Flow calculations with the USDA-NRCS Engineer Field Tools
Hydraulics Formulas

J. One of the more important steps of using the Manning’s equation is selecting the correct n
value. Figure 4-3 gives a range of values for various materials. Manning’s n values, shown in
figure 4-3, are assembled from many hydraulic references, including Brater and King (1996),
FHWA (2001), and USACE (2008).

(210-634-H, 1st Ed., Feb 2022)


634-4.4
Title 210 – National Engineering Handbook

Figure 4-3: Values of Manning’s resistance coefficient for pipe


Manning's n
Pipe material
Minimum Design Maximum
Ductile iron, unlined 0.012 0.012 - 0.015 0.016
Ductile iron, polyurethane coated 0.010 0.012 - 0.014 0.014
Ductile iron, cement coated 0.011 0.012 - 0.014 0.015
Cast-iron, coated 0.010 0.012 - 0.014 0.014
Cast-iron, uncoated 0.011 0.013 - 0.015 0.015
Wrought iron, galvanized 0.013 0.015 - 0.017 0.017
Wrought iron, black 0.012 0.013 – 0.014 0.015
Steel, riveted and spiral 0.013 0.015 - 0.017 0.017
Annular corrugated metal (1) 0.021 0.021 - 0.025 0.025
Helical corrugated metal (1) 0.013 0.015 - 0.020 0.021
Wood stave 0.010 0.012 - 0.013 0.014
Neat cement surface 0.010 0.013
Concrete 0.010 0.012 - 0.017 0.017
Vitrified sewer pipe 0.010 0.013 - 0.015 0.017
Clay, common drainage tile 0.011 0.012 - 0.014 0.017
Corrugated plastic 0.014 0.015 - 0.016 0.017
PVC 0.009 -0.011
Smooth Interior PE 0.009 - 0.015 0.02
Aluminum 0.01
Gated Aluminum Pipe 0.013
(1)N-values for corrugated metal pipe vary with pipe diameter. See FHWA (2001) or USACE (2008) to select
a refined n-value.

K. Other examples of using the Manning’s equation to solve for different variables are
shown below.
(1) Example 4-3 – Using equation 4-5, determine flow velocity in a helical corrugated
metal pipe. Use n = 0.016 (from fig. 4-3), for a 5-in diameter pipe that dissipates a
head of hf =6.5 ft in a length L = 300 ft. Using D = 5 in = 5/12 ft = 0.4167 ft, and Cu
= 1.486, equation 4-5 produces the following result:

𝐶 𝐷 ℎ 1.486 0.4167 6.5


𝑉 3.0 𝑓𝑡⁄𝑠
𝑛 4 𝐿 0.016 4 300
(2) Example 4-4 – Using equation 4-12, determine the head loss in 1500 ft of a riveted
and spiral steel pipe (use n = 0.015) with a 24-in (2-ft) diameter that carries a
discharge of 10 cfs. Using equation 4-12:
𝑛 𝐿 0.015 1500
ℎ 4.6615 𝑄 4.6615 10 3.9𝑓𝑡
𝐷 2
(3) Example 4-5 – Determine the velocity in a pipeline draining a reservoir using
Manning’s equation. Consider a reservoir whose free surface is located at an
elevation z1 = 60 ft, draining through a 0.5-ft-diameter, 100-ft-long, concrete pipe (n
= 0.012) open to the atmosphere whose outlet is located at an elevation z2 = 55 ft.
The system is depicted in figure 4-4. Entrance losses will be ignored.

(210-634-H, 1st Ed., Feb 2022)


634-4.5
Title 210 – National Engineering Handbook

Figure 4-4: Determining Velocity in a Pipeline Draining a Reservoir

Point 1 in the energy equation is at the reservoir free surface where p1 = 0 and V1 =
0. Point 2 is at the pipe outlet where p2 = 0 and V2 = V, the pipe velocity. The energy
head, H = z1 – z2 = 60 ft – 55 ft = 5 ft, for this case. Applying the energy equation
(634.03 equation 3-19) between points 1 and 2:
𝑝 𝑉 𝑝 𝑉
𝑧 𝑧 ℎ
𝜔 2𝑔 𝜔 2𝑔
Using Manning’s equation (eq. 4-8) to represent friction losses:
0 0 0 𝑉 𝑛 𝐿
𝐻 2.8755 𝑉
𝜔 2𝑔 𝜔 2𝑔 𝐷
This simplifies to:
1 2.8755𝑛 𝐿
𝐻 𝑉
2𝑔 𝐷
Solving for the velocity, V, and using H = 5 ft, L = 100 ft, n = 0.012, D = 0.5 ft, g =
32.2

𝐻 5
𝑉 6.46 𝑓𝑡/𝑠
1 2.8755𝑛 𝐿 1 2.8755 0.012 100
2𝑔 𝐷 2 32.2 0.5

634.0403 Darcy-Weisbach Equation and Friction Factor

A. A second method for calculating friction losses in pipes is the Darcy-Weisbach equation
written, in terms of the flow velocity V as:

ℎ 𝑓 (eq. 4-13)

B. The friction factor, f, is a function of a relative roughness, e/D, where “e” is known as the
absolute roughness or equivalent sand roughness, and the Reynolds number of the flow.
C. The Reynolds number, defined in 634.0104 equation 1-10, is repeated here:
𝜌𝑉𝐷 𝑉𝐷
𝑅
𝜇 
where ρ is the density (mass per unit volume) of water, μ is the absolute or dynamic viscosity
of water, and ν is the kinematic viscosity of water, defined by 634.01 equation 1-9, ν= μ/ρ.

(210-634-H, 1st Ed., Feb 2022)


634-4.6
Title 210 – National Engineering Handbook

Values of the density and viscosity of water related to temperature are available in 634.0107
Exhibit D, figures 1-9 (English Units) and 1-10 (SI Units).
D. The absolute roughness of a pipe material is the average height of the irregularities of the
inner wall of the pipe. The first experiments on head losses in pipes were conducted in the
early 20th century by coating glass pipes with uniform-size sand grains. The diameter of the
sand grains was used to represent the absolute roughness of the pipe, e. Typical values of the
absolute roughness of various pipe materials are presented in Figure 4-5. Absolute roughness
values may be found in Streeter and Wylie (1998) and other similar fluid mechanics texts.
Figure 4-5: Absolute roughness values for pipe materials
Pipe material e (mm) e (ft)
Centrifugally spun cement 0.0015 0.000005
Concrete 0.3-3 0.001-0.01
Bituminous lining 0.0015 0.000005
Asphalt-dipped cast iron 0.12 0.0004
Cast iron 0.25 0.00085
Ductile iron 0.061 0.0002
Galvanized iron 0.15 0.0005
Wrought iron 0.046 0.00015
Commercial steel 0.046 0.00015
Riveted steel 0.9-9 0.003-0.03
Welded-steel pipe 0.046 0.00015
PVC, PE, HDPE pipe 0.002 0.0000066
Aluminum, with couplers 0.13 0.00043
Smooth surface (glass, plastic) 0.0015 0.000005
Drawn tubing, brass, lead, copper 0.0015 0.000005
Wood stave 0.18-0.9 0.0006-0.003

E. The Darcy-Weisbach friction factor has different expressions for laminar or turbulent pipe
flow. Water flowing at a very low velocity, or in a very small-diameter pipe, typically flows
in a laminar flow regime. In laminar flow the flow takes place in layers (Latin, laminae) that
remain very stable and are easily identifiable by dye injected into the flow. As the velocity
increases, conditions are reached in which the flow becomes turbulent. In turbulent flow,
layers of flow are no longer identifiable, and the flow tends to break down into eddies that
facilitate mixing. Laminar flow is rare, occurring at low flow velocity or in small pipe
diameter. Most pipe flows of interest are turbulent.
F. Experiments to determine the laminar or turbulent nature of flow were carried out by
Osborne Reynolds in the 19th century. Reynolds identified a dimensionless parameter, now
known as the Reynolds number (634.0104 equation 1-10), to determine if a flow is laminar or
turbulent. There is also a transitional flow regime in which the flow is neither laminar nor
turbulent but shifts between conditions. The typical values for classifying flows according to
laminar, transitional, or turbulent regimes are shown next:
(1) Laminar flow: Re <2000
(2) Transitional flow: 2000 < Re < 4000
(3) Turbulent flow: Re > 4000
The range Re < 4000 that encompasses laminar and transitional flow is a small range
considering that Re can take values to 109 or larger. Thus, turbulent flow is more likely to
occur in pipelines.

(210-634-H, 1st Ed., Feb 2022)


634-4.7
Title 210 – National Engineering Handbook

G. Example 4-6 How Reynolds number is used to classify pipe flow. Water is flowing in a 2-
in diameter pipe at a rate of 0.08 cfs. The input data given is D = 2 in = 2in/12in/ ft = 0.167
ft, Q = 0.08 cfs, and from Title 210, National Engineering Handbook, Part 634, Chapter 1,
(210-634-1), Section 634.0107 Exhibit D (634.01.7D), figures 1-8 (English Units) and 1-9 (SI
Units) at a temperature of 70ºF the kinematic viscosity of water is ν =.00001059 ft2/s =
1.059x10 -5 ft2/s.
(1) If the water temperature is 70ºF, determine the Reynolds number of the flow and
classify it as laminar, transitional, or turbulent.
(i) The flow velocity is calculated as:
𝑓𝑡
𝑄 4𝑄 4 0.08 𝑠 𝑓𝑡
𝑉 3.7 𝑠
𝐴 𝜋𝐷 3.1416 0.167
(ii) And the Reynolds number is:
𝑓𝑡
𝑉𝐷 3.7 𝑠 0.167𝑓𝑡
𝑅 58347 5.83𝑥10
 1.059𝑥10
𝑓𝑡
𝑠
(iv) Since Re > 4000, the flow is turbulent.

(2) In the laminar regime, the Darcy-Weisbach friction factor, f, is a function of the
Reynolds number, given as:

𝑓 (eq. 4-14)

(3) In the turbulent regime, an equation relating the friction factor, f, the Reynolds
number Re, and the relative roughness, e/D, is the Colebrook-White equation:
.
2 log (eq. 4-15)
.
(4) The difficulty in using this equation is that it is not explicit in f and any solution
involving this equation requires a numerical approach. To facilitate solving explicitly
for f, a close approximation (2 – 5% error) to the Colebrook-White equation is
provided by the Swamee-Jain equation:
.
𝑓 . (eq. 4-16)
. .

H. Pipe Flow Solutions Using the Darcy-Weisbach Equation –


(1) Solutions of turbulent pipe flow using the Darcy-Weisbach equation are often
convenient to write in terms of discharge since the discharge, Q, is often known or
can be calculated (equation 4-17).

ℎ (eq. 4-17)

(2) The Reynolds number Re can also be written in terms of the discharge Q as follows:

𝑅 (eq. 4-18)

(210-634-H, 1st Ed., Feb 2022)


634-4.8
Title 210 – National Engineering Handbook

(3) With this definition of the Reynolds number, the Swamee-Jain equation becomes:
.
𝑓  . (eq. 4-19)
. .

(4) Combining the Darcy-Weisbach equation (equation 4-17) with the Swamee-Jain
equation (equation 4-19), and solving for the discharge, Q, produces the following
equation:
 .
𝑄 2.22 𝑙𝑜𝑔 0.27 4.62 (eq. 4-20)

(5) This equation’s variables are best solved by using a numerical spreadsheet
application. Typically, there are three types of problems involving pipe friction
losses, namely:
 Head loss problem: calculate hf given D, Q or V, and g, L, e, ν.
 Discharge problem: calculate Q or V, given D, hf and g, L, e, ν.
 Sizing problem: calculate D, given Q, hf and g, L, e, ν.
Where hf is the friction loss, D is the pipe diameter, Q is the discharge, V is the
velocity, g is the acceleration of gravity, L is the pipe length, e is the absolute
roughness, and ν is the kinematic viscosity.
I. Examples 4-7 through 4-9 demonstrate solving for pipe flow solutions with the Darcy-
Weisbach equation
(1) Example 4-7 - Given D = 0.3 ft, Q = 0.20 cfs, g = 32.2 ft/s2, L = 1000 ft, e = 0.002 in
= 0.000166 ft, and ν = 1.13x10-5 ft2/s, find hf.
(i) A numerical spreadsheet application of equation 4-20 gives hf = 8.86 ft.
(ii) Alternately, equation 4-20 can be re-arranged and solved directly for hf, with a
hand calculator.
(2) Example 4-8 - Given D = 0.7 ft, hf = 15 ft, g = 32.2 ft/s2, L = 750 ft, e = 0.005 in =
0.000416 ft, and ν = 1.2x10-5 ft2/s, find Q.
(i) A numerical spreadsheet application of equation 4-20 gives Q = 2.68 cfs.
(ii) Alternately, equation 4-20 can be solved iteratively for Q, with a hand calculator,
as shown in figure 4-6:

Figure 4-6: Iterative Solution for Q


Trial Q Calculated Q
1.5 2.63
2.1 2.66
2.6 2.681
2.7 2.684
2.68 2.683

(3) Example 4-9 - Given Q = 3 cfs, hf = 10 ft, g = 32.2 ft/s2, L = 1500 ft, e = 0.01 in =
0.000833 ft, and ν = 1.5x10-5 ft2/s, find D.
A numerical spreadsheet application of equation 4-20 gives D = 0.8591 ft x 12 in/ft =
10.29 in. The most likely value of commercial pipeline that can be used is 10 or 12
inch.

(210-634-H, 1st Ed., Feb 2022)


634-4.9
Title 210 – National Engineering Handbook

(4) Example 4-10 - Flow in pipeline draining a reservoir using the Darcy-Weisbach
equation. Consider again (see example 4-5) a reservoir whose free surface is located
at an elevation z1 = 60 ft, draining through a 0.5-ft-diameter, 100-ft-long, concrete
pipe (e = 0.003 ft) open to the atmosphere whose outlet is located at an elevation z2 =
55 ft. The system, which carries water at a temperature of 55ºF, is depicted in figure
4-7. Minor losses at the entrance from the reservoir into the pipe are ignored.
Determine the discharge and velocity in the pipeline.

Figure 4-7: Determining Flow in a Pipeline Draining a Reservoir using the


Darcy-Weisbach Equation

(i) As in example 4-5, the energy equation (634.03 equation 3-20) is applied, but the
friction losses are estimated with the Darcy-Weisbach equation instead of
Manning’s equation. The energy equation, in terms of discharge, Q, simplifies to:
8𝑄 𝐿
𝐻 1 𝑓
𝜋 𝑔𝐷 𝐷
(ii) Re-arrange the equation to solve for Q and approximate the friction factor, f, by
the Swamee-Jain equation (equation 4-19). Then using H = 5 ft, L = 100 ft, e =
0.003 ft, ν = 1.3135 x 10-5 ft2/s, D = 0.5 ft, g = 32.2 ft/s2, and applying a
numerical spreadsheet solution gives Q =1.285 cfs.
(iii) And the flow velocity is:
𝑓𝑡
𝑄 4𝑄 4 1.285 𝑠
𝑉 6.54 ft/s
𝐴 𝜋𝐷 3.1416 0.5 𝑓𝑡

634.0404 Pipe Flow Solutions Using Hazen-Williams Formula

A. A third method for calculating friction head losses in pipes is the Hazen-Williams formula.
The Hazen-Williams formula was developed from empirical data of water flow in pipes. In
terms of flow velocity, V, the Hazen-William formula is expressed as follows:
(1) International System (SI): V(m/s)
. .
𝑉 0.849𝐶 𝑅 𝑆 (eq. 4-21)
(2) English System (ES): V (ft/s, fps)
. .
𝑉 1.318𝐶 𝑅 𝑆 (eq. 4-22)
where V is velocity, CHW is the Hazen-Williams coefficient, R is the hydraulic radius
(R=D/4), and Sf is the energy slope (Sf=hf /L).

