Download as pdf or txt
Download as pdf or txt
You are on page 1of 42

3.

Hydraulics
1.0. Introduction
After a treatment concept has been selected, the next step is to determine the hydraulic profile, or hydraulic grade line, needed
for the wastewater to flow through the water resource recovery facility (WRRF) and its unit operations. The objective is to
provide optimum hydraulic head to drive the facility. This may include making pipes and channels large enough to meet future
expansion beyond the capacity required during the design year. Sufficient hydraulic head should be provided to permit good
distribution of the facility flow to all treatment processes over the range of expected conditions, without being excessive. During
calculation of the hydraulic profile, the economics of building deeper facility structures should be considered in relation to
building pumping stations. Pumping stations are an important consideration because their use leads to higher operation and
maintenance costs and reduced reliability.

The engineer must distinguish the difference between hydraulic capacity and process capacity. The hydraulic capacity of a
facility is the maximum flowrate the hydraulic structures of the facility are able to carry through the facility without spilling the
water out of these structures. Whereas, the process capacity is the maximum flowrate the facility can treat while meeting
current effluent quality criteria. The hydraulic capacity must at least be equal to or (ideally) greater than the process capacity.
Facility capacity, facility treatment capacity, and process capacity are usually used as synonyms. This chapter deals mostly
with hydraulic capacity.

2.0. Hydraulic Considerations


2.1. Hydraulic Profile
The hydraulic profile describes the water level required for the wastewater to flow through the facility. A hydraulic profile is
normally prepared for the main liquid flow path that extends from the facility inlet to the receiving water, but can also be
prepared for ancillary flow trains such as sludge treatment and disposal facilities. The hydraulic profile is necessary for
establishing water surface elevations and hydraulic structure elevations and should include water surface elevations, hydraulic
control devices such as control valves and weirs, and critical elevations of process structures, channels, and pipelines. The
profile may also include ground surface elevations, pipeline sizes, and other special features that will enhance operation of the
WRRF. Figures 3.1, 3.2, and 3.3 present examples of hydraulic profiles for three commonly used waste-water treatment
processes.

Figure 3.1 Typical hydraulic profile for influent pumping and primary treatment. Water surface
elevations in feet above mean sea level represent flow of 160 000 m3 /d (42 mgd).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.2 Typical hydraulic profile for an activated sludge plant. Water surface elevations in feet
above mean sea level represent flow of 160 000 m3 /d (42 mgd).

Figure 3.3 Typical hydraulic profile for a trickling filter plant. Water surface elevations represent
flow of 85 000 m3 /d (22 mgd).

Hydraulic calculations should begin at a control point where the facility discharges to a receiving watercourse, body of water, or
the wet well of an effluent pumping station. A facility's water surface profile and pumping requirements are established from
this control point. For example, if the receiving body is a river or stream, the cotrolling elevation is the required flood elevation,
as calculated by accepted hydrological methods or as obtained from an agency such as the U.S. Army Corps of Engineers. The
level of flood protection (e.g., the 100-year flood elevation) is established by the environmental regulatory agency that governs
the respective area's facility design. With a storage basin or pond installed for 100% reuse applications, the controlling water
level is the pond-overflow elevation. For a larger body of water such as a lake, wind setup should be considered in addition to
the highest water surface elevation. With discharge to an ocean, the controlling water level is the high-tide elevation based on
the selected design storm. When discharge is to an ocean, the higher specific gravity of seawater (1.025) compared with
wastewater (1.00) is also a consideration. For every 12 m (40 ft) of depth below high tide, an additional 0.3 m (1 ft) of head
should be provided to overcome the heavier weight of seawater.

In instances where facility effluent can flow to the receiving water body by gravity most of the time, except during high-flood
stages, effluent pumping may be economically justified. In such cases, the control point of the hydraulic profile for the WRRF
will be the maximum wet-well elevation at the effluent pumping station.

The hydraulic profile is then calculated through each unit process up to the influent sewer. The calculation of the hydraulic
profile is more involved than simply adding the head losses of various components, and it should take into account the need for
equal flow distribution to and among all treatment processes. The hydraulic profile should also account for the interaction
between water depth and velocity head as it relates to the hydraulic and energy grade line.

If the available hydraulic head at the facility inlet is not sufficient to meet the required head based on the hydraulic profile,
revisions to the sizes and elevations of the hydraulic structures are necessary. If revisions do not produce the desired results,
pumping should be used at the facility inlet or elsewhere along the flow path through the WRRF. A cost-effectiveness analysis
should be performed to determine the appropriate pumping location.

2.2. Flowrates
© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
2.2. Flowrates
Peak flow is used for hydraulic design to establish maximum water levels and reveal minimum freeboards at tanks and
channels. The average flow during the month in which the maximum water level occurs is commonly used for treatment
process design. Minimum flow is used to show minimum levels throughout the facility to reveal maximum heights of free-falling
water at weirs and channels and minimum submergences on equipment. The unit processes should convey the maximum flow
unless this flow would cause a hydraulic washout of the WRRF. Equalization basins to minimize any negative effect on the
treatment process may be necessary.

To calculate the hydraulic profile through the secondary facilities, pertinent channels and piping should be designed to
accommodate the return activated sludge flow, other types of process recycle flows, and the main facility flow. When they are
substantial, process sidestream flows can also be included in the calculation of the hydraulic profile for the entire facility. The
hydraulic profile associated with the peak flow should address the probability of treatment units being out of service. For this
reason, unit process redundancy should be considered. Future facility expansion should be