(210-634-H, 1st Ed., Feb 2022)


634-4.10
Title 210 – National Engineering Handbook

B. Using the definitions of R and Sf, the Hazen-Williams formula may be written as:
(1) International System (SI): hf(m), L(m), D(m), V(m/s)
.
.
𝑉 0.354𝐶 𝐷 (eq. 4-23)

(2) English System (ES): hf (ft), L (ft), D (ft), V (ft/s, fps)


.
.
𝑉 0.550𝐶 𝐷 (eq. 4-24)

B. In terms of the discharge, the Hazen-Williams formula is written as:


(1) International System (SI): hf(m), L(m), D(m), Q(m3/s)
.
.
𝑄 0.278𝐶 𝐷 (eq. 4-25)

(2) English System (ES): hf(ft), L(ft), D(ft), Q(cfs)


.
.
𝑄 0.432𝐶 𝐷 (eq. 4-26)

C. Solving for the head loss, hf, in the Q-based equations:


(1) International System (SI): hf(m), L(m), D(m), Q(m3/s)
.
ℎ 10.704 .
(eq. 4-27)

(2) English System (ES): hf(ft), L(ft), D(ft), Q(cfs)


.
ℎ 4.732 .
(eq. 4-28)

D. Solving for the diameter, D, in the Q-based equations:


(1) International System (SI): hf(m), L(m), D(m), Q(m3/s)
. .
𝐷 1.627 (eq. 4-29)

(2) English System (ES): hf(ft), L(ft), D(ft), Q(cfs)


. .
𝐷 1.376 (eq. 4-30)

(210-634-H, 1st Ed., Feb 2022)


634-4.11
Title 210 – National Engineering Handbook

E. Figure 4-8, shows typical values for the Hazen-Williams coefficient. Hazen-Williams
coefficients may be found in Lamont (1981), Mays (1991) and other similar publications.

Figure 4-8: Values of the Hazen-Williams coefficient.


Pipe description Condition CHW
Very smooth Straight alignment 140
Slight curvature 130
Cast iron, uncoated or New 130
steel pipe 5 years old 120
or steel pipe 10 years old 110
15 years old 100
20 years old 90
30 years old 80
Cast iron, coated All ages 130
Wrought iron or Diameter 12 in. and up 110
standard galvanized Diameter 4 to 12 in. 100
steel Diameter 4 in or less 80
Brass or lead New 140
Concrete Very smooth, excellent joints 140
Smooth, good joints 120
Rough 110
Vitrified clays 110
Smooth wooden; wood 120
stave
Asbestos cement 140
Corrugated metal 60
Pipes of small diameter old, rough inside surface, as low as 40
PVC, PE, HDPE 150
Aluminum 120
Aluminum gated pipe 110

F. Solutions to pipe-flow problems with the Hazen-Williams formula require the use of
equations (equation 4-21) through (equation 4-30), depending on the data given and which
variable is being solved for. This is demonstrated in examples 4-11 through 4-15.
(1) Example 4-11 – Pipe velocity calculation using the Hazen-Williams formula. A
pipeline with a length L = 1200 ft, diameter D = 0.75 ft, suffers a head loss hf = 12 ft.
If the pipe is made of 5-year-old steel pipe (CHW = 120), determine the pipe velocity.
Using equation 4-24 as follows:
. .
.
ℎ .
12
𝑉 0.550𝐶 𝐷 0.550 120 0.75 4.58 𝑓𝑡/𝑠
𝐿 1200

(210-634-H, 1st Ed., Feb 2022)


634-4.12
Title 210 – National Engineering Handbook

(2) Example 4-12 – Pipe discharge calculation using the Hazen-Williams formula.
Determine the discharge Q for a rough concrete pipeline (CHW = 110) with a diameter
D = 1.5 ft that produces a head loss hf = 8.5 ft on a length of L = 650 ft.
Using equation 4-26 as follows:
. .
.
ℎ .
8.5
𝑄 0.432𝐶 𝐷 0.432 110 1.5 13.27 𝑐𝑓𝑠
𝐿 650
(3) Example 4-13 – Head loss calculation using the Hazen-Williams formula. Determine
the head loss hf in a new PVC pipeline (CHW = 150) of length L = 2000 ft and
diameter D = 3.0 ft, that carries a discharge Q = 20 cfs.
Using equation 4-28 as follows:
. .
𝑄 𝐿 20 2000
ℎ 4.732 .
4.732 1.08𝑓𝑡
𝐶 𝐷 150 3 .
(4) Example 4-14 – Diameter D calculation using the Hazen-Williams formula.
Determine the diameter of coated cast iron (CHW = 130) pipe that produces a friction
head loss hf = 20 ft in a length L = 500 ft while carrying a discharge Q = 25 cfs.
(i) Using equation 4-30 as follows:
. . . .
𝑄 𝐿 25 500
𝐷 1.376 1.376 1.42𝑓𝑡
𝐶 ℎ 130 20
(ii) The most likely value of commercial pipeline that can be used is 1.5ft.

(5) Example 4-15 – Flow in pipeline draining a reservoir using Hazen-Williams formula.
Consider once more (see examples 4-5 and 4-10) the case of a reservoir whose free
surface is located at an elevation z1 = 60 ft, draining through a 0.5-ft-diameter, 100-ft-
long, rough concrete pipe (CHW = 110) open to the atmosphere whose outlet is
located at an elevation z2 = 55 ft. The system is depicted in figure 4-9. Local losses at
the entrance from the reservoir into the pipe are ignored. Determine the discharge and
velocity in the pipeline.

Figure 4-9: Determining Discharge and Velocity in a Pipeline Draining a


Reservoir using the Hazen-Williams Formula

(i) As in examples 4-5 and 4-10, the energy equation (634.0302 equation 3-13) is
applied, but the friction losses are estimated with the Hazen-Williams formula.
The energy equation, in terms of discharge, Q, simplifies to:
.
8𝑄 𝑄 𝐿
𝐻 4.732 .
𝜋 𝑔𝐷 𝐶 𝐷

(210-634-H, 1st Ed., Feb 2022)


634-4.13
Title 210 – National Engineering Handbook

(ii) Using H = 5 ft, L = 100 ft, CHW = 110, D = 0.5 ft, g = 32.2 ft/s2, and applying a
numerical spreadsheet solution gives Q = 1.385 cfs, and the flow velocity is:
𝑓𝑡
4𝑄 4 1.385 𝑠 𝑓𝑡
𝑉 7.05 𝑠
𝜋𝐷 3.1416 0.5𝑓𝑡
(6) The same concrete pipe material, diameter, length, and elevations were used in
examples 4-5, 4-10, and 4-15. These examples illustrate the use of Manning’s
equation, the Darcy- Weisbach equation (using Swamee-Jain equation for the friction
factor), and the Hazen- Williams formula in estimating friction loss in pipe. The
values of the flow velocity found using these three methods are listed below.
(i) Manning’s equation: V = 6.46 fps
(ii) Darcy-Weisbach (with Swamee-Jain): V = 6.54 fps
(iii) Hazen-Williams formula: V = 7.05 fps

634.0405 Local Losses in Pipelines

A. Local losses are energy losses due to the presence of appurtenances or changes in the
pipeline such as valves, elbows, curves, reductions, or expansions. The term local losses
indicate that these energy losses are concentrated at a location (rather than distributed along a
pipeline as are friction losses). Local losses in pipelines are calculated using an equation of
the form:

ℎ 𝐾 𝐾 (eq. 4-31)

Where the local loss coefficient K, depends on the nature of the device or pipeline change
producing the loss. Note that local losses are sometimes referred to as minor losses.
B. In addition to losses at appurtenances, local losses occur at pipeline expansions and
contractions. Local losses also occur at entrances and discharge ends of pipe.
C. Discharge loss - Figure 4-10, below, shows the conditions of flow at the discharge end of
a pipeline into a reservoir.
Figure 4-10. Discharge from a pipe into a reservoir.

(1) At section (1), right before the entrance, the pressure head is p1/ω = H, the elevation
of the pipe centerline can be taken as z1 = 0, and the velocity is V1 = V. The curved
line in the reservoir depicts a streamline of flow connecting the pipe outlet to the free

(210-634-H, 1st Ed., Feb 2022)


634-4.14
Title 210 – National Engineering Handbook

surface of the reservoir at section (2). In this section the pressure head is p2/ω = 0, the
elevation is z2 = H, and the local velocity is V2 = 0. Energy head loss between
sections (1) and (2) is a local head loss referred to as a discharge loss, and given by
equation 4-31:
𝑉
ℎ 𝐾
2𝑔
(2) Writing the energy equation (equation 3-20) between points (1) and (2), gives:
𝑝 𝑉 𝑝 𝑉
𝑧 𝑧 ℎ
𝜔 2𝑔 𝜔 2𝑔
(3) Replacing the data for sections (1) and (2) as described above, the equation reduces
to:
𝑉 𝑉
0 𝐻 𝐻 0 0 𝐾
2𝑔 2𝑔
(4) From which it follows that the discharge loss coefficient, Kd = 1.0 (standard value
used in pipe flow analyses).
D. Entrance loss - Three different conditions of entrance from a reservoir into a pipe, and
their local loss coefficients, are depicted in figure 4-11.

Figure 4-11. Entrance loss coefficients for typical pipe entrance shapes

(1) The Pipe Flow tab in the USDA-NRCS Hydraulics Formula software provides
additional values for the entrance coefficient Ke for other specific entrance
conditions. These values are summarized in figure 4-12. Entrance loss coefficients
may also be found in FHWA (2001) and USACE (2008), related to culvert analysis.
(2) Entrance losses are calculated using the expression (equation 4-31):
𝑉 8𝑄
ℎ 𝐾 𝐾
2𝑔 𝜋𝑔𝐷

(210-634-H, 1st Ed., Feb 2022)


634-4.15
Title 210 – National Engineering Handbook

Figure 4-12. Pipe entrance loss coefficients.


Type of Structure and Entrance Design Ke
Pipe, Concrete
Projecting from fill, socket end [groove-end] 0.2
Projecting from fill, square cut end 0.5
Headwall or Headwall and wingwalls with
Socket end of pipe [groove-end] 0.2
Square end 0.5
Rounded (radius = 1/12D) 0.2
Mitered to conform to fill slope 0.7
End section conforming to fill slope 0.5
Pipe or Pipe-Arch, Corrugated Metal
Projecting from fill (no headwall) 0.9
Headwall or Headwall and wingwalls square edge 0.5
Mitered to conform to fill slope 0.7
End section conforming to fill slope 0.5

(3) The USDA-NRCS Hydraulics Formula software can be used to determine pipe flow
calculation including local losses as shown in Example 4-16.
(i) Example 4-16 -Using the Pipe Flow tab in the Hydraulics Formula software,
calculate the discharge Q and flow velocity V for a 18-in-diameter, 1200-ft-long,
corrugated plastic pipe (Manning’s n = 0.015) whose entrance is mitered to
conform to a fill slope (Ke = 0.7). Head on the pipe is 10 ft.
(ii) The solution, illustrated in the figure 4-13 indicates that Q = 8.1 cfs, and V = 4.6
ft/s.

Figure 4-13: Solution for pipe flow with entrance loss using USDA-NRCS
Engineering Field Tools Hydraulics Formula.

(210-634-H, 1st Ed., Feb 2022)


634-4.16
Title 210 – National Engineering Handbook

E. Losses due to pipe fittings


(1) Pipe fittings such as valves, elbows, and bends produce local losses according to
(equation 4-31). The values of selected pipe fittings are shown in figure 4-14.

Figure 4-14: Local loss coefficients for selected pipe fittings.


Pipe fitting K
Globe valve, wide open 10
Alfalfa or stub valves 2.80
Close-return bend 2.20
Gate valve, half open 2.06
Portable hydrants 1.00
Valve opening elbows 1.00
Tee, through side outlet 1.80
90° short-radius elbow 0.90
90° medium-radius elbow 0.75
90° long-radius elbow 0.60
45° elbow 0.42
Gate valve, wide open 0.19
Ball valve, wide open 0.06

For a more complete set of values for local losses, refer to Brater and King (1996) or
Idelchik (1999).
(2) Calculating Pipe flow between two reservoirs including local losses using Manning’s
equation is found in Example 4-17
(3) Example 4-17
(i) The figure 4-15 shows the steady flow between two reservoirs whose free
surfaces have an elevation difference, H. The pipe has length L, diameter D, and
resistance coefficient; the pipe carries a discharge Q and flows with a velocity, V.

Figure 4-15: Schematic of pipe flow between two reservoirs.

(ii) The velocity at the free surfaces (A) and (B) is practically zero, i.e., VA = VB = 0,
and with the reservoirs open to the atmosphere, the gage pressure at those points
are also zero, pA = pB = 0. Finally, the elevations of the free surfaces at points
(A) and (B) can be taken as zA = H and zB = 0 (i.e., the reference level for
elevation is free surface (B)). The energy equation between points (A) and (B),

(210-634-H, 1st Ed., Feb 2022)


634-4.17
Title 210 – National Engineering Handbook

including friction and local losses in the pipe, is written as (see 634.0302
equation 3-19):
𝑝 𝑉 𝑝 𝑉
𝑧 𝑧 ℎ
𝜔 2𝑔 𝜔 2𝑔
(iii) Introducing Manning’s equation (equation 4-8) to estimate pipe friction losses,
the local loss equation (equation 4-31), and the data presented above, the energy
equation is written as:
0 0 0 0 𝑛 𝐿 𝑉
𝐻 2.8755 𝑉 𝐾
𝜔 2𝑔 𝜔 2𝑔 𝐷 2𝑔
Results in:
𝑛 𝐿 ∑𝐾
𝐻 𝑉 2.8755
𝐷 2𝑔
In terms of the flow discharge, this equation is written as:
16𝑄 𝑛 𝐿 ∑𝐾
𝐻 2.8755
𝜋 𝐷 𝐷 2𝑔
(iv) The term ∑K includes the sum of all local loss coefficients in the pipeline. Note
that the Darcy-Weisbach equation or the Hazen-Williams formula is always an
option to use in estimating pipe friction loss.
(v) Using the values L = 550 ft, D = 2 ft, n = 0.012, Q = 35 cfs, with the following
local loss coefficients:
 Entrance: Ke = 0.04 (bell-mouth entrance)
 Elbow 1: Kb1 = 0.90 (short-radius elbow)
 Elbow 2: Kb2 = 0.75 (medium-radius elbow)
 Valve: Kv = 0.19 (gate valve, wide open)
 Discharge: Kd = 1.00 (standard value for discharge coefficient)
(vi) Thus, for this case

𝐾 𝐾 𝐾 𝐾 𝐾 𝐾 0.04 0.90 0.75 0.19 1.0

2.88
(vii) Find the difference in elevations between the reservoirs, H. Using the equation
developed above:
16 35 0.012 550 2.88
𝐻 2.8755 16.77𝑓𝑡
𝜋 2 2 2 32.2
(viii) If the elevation difference between reservoirs is known, the resulting equation
above may be solved for V to determine the velocity in the pipe. Then the
discharge may be determined with the continuity equation, Q = VA.

F. Development of the Pipe Flow equation


(1) The equation from the Pipe Flow tab in the USDA-NRCS Engineering Field Tools
Hydraulics Formula software is shown figure 4-13 and repeated here:

(210-634-H, 1st Ed., Feb 2022)


634-4.18
Title 210 – National Engineering Handbook

𝑄 𝐴 ;𝐾 ; d(in) (eq. 4-32)

(2) To develop this equation:

𝐾 𝐾 𝐾 1

with Ke being the entrance loss coefficient, Kb accounting for devices such as elbows
and valves (referred to as Bend Coefficient in figure 4-13), and Kd = 1 being the
discharge coefficient. Thus, the energy equation presented in example 4-17
(Manning’s equation is used to estimate pipe friction losses) is written as:
𝑛 𝐿 ∑𝐾 𝑉 𝑛 𝐿
𝐻 𝑉 2.8755 2𝑔 2.8755 𝐾 𝐾 1
𝐷 2𝑔 2𝑔 𝐷
(3) Replacing the diameter D (in feet) with d (in inches) and defining:
5087𝑛
𝐾
𝑑
(4) The energy equation becomes:
𝑉 𝑉
𝐻 𝐾 𝐿 𝐾 𝐾 1 1 𝐾 𝐾 𝐾 𝐿
2𝑔 2𝑔
(5) Solving for V, gives the following equation:
2𝑔𝐻
𝑉
1 𝐾 𝐾 𝐾 𝐿
(6) Multiplying this result by the area of the cross-section and substituting Q for VA
gives equation 4-32.
G. Pipe flow calculation including local losses using USDA-NRCS Engineering Field Tools
Hydraulics Formula are shown in Example 4-18
(1) Example 4-18 - Use the Pipe Flow tab in the USDA-NRCS Engineering Field Tools
Hydraulics Formula software to determine the discharge in a pipe with L = 550 ft, D
= 2 ft, n = 0.012, H = 10 ft, with the following local loss coefficients, Ke = 0.04
(bell-mouth entrance), Kb2 = 0.75 (medium-radius elbow), and Kd = 1.00 (standard
value for discharge coefficient).
(2) The solution is shown in the following figure 4-16. The results shown are Q = 28.9
cfs and V =9.2 ft/s. Also, the friction coefficient Kp = 0.0106.

(210-634-H, 1st Ed., Feb 2022)


634-4.19
Title 210 – National Engineering Handbook

Figure 4-16: Solution for pipe flow including all local losses using USDA-
NRCS Engineering Field Tools Hydraulics Formula

634.0406 Pumps in Pipelines

A. Pumps are mechanical devices used to introduce energy into a pipeline system. Pumps
can be used, for example, to lift water from a lower elevation to a higher one, or to overcome
friction losses between reservoirs. The most common types of pumps used in pipelines are
centrifugal pumps. In this section the analysis of pipeline systems with centrifugal pumps is
presented.
B. A pump introduces an energy head hp into a pipeline system. The energy equation
including a pump head is written as:

𝑧 ℎ 𝑧 ℎ ∑ℎ (eq. 4-33)

The normal convention is that energy added to the flow, such as the pump head hp, is placed
on the left-hand side (upstream side) of the equation, while, energy extracted from or lost by
the flow, such as the friction head hf and the sum of local losses ∑hL, is placed on the right-
hand side (downstream side) of the equation.
C. Pump Operational Characteristics
(1) In pipeline system hydraulics, an important operational characteristic of a pump is
the variation of the pump head hp with the discharge Q passing through the pump.
For centrifugal pumps the relationship between hp and Q is given by a quadratic
equation (also called a polynomial equation of the second degree) of the form:
ℎ 𝑎𝑄 𝑏𝑄 𝑐 (eq. 4-34)
(2) The coefficients a, b, and c, in this equation, can be determined by fitting data from
tests performed on a given pump. The equation describing a given pump, or at least a
graph of the relationship between hp and Q, should be available from the pump
manufacturer.

(210-634-H, 1st Ed., Feb 2022)


634-4.20
Title 210 – National Engineering Handbook

(3) Example 4-19 – Pump head and discharge analysis


(i) Tests on a pump produce the following discharge-head data shown in figure 4-17:

Figure 4-17: Discharge versus Head from Pump Tests


Q(cfs) Hp(ft)
0.0 25.3
5.0 22.0
10.0 18.3
15.0 12.0
20.0 4.2

(ii) The pump discharge graph and equation were produced with curve-fitting
software and are shown in figure 4-18.

Figure 4-18: Pump discharge-head graph

Pump discharge characteristics


30
25
20
H(ft)

15
10
H = ‐0.0331Q2 ‐ 0.3811Q + 25.143
5
0
0 5 10 15 20 25

Q(cfs)

(iii) Notice from the graph that the energy head for zero discharge is H = 25.143 ft.
This is the “shut-off head” or maximum head for this pump. If valves are closed
and no flow occurs, the pump is capable of building 25.143 ft of head.
(iv) The maximum pump capacity at free discharge is calculated from the pump
curve equation by setting H = 0 and solving for Q. This gives
0 0.0331𝑄 0.3811𝑄 25.143
The positive solution to this quadratic equation produces the free-discharge, Q =
22.4 cfs, which the pump delivers with H = 0 (no discharge pressure). This
discharge is what the pump would deliver if it were disconnected from any
pipeline and allowed to discharge freely to the atmosphere.
(v) Example 4-20 illustrates the use of manufacturer-provided pump curves in pump
selection.
 The figure 4-19 shows the pump curves provided by a manufacturer for
various pumps whose sizes are specified in the figure.

(210-634-H, 1st Ed., Feb 2022)


634-4.21
Title 210 – National Engineering Handbook

Figure 4-19: Manufacturer-provided Pump Curves and Pump Sizes


OVERALL DIMENSIONS MODEL Motor power Pipe size Dimension “A”

HP KW inches inches mm

SP3005X7AZ ¾ 0.55 1½ 105/8 270

SP3007X10AZ 1 0.75 1½ 11 280

SP3010X15AZ* 1½ 1.10 2 121/8 308

SP3015X20AZ* 2 1.55 2 131/16 332

SP3020X25AZ* 2½ 1.88 2 131/16 332

SP3025X30AZ 3 2.20 2 1413/64 361

* Super II Pump available with dual‐speed motors

 Suppose that the pump is to be installed in a system as that shown in figure 4-


20. The parameters of the system are the following: L = 1000 ft, D = 2 in, e
= 0.000005 ft, g = 32.2 ft/s2, = 1.2010-5ft/s, Ke = 0.5, Kd = 1.0, (i.e., K
= 1.5), and z = 20 ft.

Figure 4-20: Pump-pipeline System for Analysis.

 A numerical spreadsheet application was used to develop the system curve,


which is the capacity and head needed for various operating conditions. The
tabular data for the system curve are shown in figure 4-21:

(210-634-H, 1st Ed., Feb 2022)


634-4.22
Title 210 – National Engineering Handbook

Figure 4-21: System Curve (Tabular)


System hp (ft) Q (gpm)
21 4
30 22
55 45
93 67
142 90
275 135
451 180

 The next step is to plot the system curve values of hp (ft) and Q(gpm) over
the manufacturer’s pump curves, as shown in the figure 4-22:

Figure 4-22: Pump System Curve (Graphical)

 The points of intersection of the system curve with the pump curves are the
operating points. The three system operating points for the pumps using 2-
inch diameter pipe are:
- Pump SP3010X15AZ, Q = 59 gpm, hP = 80 ft
- Pump SP3015X20AZ, Q = 64 gpm, h P = 88 ft
- Pumps SP3020X25AZ and SP3025X30AZ, Q = 69 gpm, h P = 99 ft
Generally, a pump should be selected with an operating point at 80 to 85% of
the shut-off head (maximum head).
D. Pump Power and Efficiency
(1) The hydraulic power developed by a pump represents the amount of energy per unit
time introduced by the pump into the flow. The hydraulic power, Ph, is calculated as:
𝑃 𝛾 𝑄ℎ (eq. 4-35)
where 𝛾 is the specific weight of water. Using the units of the English system,
namely, 𝛾 (lb/ft3), Q (cfs), and hp(ft), the power is given in units of lb∙ft/s. A more

(210-634-H, 1st Ed., Feb 2022)


634-4.23
Title 210 – National Engineering Handbook

commonly used unit of power in the English System is the horsepower (hp) defined
as 1 hp = 550 lbꞏft/s. Thus, in terms of horsepower (Ph) the equation for the hydraulic
power of a pump is:

𝑃 (eq. 4-36)

(2) In units of the International System (SI), namely, 𝛾 (N/m3), Q (m3/s), and hp(m), the
pump power Ph is given in Watts (W). For the SI use equation 4-35 to calculate the
power.
(3) To provide hydraulic power, Ph, the pump P is activated by a motor M as illustrated
in Figure 4-23. The motor can be powered by electricity at connection E.

Figure 4-23: System Schematic showing Motor M, the Electric Supply E, and
Pump P

(4) The motor provides the pump with input power Pm, however, due to energy losses by
friction in the pump shaft (which is dissipated as heat in the environment), and other
losses, not all the motor power Pm is utilized by the pump to produce hydraulic
power. Thus, in general, Ph < Pm, and the ratio between the hydraulic power and the
motor power is a quantity smaller than one (Ph/Pm<1) known as the efficiency of the
pump, ηp:
 1 (eq. 4-37)

(5) The motor itself is provided with a certain amount of electric power Pe; however, due
to losses in the transmission line as well as in the motor mechanism the motor power
Pm is smaller than the electric power, i.e., Pm<Pe. Thus, the efficiency of the motor,
ηm, is a quantity smaller than one (Pm/Pe <1), defined as:
 1 (eq. 4-38)

(6) The combined motor-pump system, in turn, has an efficiency ηm-p defined as
   1 (eq. 4-39)

(7) The various efficiencies, ηi, can be expressed as a percent by multiplying by 100.
(8) Example 4-21 - Calculating Pump power and efficiency.
(i) A pump produces a head hp = 41.67 ft with a discharge Q = 1.24 cfs. The pump
efficiency is 80% (ηp = 0.80) and the motor efficiency is 90% (ηm = 0.90).

(210-634-H, 1st Ed., Feb 2022)


634-4.24
Title 210 – National Engineering Handbook

(ii) Determine: (a) the hydraulic power provided by the pump to the flow, Ph; (b) the
power that the motor needs to provide to the pump, Pm; (c) the electric power
needed to be provided to the motor, Pe; and (d) the efficiency of the motor-pump
system, ηm-p.
 The hydraulic power (in hp) is calculated according to equation 4-36:
𝜔𝑄ℎ 62.4 1.24 41.67
𝑃 5.86ℎ𝑝
550 550
 The power provided by the motor follows from the definition of the pump
efficiency (equation 4-37):
𝑃 5.86
𝑃 7.33ℎ𝑝
 0.80
 The electric power provided to the motor follows from the definition of the
motor efficiency (equation 4-38):
𝑃 7.33
𝑃 8.14ℎ𝑝
 0.90
 The efficiency of the motor-pump system can be calculated with equation 4-
39:
𝜂 𝜂 𝜂 0.80 0.90 0.72 72%
or
𝑃 5.86
 0.72 72%
𝑃 8.14

634.0407 Pipelines and Networks

A. Pipelines of different diameters can be combined into different configurations including


pipes in series, pipes in parallel, pipes converging to a single point, and more complex pipe
networks. These configurations are discussed in the following sections.
B. Pipelines in Series
(1) Pipelines in series consist of segments of pipeline connected one after the other so
that the flow discharge follows a single path. An arrangement of pipelines in series is
presented in figure 4-24.

Figure 4-24: Three Pipelines in Series Connecting Two Reservoirs

N.T.S.

(210-634-H, 1st Ed., Feb 2022)


634-4.25
Title 210 – National Engineering Handbook

(2) Three pipelines of different diameters and lengths are shown carrying water between
reservoirs (A) and (B). The pipelines have lengths L1, L2, and L3, with corresponding
diameters D1, D2, and D3. In addition, the pipelines have different roughness
coefficients. The elevation difference between reservoirs (A) and (B) is given by ∆H.
C. Local head losses due to an expansion or a contraction
(1) The system illustrated in figure 4-24 shows an expansion in diameter between
pipelines 1 and 2 (point X), and a contraction in diameter between pipelines 2 and 3
(point C). Expansions and contraction in pipelines produce local head losses.
(2) For a sudden expansion, the local head losses are calculated as:

ℎ𝑙 1 1 (eq. 4-40)

(3) The subscripts, u and d refer to the upstream and downstream pipelines. Notice that
the head loss in a sudden expansion depends only on the diameters upstream and
downstream of the expansion (Du, Dd). Defining expansion loss coefficients as:

𝐾 1 (eq. 4-41)

or

𝐾 1 (eq. 4-42)

(4) The general form of the local losses equation is written as:

ℎ 𝐾 𝐾 (eq. 4-43)

(5) For a sudden contraction, as the one between pipes 2 and 3 in figure 4-24, the local
head losses are given by the equation:

ℎ 𝐾 (eq. 4-44)

where Kc is the contraction loss coefficient given in figure 4-25, below, and Vd is the
flow velocity in the downstream pipeline, i.e., the one with the smallest diameter. In
figure 4-24, Dd and Du are the diameters of the pipe downstream and upstream of the
contraction. See Brater and King (1996) or Idelchik (1999) for more information on
pipeline contraction losses.

Figure 4-25 Local Head Loss Coefficients for a Sudden Pipe Contraction
Dd /Du Kc Dd /Du Kc
0.1 0.45 0.6 0.28
0.2 0.42 0.7 0.22
0.3 0.39 0.8 0.15
0.4 0.36 0.9 0.06
0.5 0.33 1.0 0.00

(210-634-H, 1st Ed., Feb 2022)


634-4.26
Title 210 – National Engineering Handbook

D. Energy equation for the three-pipeline system


(1) In writing the energy equation between reservoirs (A) and (B) in figure 4-24 friction
head losses in each pipeline will be included, as well as entrance losses from
reservoir (A) into pipeline 1, expansion losses at point (X), contraction losses at point
(C), and discharge losses from pipeline 3 into reservoir (B). For the points (A) and
(B) in the surface of the reservoirs, VA = VB = 0, pA = pB = 0, zA = zB + ∆H. The
expressions for the local losses are the following:
(i) Entrance loss from reservoir (A) to pipe 1:

ℎ 𝐾 ; with Ke=0.5. (See figure 4-11.)

(ii) Expansion from pipe 1 to pipe 2:

ℎ 𝐾 ; with 𝐾 1

(iii) Contraction from pipe 2 to pipe 3:

ℎ 𝐾 ; with Kc from figure 4-25 as a function of the ratio of


D3/D2.
(iv) Discharge loss from pipe 3 into reservoir (B):

ℎ 𝐾 , with Kd=1.0 always.

(2) The friction losses in each pipe may be calculated using either the Darcy-Weisbach
equation, the Manning’s equation, or the Hazen-Williams formula. The total friction
loss is the sum of each pipe’s friction loss.
(3) The energy equation for the system of figure 4-24 is written as:
𝑝 𝑉 𝑝 𝑉
𝑧 𝑧 ℎ ℎ
𝜔 2𝑔 𝜔 2𝑔
(4) From which it follows that:
0 0 0 0 𝑉 𝑉 𝑉 𝑉
𝑧 ∆𝐻 𝑧 ℎ 𝐾 𝐾 𝐾 𝐾
𝜔 2𝑔 𝜔 2𝑔 2𝑔 2𝑔 2𝑔 2𝑔
or,

∆𝐻 ∑ℎ ∑ℎ ∑ℎ 𝐾 𝐾 𝐾 𝐾 (eq. 4-45)

(5) In most analyses of pipelines in series the diameters of the pipelines are known, and
the problem consists in determining the discharge Q for a given available head ∆H, or
vice versa. The three examples of pipes in series presented below include the effect
of local losses.

(i) Example 4-22 – Pipes in series using the Darcy-Weisbach equation. The system
of Figure 4-24 has pipelines lengths, diameters, and roughness values: L1 = 200
ft, L2 = 400 ft, L3 = 150 ft, D1 = 1.00 ft, D2 = 1.50 ft, D3 = 1.00 ft, e1 = 0.0001ft
(concrete), e2 = 0.00004 ft (PVC), and e3 = 0.00025 ft (welded steel). The
kinematic viscosity is ν = 1x10-5 ft2/s. If the discharge Q = 5 cfs, determine the
needed head H. If the available head between the reservoirs is H = 30 ft,
determine the discharge through the pipes.

(210-634-H, 1st Ed., Feb 2022)


634-4.27
Title 210 – National Engineering Handbook

A spreadsheet application gives the results. If the discharge is 5 cfs, the needed
head is coincidentally 5 ft.

If the available head is 30 ft, the discharge, Q = 12.58 cfs.

(ii) Example 4-23 - Pipes in series using the Manning’s equation. Using the same
data as example 4-22, the pipe lengths, diameters, and Manning’s n coefficients
are given by L1 = 200 ft, L2 = 400 ft, L3 = 150 ft, D1 = 1.00 ft, D2 = 1.50 ft, D3 =
1.00 ft, n1 = 0.012(concrete), n2 = 0.01(PVC), and n3 = 0.013(welded steel). If
the discharge Q = 5 cfs, determine the needed head ∆H. If the available head
between the reservoirs is ∆H = 30 ft, determine the discharge through the pipes.
A spreadsheet application gives the results. If the discharge is 5 cfs, the needed
head is 8.12 ft.

If the available head is 30 ft, the discharge, Q = 9.61 cfs.

(iii) Example 4-24 - Pipes in series using the Hazen-Williams formula. Using the
same data as examples 4-22 and 4-23, the pipe lengths, diameters, and Hazen-
Williams coefficients are given by L1 = 200 ft, L2 = 400 ft, L3 = 150 ft, D1 = 1.00
ft, D2 =1.50 ft, D3 = 1.00 ft, CHW1 = 120(concrete), CHW2 = 150(PVC), and
CHW3 = 120(welded steel). If the discharge Q = 5 cfs, determine the needed head
∆H. If the available head between the reservoirs is ∆H = 30 ft, determine the
discharge through the pipes.

(210-634-H, 1st Ed., Feb 2022)


634-4.28
Title 210 – National Engineering Handbook

A spreadsheet application gives the results. If the discharge is 5 cfs, the needed
head is 6.37 ft.

If the available head is 30 ft, the discharge, Q = 11.39 cfs.

(6) Calculations were also made ignoring local losses in examples 4-22, 4-23, and 4-24.
Figure 4-26 shows the effect of neglecting local losses on discharge.

Figure 4-26: Effect of neglecting local losses on an example pipeline system


Friction loss Q (cfs) including Q(cfs) neglecting Percentage
method local losses local losses difference
Darcy-Weisbach 12.58 14.79 17.56%
Manning's 9.61 10.47 8.95%
Hazen-Williams 11.39 13.03 14.40%

(7) The percentage differences in example 4-23 are too large to neglect local loss. A
generally accepted criterion is that local losses should be included in a pipe
design analysis if the local losses exceed 5% of the total head loss.

E. Pipelines in Parallel
(1) Figure 4-27 illustrates reservoirs connected by three different parallel pipelines
conducting water between the reservoirs. An energy equation can be written
separately for each of the pipelines. Including entrance (he), discharge (exit) (hd), and
friction (hf) losses for each pipeline, the corresponding energy equations can be
written as:
∆𝐻 ℎ ℎ ℎ
∆𝐻 ℎ ℎ ℎ
∆𝐻 ℎ ℎ ℎ (eq. 4-46)
Where the subscripts (1), (2), and (3) refer to each of the pipelines as illustrated in
figure 4-27.

(210-634-H, 1st Ed., Feb 2022)


634-4.29
Title 210 – National Engineering Handbook

Figure 4-27: Three pipelines in parallel connecting two reservoirs.

N.T.S.

(2) A simplified analysis in which the local losses are neglected simplifies equation 4-46
to:
∆𝐻 ℎ ℎ ℎ (eq. 4-47)
(3) A typical problem of parallel pipes consists in determining the total discharge Q
delivered from the upstream to the downstream reservoir given the available head ∆H
between the reservoirs. Given ∆H, the available head, the friction loss, hf, is known.
Then using either the Darcy-Weisbach equation (requires spreadsheet application for
efficient calculation), the Manning’s equation, or the Hazen-Williams formula, the
individual discharges Q1, Q2, and Q3 can be obtained. The individual discharges are
summed to obtain the total discharge. The examples using Manning’s equation and
the Hazen-Williams formula presented below neglect local losses.

(i) Example 4-25 - Pipes in parallel using the Manning’s equation. The system of
figure 4-27 has pipelines lengths, diameters, and Manning’s n coefficients: L1 =
200 ft, L2 = 400 ft, L3 = 150 ft, D1 = 1.00 ft, D2 = 1.50 ft, D3 = 1.00 ft, n1 =
0.012, n2 = 0.018, and n3 = 0.010. If the available head is ∆H = 30 ft, determine
the total discharge between the reservoirs.
 The head loss through each pipeline is 30 ft. Manning’s equation (equation 4-
9) is used to calculate the individual discharges:

𝐷 ℎ 1.0 30
𝑄 0.4632 0.4632 14.95𝑐𝑓𝑠
𝑛 𝐿 0.012 200

𝐷 ℎ 1.5 30
𝑄 0.4632 0.4632 20.78𝑐𝑓𝑠
𝑛 𝐿 0.018 400

𝐷 ℎ 1.0 30
𝑄 0.4632 0.4632 20.71𝑐𝑓𝑠
𝑛 𝐿 0.010 150

 The total discharge is Q = 14.95 cfs + 20.78 cfs + 20.71 cfs = 56.44 cfs.
 Discharge calculations may also be made for individual pipes with Pipe
Flow, NRCS Hydraulics Formula program, which uses Manning’s equation
for friction losses. The discharges calculated will be slightly less than those
shown above because Pipe Flow, USDA-NRCS Hydraulics Formula
automatically accounts for a local loss at the pipe exit (Kd = 1.0).
(ii) Example 4-26 - Pipes in parallel using the Hazen-Williams formula. The system
of figure 4-27 has pipelines lengths, diameters, and Hazen-Williams coefficients:

(210-634-H, 1st Ed., Feb 2022)


634-4.30
Title 210 – National Engineering Handbook

L1 = 200 ft, L2 = 400 ft, L3 = 150 ft, D1 = 1.00 ft, D2 = 1.50 ft, D3 = 1.00 ft, CHW1
= 100, CHW2 = 80, and CHW3 = 120. If the available head is ∆H = 30 ft, determine
the total discharge between the reservoirs.
 The head loss through each pipeline is 30ft. Using (equation 4-26), the
individual pipe discharges are calculated as follows:
. .
.
ℎ .
30
𝑄 0.432𝐶 𝐷 0.432 100 1.00
𝐿 200
15.50𝑐𝑓𝑠
. .
.
ℎ .
30
𝑄 0.432𝐶 𝐷 0.432 80 1.50
𝐿 400
24.79𝑐𝑓𝑠
. .
.
ℎ .
30
𝑄 0.432𝐶 𝐷 0.432 120 1.00
𝐿 150
21.74𝑐𝑓𝑠
 The total discharge is Q = 15.50 cfs + 24.79 cfs + 21.74 cfs = 62.03 cfs.
F. Pipelines Converging at a Single Point
(1) Figure 4-28 shows a pipeline system consisting of three pipelines connecting
reservoirs A, B, and C to a single delivery point J (for Junction), where a total
discharge Q is to be delivered. Shown in the figure are also the elevations of the free
surface of the reservoirs, namely, HA, HB, and HC, as well as the piezometric head
elevation of the junction point, HJ.

Figure 4-28: Three pipelines converging to a single delivery point.

N.T.S.

(2) Neglecting local or minor losses, the energy equations for the three pipelines shown
above are written as:
𝐻 𝐻 ℎ
𝐻 𝐻 ℎ
𝐻 𝐻 ℎ (eqs. 4-48)
Where (hf )1, (hf )2, and (hf )3 are the friction losses in pipelines [1], [2], and [3],
respectively.

(210-634-H, 1st Ed., Feb 2022)


634-4.31
Title 210 – National Engineering Handbook

(3) A typical problem for the system illustrated in figure 4-28 consists in determining the
discharge Q given the elevations of the free surfaces in reservoirs A, B, and C (HA,
HB, HC) and that of point J (HJ). The individual pipe discharges are summed to obtain
the total. The examples using Manning’s equation and the Hazen-Williams formula
presented below neglect local losses.
(i) Example 4-27, Converging pipelines using Manning’s equation. The system of
figure 4-28 has pipelines lengths, diameters, and Manning’s n coefficients: L1 =
200 ft, L2 = 400 ft, L3 = 150 ft, D1 = 1.00 ft, D2 = 1.50 ft, D3 = 1.00 ft, n1 =
0.012, n2 = 0.018, and n3 = 0.010. The elevations of interest are HA = 280 ft, HB
= 290 ft, HC = 310 ft, and HJ = 250 ft.
 Determine the total discharge delivered to junction J. The energy losses in
each pipeline are calculated as follows:
(hf )1 = HA – HJ = 280 ft – 250 ft = 30 ft,
(hf )2 = HB – HJ = 290 ft – 250 ft = 40 ft,
(hf )3 = HC – HJ = 310 ft – 250 ft = 60 ft.

 As in example 4-25, Manning’s equation is used to calculate the individual


discharges as follows:

𝐷 ℎ 1.0 30
𝑄 0.4632 0.4632 14.95𝑐𝑓𝑠
𝑛 𝐿 0.012 200

𝐷 ℎ 1.5 40
𝑄 0.4632 0.4632 23.99𝑐𝑓𝑠
𝑛 𝐿 0.018 400

𝐷 ℎ 1.0 60
𝑄 0.4632 0.4632 29.3𝑐𝑓𝑠
𝑛 𝐿 0.010 150

 The total discharge is Q = 14.95 cfs + 23.99 cfs + 29.3 cfs = 68.24 cfs
 As in the previous parallel pipes example 4-25, using Manning’s formula for
friction losses, discharge calculations may also be made for individual pipes
with Pipe Flow, USDA-NRCS Hydraulics Formula program. The discharges
calculated will be slightly less than those shown above because Pipe Flow,
USDA-NRCS Hydraulics Formula automatically accounts for a local loss at
the pipe exit (Kd = 1.0).
(ii) Example 4-28, Converging pipelines using Hazen-William’s formula. The system
of figure 4-28 has pipelines lengths, diameters, and Hazen-William’s
coefficients: L1 = 200 ft, L2 = 400 ft, L3 = 150 ft, D1 = 1.00 ft, D2 = 1.50 ft, D3 =
1.00 ft, CHW1 = 100, CHW2 = 80, and CHW3 = 120. The elevations of interest
are HA = 280 ft, HB = 290 ft, HC = 310 ft, and HJ = 250 ft. Determine the total
discharge delivered to junction J.
 The energy losses in each pipeline are calculated as follows:
(hf)1 = HA – HJ = 280 ft – 250 ft = 30 ft,
(hf)2 = HB – HJ = 290 ft – 250 ft = 40 ft,
(hf)3 = HC – HJ = 310 ft – 250 ft = 60 ft.

(210-634-H, 1st Ed., Feb 2022)


634-4.32
Title 210 – National Engineering Handbook

 As in example 4-25, the Hazen-Williams formula is used to calculate the


individual pipe discharges as follows:
. .
.
ℎ .
30
𝑄 0.432𝐶 𝐷 0.432 100 1.00
𝐿 200
15.50𝑐𝑓𝑠
. .
.
ℎ .
40
𝑄 0.432𝐶 𝐷 0.432 80 1.50
𝐿 400
28.95𝑐𝑓𝑠
. .
.
ℎ .
60
𝑄 0.432𝐶 𝐷 0.432 120 1.00
𝐿 150
31.60𝑐𝑓𝑠
 The total discharge is Q = 15.50 cfs + 28.95 cfs + 31.60 cfs = 76.05 cfs.
G. Pipeline Networks
(1) Pipe networks are utilized to supply water for urban and rural domestic and industrial
uses. Networks are also used by irrigation districts and occasionally in livestock
watering systems. A pipe network is simply a collection of pipes connected in a given
geometric pattern and having at least one supply point and one delivery or
consumption point. The points where individual pipelines join each other are known
as junctions or nodes. Pipe networks are schematized as geometric constructs of lines
representing the component pipes.
(2) The main purpose of analyzing a given pipe network is to determine the discharge in
each pipe and the piezometric head at each node. These results can be used then to
verify that certain design guidelines, such as minimum flow velocities and minimum
junction pressures, are satisfied. The analysis of pipe networks requires the
simultaneous solution of several mathematical equations that represent equations of
continuity at the nodes, and equations of energy around closed loops and/or pseudo-
loops.
(3) Analyses of simple pipe networks may be performed with a spreadsheet application,
but spreadsheet use would become very tedious as the number of pipelines and loops
increase. There are several publicly available and commercial software that can be
used for network solution in a more efficient manner, such as the U.S. Environmental
Protection Agency’s EPANET software.

634.0407 EPANET

A. A software program available for performing pipeline hydraulic calculations is called


EPANET. Developed and supported by the US Environmental Protection Agency, this
program can be used for full pipe design. It is available online and an approved CCE version
can be loaded on NRCS computers (at this time it must be installed by an IT specialist). It is
free to use, and training can be self-based through tutorials or the web site.
B. This program uses pipes, junctions, pumps, valves, storage tanks and reservoirs as the
basis for design. It has the capability of carry out simulations over defined time frames
computing flow, velocity, source, chemical tracking and water age. The pipe system can be
simply a single pipeline or more complex using multiple pumps, water sources, networks and
loops. Units can be English or metric as well as the results. Maps (as either *.bmp, *.emf, or

(210-634-H, 1st Ed., Feb 2022)


634-4.33
Title 210 – National Engineering Handbook

*.wmf files) can be used as backdrops for lay out, which can be used to scale the system or
enter points and define the layout as the system is developed. There is no limit to size of the
network being analyzed and calculations can be made using either Hazen-Williams, Darcy-
Weisbach or Chezy-Mannings to compute friction.
C. Valves, pumps and junctions (or turnouts) can be set to turn on or off at different times of
the day. Pumps can be entered with one or more flow/head points on the curve or the pump
characteristics can be estimated using a one-point curve built by EPANET. Modeling of the
flow, demands, quality or age can be estimated and colorized and seen as flow changes.
There is a tutorial available that has step by step instructions to build a few different types of
scenarios and explanations on how to use them. This program is very easy to use and quick to
enter data to get hydraulic flow information.
(1) Example 4-29 Use of EPANET for a single pipeline. For a pumped pipeline from a
water source at point 1 (reservoir) to 3 turnouts across a field using EPANET in the
figure 4-29.

Figure 4-29 Hydraulic example for EPANET

(i) After opening the software, it is easy to set up the initial defaults. From the menu
choose the View, Options and select the notation, flow arrows, symbols to use.
Under Notation checking the Display Node Values, Display Link ID’s and
Display Link Values turns on the values to the screen for the user to see. For
more complex systems with loops or multiple pumps or inlets, the flow arrows
are good to see how the system is operating. All these settings can be selected at
any time during the program operation.
(ii) Next, from the Project menu, choose Defaults, and under the Properties and the
Hydraulics tabs to choose the hydraulic calculation method, the typical pipe size,
length, and friction loss value. It is possible to have the software scale the
dimensions based on an inserted background photo or you can adjust the pipe
length at any time under the properties list for each pipe. You may also edit the
view of the layout by changing the nodes under the view. This makes it easier to
see water source, direction, which valves are open or closed and where possible
design concerns are located.
(iii) For this example, enter a reservoir as a water source (Pt 1) and nodes for each of
the above points (Pt 2-7) into EPANET. The screen shot of what this could look
like is shown above with flows in gpm at each of the turnout nodes (Pts. 5, 6 and
7). After entering the reservoir and nodes for the system, select the pump and
connect it to the reservoir and to point 2. Then connect the pipes from node to
node. Edit all the information in the properties table for each node elevation and
pipe diameter, friction and length. All elevations are as shown. Use 500 ft. of 10”
diameter pipe from the pump outlet at point 2 to the first turnout at point 5. From
point 5 to the next turnout at point 6 use 200 ft. of 8” diameter pipe and from the

(210-634-H, 1st Ed., Feb 2022)


634-4.34
Title 210 – National Engineering Handbook

second turnout (Pt 6) use 300 ft. of 6” diameter pipe the end (Pt 7), calculate the
pump head needed in feet to get 45 psi of pressure at the third turnout (Pt 7).
(iv) With the pump located between the reservoir and point 2 it is connected between
these two points and just ties into the line and does not take up length in the
system. To get the system to run, at least a one-point pump curve (Q in gpm and
H in ft) must be entered. This is done in two parts. First a pump curve needs to be
entered under curves on the right under Data on the Browser menu. Next enter
the pump curve number into the pump properties table. For this example, we
want to produce a pressure of (45 psi) at the end of the line. To get this pressure,
we can do trial and error for the pump. For the first estimate try a flow of 650
gpm and head 150 ft entered. The figure 4-30 is of the pump 1 Curve editor.
(v) Edits can be made by either selecting an item or by selecting from the Browser,
Data menu on the right. After entering the correct friction coefficient, friction
loss equation to use, pipe size and lengths, select the run button (lightning bolt) to
see if it works or if there is an error. Errors usually come from pipes not
connected to nodes or junctions, pump curves not entered correctly or a possible
hydraulic issue. Support and troubleshooting can be found by using the support
contained in the software.
(vi) The pump flow rate may show up as negative since it is losing that amount of
water, as will the reservoir.

Figure 4-30: Pump Curve editor screen shot. Only one point is entered for
this example and the curve is calculated from the program.

(vii) Velocity and demand graphics are included for the example below, which can
be turned on from the Browser, Map menu. This color codes the EPANET results
for each section of pipe and locations of turnouts with flow amounts.
(viii) Figure 4-31 shows an EPANET screen shot with flow direction and velocity
shown next to the direction arrows, turnout flows in red in gpm, plus reservoir
location and pipes. The pump is located between points 1 and 2.

(210-634-H, 1st Ed., Feb 2022)


634-4.35
Title 210 – National Engineering Handbook

Figure 4-31: Screen shot of EPANET output

(ix) To have 45 psi (44.9 psi in circle below) at the end of the line (Point 7) the pump
would need to have an outlet head of 150 ft. (62.83 psi) as shown at Junction 2 in
figure 4-28.
(2) The main disadvantage of EPANET is the lack of reports that are user friendly to
NRCS. Reviewing the file in EPANET is the best way to see what is going on.
Reports do not show all the details that use common NRCS nomenclature (see fig 4-
32 and 4-33).

Figure 4-32: Output table of Junction information for EPANET example

Figure 4-33: Output table of links (Pipelines) for EPANET example.

(210-634-H, 1st Ed., Feb 2022)


634-4.36
Title 210 – National Engineering Handbook

(3) What they do show is the operating characteristics of the system and how it will
perform. The reports can be used to build a design and include the appropriate data.
D. Friction, velocities, pressure, flows can be found in the EPANET program for each node
and line. This is useful and can be transferred to other programs for graphing or
documentation. Printing reports and explaining system hydraulics by the user may be the best
way to show results. Another option with the program is to inject chemicals (chemigation and
fertigation) and look at aging and distribution with varying flow parameters. Concentrations,
timing and direction can be quickly shown in real time system operation.
E. EPANET can be used with multiple pumps, loops and varying flow requirements. It is
easier to use than most spreadsheets, more flexible and files that can be transferred to others
who can easily view and use.

634.0408 WaterCAD

A. Bentley WaterCAD is a commercial software program written to do hydraulic


calculations in piping systems. This tool has been available through a limited number of
licenses to NRCS staff, or it can be purchased (see disclaimer page 634-4.i.)
B. This program is more robust than EPANET and provides added flexibility in hydraulic
analysis. There are examples and lessons on how to use the program as a part of the software
package. With more ability in a program requires more challenges for using. The following
example is given to provide a bit of insight on its use.
(1) Example 4-30 Use of WaterCAD for a single pipeline. In example 4-29, inputs from
the EPANET section will be used to show the WaterCAD program. The first thing
after opening the tool is to set up the system parameters with Tools, Options that are
appropriate. The Drawing tab in the Options table is also a good item to check to set
the scale and text alignment. Lastly the Labeling tab in this menu has the names of
the different codes for each of the items that can be inserted such as pipes, valves,
and pumps.
(2) One option to make the project easier to follow is to insert a background image. This
can be from many different formats including jpeg, dxf, shp, jif, or sid. There can
also be many layers that can be turned on and off as needed. Multiple images can be
placed side by side to sequence a series of views, such as for a long pipeline.
Switching images off and on can help in layout and in building a longer plan view of
a system into a complete file. Look under the View, Background Layer to start a new
file and add to your background drawing. The scale of the background can be
changed by adjusting the XY coordinates of the image. Be sure and adjust the image
the same in both dimensions. It can be moved by adjusting the 0:0 location.
(3) Next look at the inputs for the example. In this case the drawing from the EPANET
example above will be used as the layout. Building this data into WaterCAD is like
inputting into EPANET. Using the different symbols on the left of the screen can
guide you to the type of item to select. First start with the reservoir , which is a
quarter of the way down from the top of the screen. After clicking on it, place it on
the screen where it belongs. Every time you left click on the mouse a reservoir will
be placed until you either choose a new item or select the edit arrow at the top of the
list. To select and edit an item after it is on the screen, left click and it will turn red
and you can delete, edit or move the item.

(210-634-H, 1st Ed., Feb 2022)


634-4.37
Title 210 – National Engineering Handbook

(4) Place a pump next to the reservoir and connect it to the reservoir with a pipe. After
that either enter junctions or pipes. Using the pipe button, will include junctions, if
needed. For example, when entering a pipe to connect the reservoir to the pump a
junction will be shown. By clicking on the reservoir, the line will show a junction
which can be ignored and clicking on the pump will link the pipe. This way the
suction side of the pump can be any length and size, which is not as easy to do in
EPANET.
(5) While still in the pipe mode and connected to the pump, move to the right and for
every mouse click a pipe and junction will be placed. This can continue for as many
lengths as necessary. The length will be scaled based on the distance you move.
Having a background image will provide a good basis and scale for the layout. An
example of a junction could be a change in grade, T in the pipe, outlet for a sprinkler,
valve or turnout.
(6) After the example has been entered and it looks like the EPANET example, it is
necessary to edit the information to the actual conditions. To do this simply click on
the item you want to edit, and a properties screen will open. This is where pipe,
junctions, reservoirs, pumps can be edited. You can move across the system and
choose items in whatever order you want. Pipe type, size, roughness factors and
length may be entered. The junctions may also be edited by clicking on the points
and changing data in the properties table. Information that needs changing includes
elevation, flow out of a point, or valve information.
(7) This program is much more robust than described here, this is just a quick overview
to introduce the subject.

634.0409 Appurtenances in Pipelines and Networks

A. In this section many appurtenances commonly used in pipelines and pipe networks are
described.
B. Air Vacuum and Release Valves
(1) Entry and entrapment of air in a pipeline during filling and operation of a pipeline
can cause development of air pockets. Air pockets tend to restrict and reduce flow or
increase pumping costs in pipelines. Air pockets that build-up in pumped pipelines
can reduce flows by 5 to 15%. In low-head, gravity-driven flow, air pocket build-ups
can reduce flow up to 50%. On the other hand, a lack of air entry can cause pipe
collapse if a vacuum develops behind flowing water as a pipeline empties. Air valves
are of four types:
(2) Air/vacuum relief valves, also known as kinetic air valves, large orifice air valves,
vacuum breakers, low-pressure air valves, and air relief (not release) valves. Large
volumes of air are discharged before a pipeline is pressurized, especially at pipe
filling. Large quantities of air are admitted when the pipe drains and at the
appearance of water column separation.
(3) Continuous acting air and vacuum valves, also known as double orifice air valves, or
combination air valves, fill the functions of the air/vacuum relief valves and air
release valves, admitting and releasing large quantities of air when needed, and
releasing air continuously when the lines are pressurized (figure 4-34).

(210-634-H, 1st Ed., Feb 2022)


634-4.38
Title 210 – National Engineering Handbook

Figure 4-34: Continuous acting air valve

(4) Air release valves are also known as automatic air valves, small orifice air valves,
continuous acting air vents, and pressure air valves. These vents continue to
discharge air, usually in smaller quantities, after the air vacuum valves close, as the
line is pressurized.
(5) Vacuum relief valves are large orifice valves for vacuum relief only. These valves
allow air to enter the pipeline when draining.
(6) Guidelines for placement of air valves:
(i) Pipelines in general (figure 4-35) - It is important to establish a smooth
pipeline grade and not follow the terrain or an excessive number of air valves
will be needed. The designer must balance the cost of air valve locations with the
cost of additional excavation. The high points and grade changes that are less
than 1 pipe diameter are typically ignored because the flow will flush
accumulated air downstream. Air valves should be placed at the following
locations:
 Place continuous-acting air and vacuum valves at all high points. The large
orifice is necessary to discharge air at pipe fill-up, and for air intake during
pipe drainage or when the pipe bursts. The small orifice releases air that
continually collects at the high points.
 On long horizontal runs air release valve should be placed at 1250-ft to 2500-
ft (380 m to 760 m) intervals on pipelines. Best practice is to alternate air
release and combination air valves
 Long descents: combination air valve at 1250-ft to 2500-ft (380 m to 760 m)
intervals.
 Long ascents: air/vacuum valve at 1250-ft to 2500-ft (380-m to 760-m)
intervals.
 Decrease in an up slope: air/vacuum valve.
 Increase in a down slope: combination air valve.
 Locate an air release valve at the end of the line

(210-634-H, 1st Ed., Feb 2022)


634-4.39
Title 210 – National Engineering Handbook

Figure 4-35 Air valve placement a long a pipeline

(ii) Locate an air-and-vacuum or vacuum valves to allow air to enter behind the
water as a pipeline is emptied. Vacuum valves are generally not needed on
pipelines of less than 3-inch diameter for protection from collapse; however, they
may be needed to insure complete drainage of the line. On the other hand,
vacuum valves are important on low head plastic irrigation pipe and large
diameter steel pipe (>24”) to prevent collapse of the line.
(iii) Air is usually present in pumps and in suction pipes. When pumps are started and
during operation, air will continue to enter through the pump, fittings, and related
accessories. A combination air/vacuum relief and continuous acting air valve
should be placed directly after the pump and before the pump check valve (figure
4-36).
Figure 4-36 Air valve placement for pumps

(iv) For isolating valves, a continuous acting air valve should be placed upstream of
the valve to release air pressure when the valve is closed and protect from air-
bubble induced cavitation. It also provides slow end-of-pipe air release for air
trapped by the closed valve when filling the pipe (figure 4-37).

(210-634-H, 1st Ed., Feb 2022)


634-4.40
Title 210 – National Engineering Handbook

(v) Downstream of an isolation valve, a continuous acting air/vacuum relief is needed


to prevent collapse during draining of the pipe and to aid in filling operation
when the valve is open (figure 4-37).
Figure 4-37 air valve placement for isolation valves

(vi) Air valve placement at flow meters or metering valves


 One of the most important places to locate air valve in irrigation systems is
before water meters and automatic metering valves. When air rushes through
a water meter or the metering component of an automatic metering valve, it
spins the velocity measuring spinning accelerator with great speed,
registering very high volumes of flow. The metering instruments cannot
distinguish between air and water. The crops may only receive a small
portion of the water that the meters registered and supposedly delivered.
Thus, creating severe, often irreversible damage to crops.
 When water lubricates and cools the spinning parts of flow meter. Air
provides little lubrication and the spinning part tend to heat. If the spinning
element is made of plastic, it may even melt. Placing an air vent before a
metering device protects it from this phenomenon.
(vii) Filter heads
 Combination Air/Vacuum Relief and Continuous Acting Air valves should
be installed at the inflow side of filter heads to prevent air from entering the
filters
 On top of the filters to release air from within and to enable draining and
backwashing.
 At the outflow side to release remaining air and to prevent vacuum
conditions and suction of filter media out of the filter.

(210-634-H, 1st Ed., Feb 2022)


634-4.41
Title 210 – National Engineering Handbook

Figure 4-38 Air valve placement on filter stations

(viii) Vacuum protection in microirrigation especially subsurface drip irrigation, is


essential, even at very low negative pressure, back siphoning of dirt into the
emitters, and causing other damage to pipe and accessories. Air intake should be
equal to maximum drainage flow rate (maximum slope). To prevent suction even
at low negative pressures, air intake should be determined at low negative
pressure, like 1.45 psi (0.1 atmosphere).

Figure 4-39 Air valve placement on a microirrigation manifold

(ix) The need for air valves should also be evaluated in these other common areas of
air accumulation
 Road crossings where larger diameter pipes dip down and then rise up again
in sharp slopes. If the crossing is wide and/or deep, air valves should be
installed on the top elbows at both sides of the crossing. If the crossing is
narrow and shallow, placing one combination air vent on the upstream top
elbow may be sufficient.
 Areas of pressure change such pressure reducing valves or sudden
enlargements can cause turbulence and allow air to be released from the
water to the pipe. A combination Air/Vacuum relief and continuous acting
air valve should be installed very close and downstream from the accessory
or disturbance.
 Control heads and other areas of disturbance or air accumulation.

(210-634-H, 1st Ed., Feb 2022)


634-4.42
Title 210 – National Engineering Handbook

(x) Avoid oversizing air release valves to lessen the possibility of water hammer. See
section 634.0410 for more information on water hammer
(7) Valve sizes may be determined for a pipeline by calculating: (1) flow for both filling
and emptying the pipeline, (2) pipe collapse pressure, and (3) valve intake or
discharge at a prescribed pressure across the valve as illustrated in example 4-31.

(i) Example 4-31 - Air valve sizing. Given: 18” diameter, 80 psi, SDR 51, PVC pipe,
on a slope of 0.05 ft/ft. Note that SDR is the pipe outside diameter divided by the
wall thickness.
 Determine flow for draining the line using the Hazen-Williams formula
(equation 4-26). CHW for PVC pipe is 150:
. .
𝑄 0.432 150 1.5 0.05 37.34𝑐𝑓𝑠
 Air vent valves must re-enter air at the same rate as the calculated water flow
of 37.34 cfs.
 Determine pipe collapse pressure using equation for PVC pipe given in
ASTM in this case DD2241:
2𝐸 1
𝑃
1 𝑣 SDR SDR 1
where: Pc = collapse pressure of PVC pipe (psi)
E = Young’s modulus of elasticity (400,000 psi)
ν = Poisson’s ratio (0.33)
SDR= Standard Dimension Ratio (do/t)
t = wall thickness (in.)
do= outside diameter of pipe (in.)
2 400,000 1
𝑃 7 𝑝𝑠𝑖
1 0.33 51 51 1
 Use the collapse pressure determined above or 5 psi, whichever is lower
(standard design practice).
 Determine orifice diameter, using the orifice flow equation (equation 4-49):
𝑄 𝐶 𝐴 2𝑔ℎ (eq. 4-49)
where: Cd = flow coefficient = 0.6
A = orifice area
ωair = specific weight of air =0.0764 lbs/ft3
Pc = (5 lbs/in2) (144 in2/ ft2) = 720 psf
h = Pc/ωair
 Rearranging the orifice formula,
37.34
𝐴 0.079𝑓𝑡
0.6 2 32.2 720⁄0.0764
𝑖𝑛
𝐴 0.08𝑓𝑡 144 11.37𝑖𝑛
𝑓𝑡

4𝐴 4 11.37
𝐷 3.80 𝑖𝑛𝑐ℎ𝑒𝑠, 𝑜𝑟𝑖𝑓𝑖𝑐𝑒 𝑑𝑖𝑎𝑚𝑒𝑡𝑒𝑟
𝜋 𝜋

(210-634-H, 1st Ed., Feb 2022)


634-4.43
Title 210 – National Engineering Handbook

 Determine the size of orifice required when filling the pipe. When filling a
line, good practice is not to exceed 1 fps (note: not all systems provide that
much control of filling velocity, see next part of this example). Using a
filling velocity of 1 fps and the 18inch diameter pipe:
𝑄 𝐴𝑉 𝜋⁄4 1.5ft 1 𝑓𝑡⁄𝑠 1.76𝑐𝑓𝑠
Therefore, air must be discharged from the line at a rate of 1.76 cfs

 Size the orifice for 2 psi pressure across the valve (standard design practice).
Using the orifice formula, as before:
𝑃 2𝑙𝑏𝑠⁄𝑖𝑛 144𝑖𝑛 ⁄𝑓𝑡 288 𝑝𝑠𝑓
1.76
𝐴 0.006𝑓𝑡
0.6 2 32.2 288⁄0.0764
0.004𝑓𝑡 ∗ 144 𝑖𝑛 ⁄𝑓𝑡 0.85 𝑖𝑛
.
𝐷 1.04 𝑖𝑛 orifice diameter.

 Alternative determination of orifice size required when filling the pipe. When
filling a line, with an electric powered pumping plant not equipped with
special controllers or valves, it may not be possible to limit filling velocity to
1 fps. The electric motor is either on or off which might produce filling
velocities of up to 5 fps. Using a filling velocity of 5 fps:
𝑄 𝐴𝑉 𝜋⁄4 1.5 5 5.89𝑐𝑓𝑠
Therefore, air must be discharged from the line at a rate of 5.89 cfs
 Size the orifice for 2 psi pressure across the valve (standard design practice).
Using the orifice formula, as before:
𝑃 2 𝑙𝑏𝑠⁄𝑖𝑛 144 𝑖𝑛 ⁄𝑓𝑡 288𝑝𝑠𝑓
5.89
𝐴 0.02𝑓𝑡
0.6 2 32.2 288⁄0.0764
0.0 𝑓𝑡 144 𝑖𝑛 ⁄𝑓𝑡 2.87𝑖𝑛
.
𝐷 1.91𝑖𝑛 orifice diameter

(ii) Comparing orifice sizes of 3.80 in. diameter for vacuum relief and 1.04 or 1.01
in. diameter for air release shows that the 3.80 in. diameter orifice should be used
if an air- and-vacuum valve, or continuous-acting air-and-vacuum valve, is to be
used. If individual air release and vacuum relief valves are used, then they each
may be of the size determined.
(8) Additional information on the operation and settings for air valves is available from
the valve manufacturers.
C. Air Vents
(1) Air vents act similarly to air vacuum or air release valves; however, they allow for
the release of large volumes of air near pumps or water intakes, pressure boxes, and
check valves. Air vents are often used in low-head pipelines. The figure below shows
a simple air vent consisting of a cylindrical chamber of diameter Dc located at a
distance L from a water intake where air entrainment occurs. The height of the air

(210-634-H, 1st Ed., Feb 2022)


634-4.44
Title 210 – National Engineering Handbook

vent chamber should be at least one-half the diameter of the pipe (D/2), and the
chamber’s area should be at least half of the mainline pipe cross sectional area (fig 4-
40). For a cylindrical chamber, the minimum diameter is, therefore,

𝐷 (eq. 4-50)

Figure 4-40: Location and dimensions of air vent chamber near a water intake.

(2) The minimum length L, measured from the pipe entrance is given by, Roberson, et.al.
(1956) Using units of the International System (SI), the corresponding equation is:
𝐿 5.77 𝑉 𝐷 (eq. 4-52)
where V is in m/s, D in m, and L in m.
(3) Using units of the English System (ES), the corresponding equation is (4) Example 4-
32 Air vent chamber sizing. A 12-in diameter pipeline is to carry a discharge Q = 2.5
cfs. Determine the minimum diameter, height, and location of an air vent chamber
from a water intake where air entrainment occurs.
(i) The pipe diameter is D = 1 ft, and the flow velocity can be calculated from
634.0301equation 3-04:
4𝑄 4 2.5𝑐𝑓𝑠
𝑉 3.18𝑓𝑝𝑠
𝜋𝐷 3.1416 1𝑓𝑡
(ii) The chamber should have a height of D/2 = 0.5 ft = 6 in, and a minimum diameter
of:

𝐷 1𝑓𝑡
𝐷 0.707𝑓𝑡
2 2
(iii) The distance from the water intake to the air vent should be at least:
𝐿 1.76 𝑉 𝐷 1.76 3.18 1 5.6𝑓𝑡
D. Pressure Control Valves
(1) Pressure relief valves (or pressure safety valves) are used in pipeline systems to
protect against excessive pressure. They are designed to open and discharge small
amounts of water at a preset pressure limit. Excess pressure opens the valve and some
water is released (fig 4-41).

(210-634-H, 1st Ed., Feb 2022)


634-4.45
Title 210 – National Engineering Handbook

Figure 4-41: Spring operated pressure relief valve

(i) Example 4-33 – Pressure relief valve selection. Given: 8” diameter, 80psi, SDR
51, plastic irrigation pipe. Pipe inside diameter is 7.84in = 0.65ft. Design pipeline
velocity is 5fps.
(ii) Select a pressure relief valve to protect against possible surge pressure at the end
of the pipeline. The valve is set to relieve pressure at 80psi, which is the pipe
pressure rating. Size the pressure relief valve to pass the full flow at 150% of the
pipe pressure rating:
(iii) First, determining the pipeline flow:
𝑄 𝐴𝑉 𝜋⁄4 0.65 5 1.66𝑐𝑓𝑠
(iv) Determining minimum orifice diameter, using orifice equation (equation 4-49):
𝑄 𝐶 𝐴 2𝑔𝐻
where: Cd = flow coefficient = 0.6
A = orifice area (ft2)
P = (150%/100) (80 lbs/in2) (144 in2/ ft2) = 17280 psf
ωw = specific weight of water = 62.4 lbs/ft3
h = P /ωw
(v) Rearranging the orifice formula,
𝑄 1.66
𝐴 0.0207𝑓𝑡
𝐶 2𝑔ℎ 0.6 2 32.2 17280⁄62.4
0.0207𝑓𝑡 144 𝑖𝑛 ⁄𝑓𝑡 2.98𝑖𝑛
𝐷 4𝐴⁄𝜋 4 2.98⁄𝜋 1.95 𝑖𝑛, 𝑜𝑟𝑖𝑓𝑖𝑐𝑒 𝑑𝑖𝑎𝑚𝑒𝑡𝑒𝑟
(vi) Review of manufacturer’s literature indicate that pressure relief valves with 2-in
and larger orifices are available. If, for safety concerns, the designer wishes to
reduce the valve escape velocity, larger-size valves are available or multiple
valves may be used. Find the dimensions of the valve opening using the
continuity equation:

(210-634-H, 1st Ed., Feb 2022)


634-4.46
Title 210 – National Engineering Handbook

(2) Pressure reducing valves are used to reduce downstream pressure to protect and
assure proper function of certain components. They maintain a constant downstream
pressure regardless of the upstream pressure. The reducing valves are usually
hydraulic controlled and are adjustable over a wide range of pressures and flows
(figure 4-42).

Figure 4-42 A hydraulic control globe type pressure reducing valve on an irrigation mainline

(3) Similar to pressure reducing valve a pressure regulating valve (PRV) are generally
smaller scale spring operated devices used for regulating the pressure in sprinkler
(figure 4-43) or microirrigation systems.

Figure 4-43 Pressure regulating valve on a center pivot

(i) Example 4-34 – Pressure reducing valve selection. Given: 3” diameter, SDR 41,
PVC pipe. Pipe inside diameter is 3.33in = 0.278ft. The pipe is a 500ft sprinkler
lateral, laid on a 2.5% uphill slope. The lateral’s design velocity is 5fps. A
minimum of 30-psi pressure is needed at the last (end-of-line) sprinkler.
(ii) Select a pressure reducing valve to maintain a maximum of 20% pressure
variation at the sprinklers located along the lateral.
(iii) First, determine the pipeline flow:
𝑄 𝐴𝑉 𝜋⁄4 0.278 5 0.302𝑐𝑓𝑠

(210-634-H, 1st Ed., Feb 2022)


634-4.47
Title 210 – National Engineering Handbook

(iv) Determine friction losses for the lateral using the Hazen-Williams formula
(equation 4-28). CHW for the PVC pipe lateral is 130.
. .
𝑄 𝐿 0.303 500
ℎ 4.732 .
4.732 .
𝐶 𝐷 130 0.278
16.27𝑓𝑡
The friction loss of 16.27 ft is equivalent to (16.27ft/2.31 lb/in2) =7.0 psi.
(v) Determine minimum lateral pressure needed upstream of the first sprinkler. This
pressure is the sum of the pressure needed at the last sprinkler plus the lateral
elevation difference plus the friction losses (local losses neglected).
𝑓𝑡
𝐸𝑙𝑒𝑣. 𝐷𝑖𝑓𝑓. 0.025 𝑓𝑡 500𝑓𝑡 12.5𝑓𝑡, equivalent to 5.4 psi

𝑃 30 5.4 7 42.4𝑝𝑠𝑖
(v) At the first sprinkler, the pressure needs to be reduced from about 43 psi to
between 30 - 36 psi (1.2 x 30psi). Review of manufacturer’s literature indicates
that a 2-in pressure reducing valve is available to meet the need for reduced
pressure. To maintain a maximum of 20% sprinkler pressure variation and 30 psi
at lateral end, pressure reducing valves are needed at about the first half of the
lateral’s sprinklers due to lower elevation and less friction loss.
(4) Pressure-sustaining valves consist of the basic valves and a three-way pressure-
sustaining pilot (figure 4-44). Pressure is sustained at the upstream side of the valve
to a preset level, while the valve outlet drains excessive pressure to maintain the
preset inlet pressure. Pressure is maintained constant, regardless of upstream
fluctuating pressure and flow rate. Pressure sustaining valves are used on hilly terrain
to maintain pressure in elevated areas and many other applications where sustained
pressure is necessary.

Figure 4-44 Pressure sustaining valve on a microirrigation filtering station

(5) Details on the operation of pressure relief, reducing, and sustaining valves are
available from the valve manufacturers. The pipeline system pressure range and
capacity are parameters needed for valve selection.

(210-634-H, 1st Ed., Feb 2022)


634-4.48
Title 210 – National Engineering Handbook

E. Surge/Air Chambers
(1) Surge or air chambers are vertical chambers, typically of cylindrical cross-section
and open to the atmosphere, attached to pipelines to allow pressure relief during
hydraulic transients. They perform a similar role as pressure relief valves, except that
instead of releasing water flow to another pipeline, they allow the water surface level
in the chamber to increase in response to a local pressure increase. During a hydraulic
transient (unsteady flow) the water surface in the chamber will oscillate until a steady
state is recovered. Figure 4-45 shows a schematic of a surge chamber in a pump-
pipeline system. The surge chamber in this case would help dissipate a pressure wave
that may result after a sudden failure of the pump. The system also includes a check
valve that shuts off flow as the flow reverses.

Figure 4-45: Schematic of pump-pipeline system protected with a surge


chamber.

(2) For the situation illustrated in the figure 4-40, if the cross-sectional area of the surge
chamber is Ac and that of the pipeline is A, the maximum height S of the water
surface in the chamber above its steady-state level (i.e., the amplitude of the
oscillation) is calculated as:

𝑆 (eq. 4-53)

where Q is the steady-state discharge, L is the length of pipeline between the


upstream reservoir and the check valve, and g is the acceleration of gravity (= 9.81
m/s2 = 32.2 ft/s2).
(3) The time in seconds required for the water level in the chamber to first reach its
maximum height (one quarter of the period of oscillation) is calculated as, Tullis
(1989):

𝑇 (eq. 4-54)

(210-634-H, 1st Ed., Feb 2022)


634-4.49
Title 210 – National Engineering Handbook

(4) The water in the chamber will oscillate until friction in the pipe, and in the chamber,
slows down the oscillation and produces a complete flow shut off in the pipe
upstream of the check valve. Notice that in the absence of a check valve, flow
reversal will occur through the pump which could affect its operation after starting
up.
(5) Example 4-35 – Surge chamber calculation. The pipeline in figure 4-40 is a 1-ft
diameter pipeline and the length from the supply reservoir to the check valve is 300
ft. If the pipe carries a steady-state discharge of 4 cfs.
(i) Determine the maximum water surface increase in a surge chamber of cross-
sectional area Ac = 0.785 ft2. If the static level at the chamber is H0 = 500 ft, how
high will the water level reach in the surge tank? What is the time T required to
reach the maximum water elevation?
(ii) With D = 1.0 ft, the pipe area is:
𝜋𝐷 1
𝐴 𝜋 0.785𝑓𝑡
4 4
(iii) With L = 300 ft, Q = 4 cfs, g = 32.2 ft/s2, then the rise in water elevation is:
𝑓𝑡
𝑄 𝜋 𝐴 𝐿 4 𝑠 𝜋 0.785𝑓𝑡 300𝑓𝑡
𝑆
𝐴 2 𝐴 𝑔 0.785𝑓𝑡 2 𝑓𝑡
0.785𝑓𝑡 32.2
𝑠

24.43𝑓𝑡
(iv) Thus, the maximum water level is:
𝐻 𝐻 𝑆 500𝑓𝑡 24.34𝑓𝑡 524.43𝑓𝑡
(v) The time required to reach the maximum water level:

𝜋 𝐴 𝐿 3.1416 0.785𝑓𝑡 300𝑓𝑡


𝑇 4.79𝑠
2 𝐴 𝑔 2 𝑓𝑡
0.785𝑓𝑡 32.2
𝑠

F. Check Valves
(1) Check valves (figure 4-46) are valves that allow flow only in one direction. The
pressure of the flowing water opens the valves in the preferred direction. A sudden
change in flow direction, such as the case of a pump failure, immediately closes the
valve avoiding a reverse flow. The lack of a check valve leaves a pump unprotected
against backflow, which may damage the pump and affect its operational efficiency.
Details on the operation of check valves can be obtained from the valve
manufacturers.

(210-634-H, 1st Ed., Feb 2022)


634-4.50
Title 210 – National Engineering Handbook

Figure 4-46: Check valve installed on a pumping plant

634.0410 Hydraulic Transients (Water Hammer)

A. The term hydraulic transient refers to a pipeline flow, and its attendant pressure, changing
rapidly with time. A hydraulic transient in a pipeline can be originated by the closing of a
valve, a pump failure, or a sudden increase in pipeline demand. Water hammers can easily be
produced by a sudden closure of a valve. In this case, a pressure wave originates at the valve
that travels through the pipe at speeds of the order of 1,440 m/s or 4,700 ft/s. The pressure
wave reflects at the other end of the pipeline, be it a reservoir or a main line connection, and
travels back towards the valve where pressure builds up again. The process repeats many
times until friction dampens the pressure wave and the water flow finally stops (figure 4-47).
Typically, a hammering noise is produced by the transient, thus the name water hammer.
Pipelines, appurtenances, and tanks subject to water hammer may suffer deformation and
damage in joints and other locations. In designing pipelines, care should be taken to prevent
the presence of the water hammer phenomenon by providing check valves, pressure relief
valves, and surge chambers in the pipeline systems.

(210-634-H, 1st Ed., Feb 2022)


634-4.51
Title 210 – National Engineering Handbook

Figure 4-47 Cycle of a hydraulic transient created by a sudden valve closing

B. Valve closure hydraulic transients can be minimized by closing the valve slowly.
Unexpected hydraulic transients, such as a pump failure, will likely produce a high-
magnitude pressure wave in pipelines with its attendant vibration. To prevent damage to the
pipeline in such a case the pipeline should be provided with valves or surge chambers. The
pipeline will be able to better survive a sudden hydraulic transient if it is buried, or otherwise
anchored securely.
C. The formal analysis of hydraulic transients in a pipeline includes the simultaneous
solution of two partial differential equations involving the discharge Q and the total head H at
different locations x along the pipeline and at different times t. This type of analysis requires
complex computer programming, considered beyond the scope of this handbook.

(210-634-H, 1st Ed., Feb 2022)


634-4.52
Title 210 – National Engineering Handbook

D. The closing of a valve is a common cause of hydraulic transients. If closing a valve


changes the flow velocity in a pipe from V to Vf, the instantaneous increase in pressure near
the valve, ∆p, is calculated by using the equation:


∆𝑝 𝑉 𝑉 (eq. 4-55)

Where ω is the specific weight of water (typically, ω = 62.4 lb/ft3), g is the acceleration of
gravity (= 32.2 ft/s2), Ev is the bulk modulus of elasticity of water, t is the pipe wall thickness,
and E is the modulus of elasticity for the pipe material. A typical value of the modulus of
elasticity for water is Ev = 311,000 psi at 60ºF (see 634.0107, Exhibit D, Figure 1-9).
Modulus of elasticity values for common pipeline materials include steel (E = 30,000,000
psi), cast iron (E = 15,000,000 psi), concrete (E = 3,000,000 psi) and PVC (E = 400,000 psi,
with some variation by manufacturing class).
E. For additional information on water hammer analyses see Finnemore and Franzini (2002)
or USDA (2005).
F. Example 4-36 – Pressure increases with sudden closing of a valve. Consider a steel pipe
(E = 30,000,000 psi) with a wall thickness t = 0.25 in = 0.25in/12 in/ft =0.02 ft, and a
diameter D = 0.5 ft.
(1) Closing a valve in a pipe that reduces the flow velocity from V = 2.5 fps to Vf = 0.5
fps for water at 60ºF in such a pipeline produces a pressure increase of:

⃓ 62.4 𝑙𝑏
⃓ 𝑓𝑡

⃓ 𝑓𝑡

⃓ 32.2
∆𝑝 2.5𝑓𝑝𝑠 0.5𝑓𝑝𝑠 ⃓ 𝑠

⃓ 1 0.5𝑓𝑡

⃓ 311,000𝑙𝑏/𝑖𝑛 144 𝑙𝑏
𝑓𝑡 0.02𝑓𝑡 30,000,000𝑙𝑏/𝑖𝑛 144 𝑙𝑏
⎷ 𝑓𝑡
16,604 𝑙𝑏⁄𝑓𝑡
16,604 𝑙𝑏 𝑓𝑡 115𝑝𝑠𝑖
144 𝑝𝑠𝑖⁄𝑙𝑏⁄𝑓𝑡
(2) The typical value for atmospheric pressure is 14.7 psi, thus, an increase of 115 psi
represents almost eight times the atmospheric pressure. Pipe should have adequate
pressure ratings or otherwise be protected to withstand hydraulic transients. This
surge pressure is in addition to the operation pressure.
(3) For more information on Calculating surge pressures refer to NEH Part 636, Chapter
52, Structural design of Flexible Conduits.

(210-634-H, 1st Ed., Feb 2022)


634-4.53
Title 210 – National Engineering Handbook

634.0411 Cavitation

A. When the local absolute pressure in a closed conduit (by its elevation and the hydraulic
characteristics of the system) falls below the vapor pressure of water, vapor bubbles, or
cavities, can form in the flow. These cavities are then carried with the flow towards zones of
higher pressure where they implode (collapse onto themselves) producing a characteristic
noise, vibration, and even pitting of the pipeline walls at the point of bubble collapse.
Cavitation conditions are to be avoided because of their damaging effects, since cavitation
can cause pitting and perforations in pipes, and even collapsing of pipes if the pressure drop
at a given location is excessively large. Locations where cavitation may occur are the high
point of a pipeline, or the suction side of a pump in a pipeline system. Cavitation is
undesirable for pump operation since it can reduce pump efficiency and cause deterioration of
the impeller (figure 4-48). An example of cavitation analysis in a siphon is presented figure
4-49. Cavitation was introduced in section 634.0104 Physical Properties of Water (5), in
relation to the vapor pressure of water.
Figure 4-48 A cavitated pump impeller

B. Figure 4-49 illustrates a conduit delivering water from reservoir (A) to reservoir (B). Such
an arrangement, where the middle part of the conduit raises above the free surface of both
reservoirs is called a siphon. To start the siphon, it is necessary to fill the conduit with water
by using, for example, a vacuum pump at point (C). Once the water flow is established the
vacuum pump can be removed and the system continues flowing as long as the free surface at
reservoir (A) is higher than that at reservoir (B). Siphon systems can be set up to overcome a
hill separating two reservoirs or to discharge water downstream over a dam.

Figure 4-49 Siphon conduit connecting two reservoirs.

(210-634-H, 1st Ed., Feb 2022)


634-4.54
Title 210 – National Engineering Handbook

C. Example 4-37 – Cavitation at high point of a siphon. Suppose that the siphon has a
diameter D = 1.0 ft, a length L =200 ft, and the material is smooth enough that the roughness
can be taken as e = 0. Let the difference of elevation between the reservoirs be H = 20 ft, and
the elevation difference of point (C) above the level of reservoir (A) be hC = 10 ft. Local loss
coefficients are Ke = 0.5 and Kd = 1.0 for the entrance and discharge points at reservoirs (A)
and (B), respectively.
(1) Find (a) The discharge in the siphon. (b) Determine the absolute pressure at high
point (C), located halfway through the conduit, and check if cavitation is likely to
occur. Assume that the water temperature is 60ºF and use an atmospheric pressure
patm = 14.7 psi. (c) What is the maximum value of hC to prevent cavitation from
taking place at point (C)?

(i) Writing the energy equation between points (A) and (B), with VA = VB = 0, using
gage pressures PA = PB = 0, taking zA = H and zB = 0, and including friction
(using the Darcy-Weisbach equation to estimate friction losses) and local losses
(∑K = 1.5), the resulting equation is:
8𝑄 𝑒 4𝑄 𝐿
𝐻 𝑓 𝐾
𝜋 𝑔𝐷 𝐷 𝜋𝐷 𝐷
634.0107D figure 1-9 gives the kinematic viscosity, ν = 1.217x10-5 ft2/s, for
water at 60ºF. A spreadsheet application gives Q = 14.73 cfs.
(ii) To determine the absolute pressure PC at point (C) write the energy equation
between points (A) and (C). LA-C =200 ft/2 = 100 ft:
𝑝 𝑉 𝑝 𝑉
𝑧 ℎ ℎ 𝑧
𝜔 2𝑔 𝜔 2𝑔
With VA = 0, gage pressure pA = 0, and elevations:
𝑧 𝐻 20𝑓𝑡
𝑧 𝐻 ℎ 20𝑓𝑡 10𝑓𝑡 30𝑓𝑡
The flow velocity at point (C) is the constant value:
4𝑄 4 14.73 𝑓𝑡 ⁄𝑠 18.75𝑓𝑡
𝑉
𝜋𝐷 3.1416 1.0 𝑓𝑡 𝑠
And the Reynolds number is:
𝑉𝐷 18.75 𝑓𝑡⁄𝑠 1.0𝑓𝑡
𝑅 1.54𝑥10
 1.217𝑥10 𝑓𝑡 ⁄𝑠
(iii) With this value of the Reynolds number (indicating turbulent flow) and relative
roughness e/D = 0, the Swamee-Jain equation (equation 4-19) produces the
following friction factor:
0.25 0.25
𝑓 . .
𝑒 𝐷 𝑓𝑡
𝑙𝑜𝑔 0.27 4.62 1.217𝑥10 𝑠 1.0𝑓𝑡
𝐷 𝑄 𝑙𝑜𝑔 0 4.62
𝑓𝑡
14.73 𝑠
0.0108

(210-634-H, 1st Ed., Feb 2022)


634-4.55
Title 210 – National Engineering Handbook

(iv) The friction losses (using the Darcy-Weisbach equation) and local head losses
(entrance losses) are calculated as follows:
200𝑓𝑡 𝑓𝑡
𝐿 𝑉 18.75 𝑠
ℎ 𝑓 0.0108 2 5.9𝑓𝑡
𝑑 2𝑔 1.0𝑓𝑡 𝑓𝑡
2 32.2
𝑠

𝑉 18.75 𝑓𝑡⁄𝑠
ℎ 𝐾 0.5 2.73𝑓𝑡
2𝑔 2 32.2𝑓𝑡⁄𝑠

(v) Substituting values calculated above into the energy equation:


𝑓𝑡
0 0 𝑝 18.75
20𝑓𝑡 5.9𝑓𝑡 2.73𝑓𝑡 30𝑓𝑡 𝑠
𝜔 2𝑔 𝜔 2 32.2𝑓𝑡⁄𝑠
(vi) Solving for the pressure head at point (C):
𝑝
24.09𝑓𝑡
𝜔
And, using ω = 62.4 lb/ft3, the gage pressure at point (C) is:
1503
𝑃 62.4 𝑙𝑏 24.09𝑓𝑡 1503𝑝𝑠𝑓 𝑝𝑠𝑖 10.44𝑝𝑠𝑖
𝑓𝑡 144
(vii) Calculating the absolute pressure at point (C) and comparing to the vapor
pressure, pv = 0.26 psi, for water at 60ºF (634.0107D figure 1-9):
𝑃 𝑃 𝑃 10.44𝑝𝑠𝑖 14.7𝑝𝑠𝑖 4.26𝑝𝑠𝑖 0.26𝑝𝑠𝑖
Since the absolute pressure at point (C) is larger than the vapor pressure of water
at a temperature of 60oF, no cavitation would occur at point (C) for the siphon
system shown.
(viii) To determine the maximum height hC to avoid cavitation at point (C), the
absolute pressure at point (C) is set equal to the vapor pressure of water at 60oF,
and the value of hC from the energy equation between points (A) and (C) is
calculated.
(ix) The gage pressure at point (C) is:
𝑃 𝑃 𝑃 𝑃 𝑃 0.26𝑝𝑠𝑖 14.7𝑝𝑠𝑖 14.44𝑝𝑠𝑖
(x) The corresponding pressure head is:
𝑝𝑠𝑓
𝑃 14.44𝑝𝑠𝑖 14.44𝑝𝑠𝑖 144 𝑝𝑠𝑖
33.32𝑓𝑡
𝜔 62.4 𝑙𝑏 62.4 𝑙𝑏
𝑓𝑡 𝑓𝑡
(xi) The velocity head, head losses, and elevation of point A are the same as in the
energy equation used above. The elevation of point (C) is:
𝑧 ℎ 20𝑓𝑡
Thus, the energy equation between (A) and (C) becomes:

(210-634-H, 1st Ed., Feb 2022)


634-4.56
Title 210 – National Engineering Handbook

0 0
20𝑓𝑡 5.9𝑓𝑡 2.73𝑓𝑡
𝜔 2𝑔
𝑓𝑡
18.75
ℎ 20𝑓𝑡 33.32𝑓𝑡 𝑠
𝑓𝑡
2 32.2
𝑠
(xii) The resulting value of hC is:
ℎ 19.23𝑓𝑡

634.0412 Culverts

A. General
(1) A culvert is a relatively short conduit connecting two open channels and typically
located under a road or highway. Figure 4-50 illustrates three possible flow regimes
when a culvert’s inlet is submerged.

Figure 4-50: Flow regimes for submerged inlet flow in culverts

(2) The flows of figure 4-50(i) and (ii) are said to be under outlet control, while that of 4-
50 (iii) is under inlet control. In figure 4-50, V stands for the flow velocity in the full-
flowing culvert; HW stands for head water (that needs to be at a minimum of 1.2 D to
provide a submerged inlet) and TW for tailwater depths, i.e., the depths of flow
upstream and downstream of the respective culverts. D is the diameter of the culvert,
h is the vertical distance from the head water surface to the centerline of the culvert,
and ∆h is the difference between the headwater and tailwater depths.

(210-634-H, 1st Ed., Feb 2022)


634-4.57
Title 210 – National Engineering Handbook

(3) Publication FHWA-NHI-01-020, HYDRAULIC DESIGN OF HIGHWAY


CULVERTS, Hydraulic Design Series No. 5, 2012, available from the Federal
Highway Administration (FHWA) provides detailed information for the analysis of
culvert flow (FHWA, HDS 5). This publication is available at:
https://www.fhwa.dot.gov/engineering/hydraulics/library_arc.cfm?pub_number=7&i
d=13
(4) The FHWA also provides a computer program, HY-8, for calculating culvert flow.
The program is available at:
http://www.fhwa.dot.gov/engineering/hydraulics/software/hy8/

B. Culvert Flow with Inlet Control


(1) This case is represented in figure 4-50 (iii). The capacity of the culvert is limited by
the capacity of the culvert opening, rather than by conditions farther downstream. In
such case, the open-channel normal depth of flow dn in the culvert is less than the
barrel height (height of the culvert section), and the entrance does not allow enough
water into the culvert to allow filling the full barrel height. Thus, the flow is under
inlet control. The discharge Q into the culvert can be calculated using the following
equation (orifice flow):
𝑄 𝐶 𝐴 2𝑔ℎ (eq. 4-49)
(2) Where A is the area of the inlet section, Cd is a discharge coefficient, g is the
acceleration of gravity, and h is the vertical distance from the HW surface to the
centerline of the inlet or the HW and TW depth difference whichever is smallest.
(3) For inlet control, the required HW head is calculated as:

ℎ (eq. 4-56)

(4) Values of the discharge coefficient Cd vary from 0.62 (for a sharp-edged inlet without
contraction suppression) to 1.0 (for a well-rounded inlet). Use the value of Cd = 0.60
as a conservative reference value. Contraction suppression refers to the absence of
contracted flow at the inlet cross-section. Suppressed contraction at the bottom of the
inlet occurs when the inlet invert is set at stream bed level. Partial contraction
suppression can be achieved by using flared wingwalls at the inlet approach.
(5) As an alternative to equation 4-49 (general orifice flow equation), FHWA, HDS 5
includes a submerged inlet control equation, based on extensive laboratory testing.
(6) Example 4-38 – Circular culvert calculation under inlet control.
A circular culvert of diameter D = 2 ft, carrying a discharge Q = 40 cfs, is to be
designed under inlet control. What is the required headwater height, h, above the
culvert centerline if the inlet discharge coefficient is Cd = 0.75?
(i) The pipe area is:

𝐴 𝜋𝐷 ⁄4 𝜋 2𝑓𝑡 /4 3.1416𝑓𝑡

(ii) And the HW head is:


𝑄 40 𝑓𝑡 ⁄𝑠
ℎ 4.48𝑓𝑡
2𝑔𝐶 𝐴 2 32.2 𝑓𝑡⁄𝑠 0.75𝑓𝑡 3.1416𝑓𝑡

(210-634-H, 1st Ed., Feb 2022)


634-4.58
Title 210 – National Engineering Handbook

𝐻𝑊 𝑑𝑒𝑝𝑡ℎ h 1⁄2 D 4.48𝑓𝑡 1.0𝑓𝑡 5.48𝑓𝑡

(iii) Checking that HW depth exceeds 1.2D:


𝐻𝑊 𝑑𝑒𝑝𝑡ℎ 4.48𝑓𝑡 1.0𝑓𝑡 5.48𝑓𝑡 1.2 2𝑓𝑡 2.4𝑓𝑡

C. Culvert Flow with Outlet Control


(1) Figures 4-50(i) and (ii) show instances of flow with outlet control. In figure 4-44(i)
the outlet is fully submerged, a condition that may result from inadequate channel
capacity downstream or by an existing backwater from a connecting stream. Even if
the tailwater depth is below the barrel height, as shown in figure 4-44(ii), the normal
flow depth dn within the culvert is greater than the culvert height D, and the culvert
flows full.
(2) Analysis of flow for the outlet control cases of figure 4-50(i) and (ii) is approached
by writing the energy equation between the free surface of the headwater section (1)
and that of the tailwater section (2). The elevation of the outlet (section 2) is taken as
a reference, so that z1 = ∆h, and z2 = 0. Since the culvert is open to the Atmosphere at
both ends, the gage pressure, P1 = P2 = 0. Also, V1 = 0, and V2 = V. Including head
losses in the culvert, the resulting energy equation is:

∆ℎ ℎ (eq. 4-57)

(3) In the simplest case, the head losses hL includes friction head losses hf and entrance
losses he. Friction head losses are calculated using Manning’s equation:

ℎ (eq. 4-58)

(4) Where n is Manning’s resistance coefficient, Cu is the coefficient in Manning’s


equation (Cu = 1.0 for units of the SI, and Cu = 1.486 for English units), and Rh is the
hydraulic radius in the culvert cross-section. Entrance losses are calculated using:

ℎ 𝐾 (eq. 4-59)

(5) A higher value for Ke, the entrance loss coefficient, gives a higher head loss. Values
of the entrance loss coefficient are presented in figure 4-11 found in section
634.0405. Also see FHWA, HDS 5 for additional values of entrance loss coefficients.
Use a value of 1 for the exit loss.
(6) Substituting the sum of equations 4-58 and 4-59 for hL in equation 4-57 produces the
equation:

∆ℎ ℎ 𝐾 1 (eq. 4-60)

(7) The approach presented in FHWA, HDS 5 includes the use of a bend coefficient Kb,
and an exit coefficient Kx (to replace the value of 1 in equation 4-60); the resulting
equation is:

𝑄 𝐴 (eq. 4-61)

(8) Where Kp is the pipe friction loss coefficient, which, in equation 4-60, is equal to:

(210-634-H, 1st Ed., Feb 2022)


634-4.59
Title 210 – National Engineering Handbook

𝐾 (eq. 4-62)

(9) A typical culvert design consists in determining the required diameter of a culvert
given the design discharge, the inlet entrance conditions, the pipe material and length,
the culvert slope, and the headwater and tailwater depths. The use of Culvert Flow,
USDA-NRCS Hydraulics Formula program is illustrated in the example 4-39.
(10) Example 4-39 – Circular culvert sizing. It is desired to install L = 50 ft of concrete culvert
pipe, n = 0.012, in a drainage channel for a road crossing. The design discharge is Q = 80 cfs
with a tailwater depth of d2 = 3.0 ft. The slope of the culvert is to be 0.002 ft/ft. The maximum
headwater depth is d1 = 5.0 ft.
(i) Determine the required pipe diameter, if the inlet has a groove edge and a
headwall.
(ii) The elevation change in 50 ft of culvert is:
∆𝑧 𝑆 𝐿 0.002 𝑓𝑡⁄𝑓𝑡 50𝑓𝑡 0.1𝑓𝑡
(iii) Since the culvert outlet elevation is taken to be “0”, the inlet elevation is 0.1 ft.
Figure 4-51 shows the calculation of the discharge for culvert flow using the
Culvert Flow tab in the USDA-NRCS Hydraulics Formula program for a guessed
diameter of 24 in = 2 ft. The result shows Q = 27.9 cfs, (see figure 4-51) which is
smaller than the required discharge of 80 cfs. The diameter is increased until Q =
80 cfs. Figure 4-52 shows the convergence of values.

Figure 4-51: Culvert discharge calculation

(210-634-H, 1st Ed., Feb 2022)


634-4.60
Title 210 – National Engineering Handbook

Figure 4-52: Culvert Discharge Convergence of Values


D (in) Q (cfs)
24 27.9
36 67.1
48 105.7
42 88.0
39 78.7
40 81.8

A standard sized concrete culvert of D = 42 in will carry the required 80 cfs.


(11) Example 4-40 Calculate the discharge in a PVC culvert with a diameter of 2 ft (24
in), and a length of 30 ft if the head wall inlet has a groove edge. The inlet invert is
located 0.23 ft above the outlet invert (set as zero), the headwater depth is 3.0 ft
(headwater elevation = 3.23 ft), while the tailwater depth is 1.5 ft (figure 4-53).
(i) The USDA-NRCS Hydraulics Formula program aids in the selection of the values
of Manning’s n and the entrance coefficient Ke. The discharge calculation is
shown below:

Figure 4-53: Culvert discharge calculation using USDA-NRCS


Hydraulics Formula Tool

(ii) Thus, the calculated discharge is Q = 22.9 cfs and the inlet controls the flow (the
equations associated with inlet control apply).
(11) In some applications the culvert headwater depth is below the barrel height
producing free or unsubmerged inlet conditions. The techniques of open channel flow
presented in 210-634-3, section 634.0316 “Gradually Varied Flow” can be used to
solve for the critical depth of flow and the headwater depth. This analysis can allow
for entrance and exit losses. As an alternative, FHWA, HDS 5 includes an
unsubmerged inlet control equation, based on extensive laboratory testing.

(210-634-H, 1st Ed., Feb 2022)


634-4.61
Title 210 – National Engineering Handbook

(12) An alternative for culvert flow calculations is the use of nomograms


D. Culvert flow solutions using nomograms
(1) This section includes nomograms for the calculation of culvert flow for different pipe
materials. A nomogram is a graphical device used for solving equations. In a
nomogram, the user pinpoints the location of two points in the scales provided and
traces a straight line to a third scale. The intersection of the straight line with the
third scale is the value of the unknown variable sought. Some nomograms include
additional scales corresponding to different conditions for the problem, as well as
turning lines to relate the solution to a third variable.
(2) The four nomograms included in this section are the same as charts in FHWA-NHI-
01- 020, HYDRAULIC DESIGN OF HIGHWAY CULVERTS, Hydraulic Design
Series No. 5, September 2001. FHWA, HDS 5 contains 55 additional charts for
hydraulic analyses of culverts of different types and flow conditions. Examples are
included for each of the nomograms.
(3) The four nomograms for this section are described as follows:
(i) Figure 4-54. Nomogram for headwater depth for concrete pipe culverts with inlet
control.
(ii) Figure 4-55. Nomogram for headwater depth for corrugated metal (CM) pipe
culverts with inlet control.
(iii) Figure 4-56. Nomogram for concrete pipe culverts flowing full with outlet
control, Manning’s n = 0.012.
(iv) Figure 4-57. Nomogram for head for corrugated metal (CM) pipe culverts
flowing full with outlet control, n = 0.024.
(4) The variables referenced in these nomograms are described as follows:
(i) D = culvert pipe diameter (in)
(ii) = Head water depth to culvert pipe diameter ratio (dimensionless)
(iii) Q = discharge (cfs)
(iv) L = culvert length (ft)
(v) H or h = head (ft), the difference in elevation between the HW depth and
downstream depth in a culvert
(5) Notice that variable h in Figure 4-50 is the same as variable H in the nomograms of
Figures 4-56 and 4-57.
(6) Example 4-41. Concrete pipe culvert with inlet control
(i) Consider a concrete pipe culvert with diameter D = 42 in = 3.5 ft, carrying a
discharge Q = 120 cfs. Determine the required head water depth (HW) for the
following conditions:
 Square edge with headwall
 Groove end with headwall
 Groove end projecting
(ii) Using Figure 4-54, first locate D = 42 in on scale (4), then Q = 120 cfs on scale
(3) and trace a straight line through these two points and extend it up to scale (1).
The point of intersection of the EXAMPLE line with scale (1) corresponds to the
value of HW/D = 2.5 for a square edge with headwall. To find the values of
HW/D for the other cases (groove end with headwall and groove end projecting),
draw a horizontal line from the point found in scale (1) to intersect scale (2) and
scale (3), respectively. The values read in scales (2) and (3) are HW/D = 2.1 and
HW/D = 2.2, respectively. The value of the corresponding headwater can be
calculated multiplying the HW/D value by the pipe diameter D. Thus, for the
three cases required in this example, the following results are obtained:

(210-634-H, 1st Ed., Feb 2022)


634-4.62
Title 210 – National Engineering Handbook

 Square edge with headwall, HW=(HW/D)*D = 2.5 × 3.5 ft = 8.75 ft ≈ 8.8 ft


 Groove end with headwall, HW=(HW/D)*D = 2.1 × 3.5 ft = 7.35 ft ≈ 7.4 ft
 Groove end projecting, HW=(HW/D)*D = 2.2 × 3.5 ft = 7.7 ft
(iii) The nomographs were computed assuming 2% culvert slopes and are accurate
within 10% for determining the required inlet control headwater.
(7) Example 4-42. Corrugated metal (CM) pipe culvert with inlet control
(i) Consider a CM pipe culvert with diameter D = 36 in = 3 ft, carrying a discharge
Q = 66 cfs. Determine the required head water depth (HW) for the following
conditions:
 Headwall
 Mitered to conform to slope
 Projecting
(ii) Using figure 4-55 first locate D = 36 in on scale (4), then Q = 66 cfs on scale (5)
and trace a straight line through these two points and extend it up to scale (1).
The point of intersection of the EXAMPLE line with scale (1) corresponds to the
value of HW/D = 1.8 for a square edge with headwall. To find the values of
HW/D for the other cases (mitered to conform to slope and projecting), draw a
horizontal line from the point found in scale to intersect scale (2) and scale (3),
respectively. The values read in scales (2) and (3) are HW/D = 2.1 and HW/D =
2.2, respectively. The value of the corresponding headwater can be calculated
multiplying the HW/D value by the pipe diameter D. Thus, for the three cases
required in this example, the following results are obtained:
 Headwall, HW=(HW/D)*D = 1.8 × 3 ft = 5.4 ft
 Mitered to conform to slope, HW=(HW/D)*D = 2.1 × 3 ft = 6.3 ft
 Projecting, HW=(HW/D)*D = 2.2 × 3 ft = 6.6 ft
(8) Example 4-43. Concrete pipe culvert flowing full with outlet control (n = 0.012)
(i) Determine the diameter of a concrete pipe culvert flowing full with a length L =
110 ft with an entrance loss coefficient ke = 0.5 if it is to carry a flow Q = 70 cfs
with a head loss H = 0.94 ft.
(ii) Using figure 4-56, first, trace a straight line from the DISCHARGE scale at point
Q = 70 cfs to the HEAD scale at H = 0.94. Next, locate the point L = 110 ft on
the LENGTH scale corresponding to ke = 0.5, and trace a straight line through
the intersection of the first line with the TURNING LINE, extending it all the
way to the DIAMETER scale. The ending point of this second line represents the
predicted value for the diameter. In this case, D = 48 in = 4 ft. Since this value
is a standard value for pipe diameters, this will be the design diameter. On the
other hand, if the diameter had been, say, D = 51 in, then, it is recommended to
use the next larger standard diameter, D =54 in = 4.5 ft.
(9) Example 4-44. Corrugated metal (CM) pipe culvert flowing full with outlet control (n
= 0.024)
(i) Determine the head loss H for a CM culvert with diameter D = 27 in and length L
= 120 ft if it is to carry a discharge Q = 35 cfs. The entrance loss coefficient is
ke = 0.9.
(ii) Using figure 4-57, first, trace a straight line from the DIAMETER scale at D =
27 in to the point L = 120 ft in the LENGTH scale corresponding to ke = 0.9.
Next, locate the point Q = 35 cfs in the DISCHARGE scale, and trace a straight
line through the intersection of the first line with the TURNING LINE,
extending all the way to the HEAD scale. The end of this line will show the
required value of H = 7.5 ft.

(210-634-H, 1st Ed., Feb 2022)


634-4.63
Title 210 – National Engineering Handbook

Figure 4-54: Nomogram for headwater depth for concrete pipe culverts with inlet control.

EXAMPLE (3)

D=42 inches (3.5 feet)


Q=120 cfs

* HW
(feet)
(1) 2.5 8.8
(2) 2.1 7.4
(3) 2.2 7.7

*D in feet

HEADWATER DEPTH IN DIAMETERS (HW/D)


DIAMETER OF CULVERT (D) IN INCHES

DISCHARGE (Q) IN CFS

SCALE ENTRANCE
TYPE
(1) Square edge with
headwall
(2) Groove end with
headwall
(3) Groove and
projecting

To use scale (2) or (3) project


horizontally to scale (1), then use
straight inclined through D and Q
scales, or reverse as illustrated.

HEADWATER SCALES 2B.3


REVISED MAY 1964
BUREAU OF PUBLIC ROADS JAN. 1963

(210-634-H, 1st Ed., Feb 2022)


634-4.64
Title 210 – National Engineering Handbook

Figure 4-55: Nomogram for headwater depth for corrugated metal (CM) pipe culverts with
inlet control.

EXAMPLE (1)
D=36 inches (3.0 feet) (2)
Q=66 cfs
(3)
* HW
(feet)
STRUCTURAL PLATE C. M.

(1) 1.8 5.4


(2) 2.1 6.3
(3) 2.2 6.6

*D in feet

HEADWATER DEPTH IN DIAMETERS (HW/D)


DISCHARGE (Q) IN CFS
DIAMETER (D) IN INCHES

SCALE ENTRANCE
TYPE
(1) Headwall
(2) Mitered to
STANDARD C. M.

conform to slope
(3) Projecting

To use scale (2) or (3) project


horizontally to scale (1), then use
straight inclined through D and Q
scales, or reverse as illustrated.

BUREAU OF PUBLIC ROADS JAN. 1963

(210-634-H, 1st Ed., Feb 2022)


634-4.65
Title 210 – National Engineering Handbook

Figure 4-56 Nomogram for concrete pipe culverts flowing full with outlet control,
Manning’s n = 0.012.

Turning

Line
For outlet crown not submerged compute HW by
other methods.

L=110
D=48
Ke=0.5

BUREAU OF PUBLIC ROADS JAN. 1963

(210-634-H, 1st Ed., Feb 2022)


634-4.66
Title 210 – National Engineering Handbook

Figure 4-57: Nomogram for head for corrugated metal (CM) pipe culverts flowing full with
outlet control, n = 0.024.

Turning

Line
For outlet crown not submerged compute HW by
other methods.

HEAD (H) IN FEET


DISCHARGE (Q) IN CFS

DIAMETER (D) IN INCHES

BUREAU OF PUBLIC ROADS JAN. 1963

(210-634-H, 1st Ed., Feb 2022)


634-4.67
Title 210 – National Engineering Handbook

634.0413 Sprinkler Irrigation

A. Sprinkler systems irrigate by spraying water in a desirable pattern. Sprinkler systems


typically include a pump, a main pipeline (or, simply, main), lateral pipelines (or, simply,
laterals), risers, and sprinkler heads.
B. Sprinkler systems can be classified as permanent, semi-permanent, portable, and
continuous move. Permanent systems have permanently located main and lateral pipelines
and a pumping plant. Semi-permanent systems consist of a permanent pumping plant and
main pipeline to which portable lateral pipelines can be attached. Portable sprinkler systems
have both portable main and lateral pipelines. Continuous move systems include center pivot
systems and linear move systems.
C. Center pivot systems consist of a continuously moving lateral pipeline that rotates about a
pivot point producing a circular irrigation pattern. A center pivot sprinkler system is shown
figure 4-58.

Figure 4-58 Central pivot irrigation system near Grace, Idaho.

D. Sprinklers can be classified as low-pressure (6-25 psi), medium-pressure (25-60 psi), and
high-pressure (> 60 psi) sprinklers. Specific characteristics for commercial sprinkler heads,
such as the pressure and discharge ranges for their operation, and the wetted diameter
(effective distance of water spraying) are typically provided by manufacturers. Sprinkler
types that operate under higher operating pressure provide a larger wetted diameter.
E. Low-pressure spray types of sprinklers include fixed sprays, spinners, and rotating sprays.
Low-pressure sprinklers have become the most common sprinkler type today.

(210-634-H, 1st Ed., Feb 2022)


634-4.68
Title 210 – National Engineering Handbook

F. A high-pressure sprinkler generally consists of one or two nozzles that rotate under the
effect of a hammer blade. The water spray impinges on a hammer blade which produces an
intermittent water spray and rotates the sprinkler head. A high-pressure sprinkler head is
typically mounted on a 1-inch, or larger (25 mm) diameter riser attached to a supply pipeline
or lateral.
G. Figure 4-59 shows a schematic for the layout of a periodic move sprinkler system (hand
line or wheel line) with a pump (P), a main line, and two lateral lines. Basic information
needed for the design of a sprinkler irrigation system includes a contour map of the plot to
irrigate (this provides information on the ground slopes on which the pipelines will be laid),
the soil characteristics (maximum application rate should not exceed the soil infiltration rate),
source and quality of water available, crops to be irrigated, and local climate.

Figure 4- 59 Schematic of a sprinkler system layout

H. Some basic guidelines for sprinkler system alignment include: (1) the laterals should be
perpendicular to the prevailing wind direction; (2) the main line should be as short as possible
to reduce head losses; (3) the pump should be located at the center of the irrigated area if
possible; and (4) provide for future expansion of the system.
I. Detailed design guidelines and calculations for sprinkler irrigation systems are available in
NRCS, Title 210, National Engineering Handbook, Part 623, Chapter 11 - Sprinkle Irrigation.
This reference may be downloaded from the engineering handbooks listed at:
http://directives.sc.egov.usda.gov/

634.0414 Microirrigation

A. Microirrigation (MI) is accomplished by the frequent application of small quantities of


water as drops, tiny streams, or miniature spray through emitters or applicators placed along a
water delivery line. MI encompasses many methods or concepts such as bubbler, drip,
subsurface drip, mist and spray. The emitters are located at or near the plant root zone thus
placing water only where the plant can use it. Thus, microirrigation can be very effective for
all crops, especially widely spaced crops (such as orchards, melons, cucumbers).

(210-634-H, 1st Ed., Feb 2022)


634-4.69
Title 210 – National Engineering Handbook

B. Microirrigation systems include a control unit or system through which water is


controlled, filtered, and possibly provided with additives. The control unit is typically located
at the highest place in the field, and the pipelines laid parallel to the terrain slope. As with
sprinkler irrigation systems, the main line in a microirrigation system is often divided into
secondary branches (manifolds) which connect to the lateral lines with the emitters. The
mains, submains, and manifolds are frequently fitted with flow or pressure regulators.
Devices are usually provided at the end of the lines to flush and clean the system (fig 4-60).
Figure 4-60 Typical parts layout of a microirrigation system

C. Water is dissipated from a pipe distribution network under low pressure in a


predetermined pattern (figure 4-61). The shape and design of the emitter reduces the
operating pressure in the supply line, and a small volume of water is discharged at the
emission point
Figure 4-61 Emitter wetted pattern on a fresh pick pea field

(210-634-H, 1st Ed., Feb 2022)


634-4.70
Title 210 – National Engineering Handbook

D. Detailed design guidelines and calculations for microirrigation systems are available in
NRCS, Title 210, National Engineering Handbook, Part 623, Chapter 7 - Microirrigation.
This reference may be downloaded from the engineering handbooks section of NRCS
eDirectives Electronic Directives System: http://directives.sc.egov.usda.gov/

634.0415 References

A. For references see 210 NEH Part 634 Chapter 1 Section 634.0105 References

(210-634-H, 1st Ed., Feb 2022)


634-4.71

You might also like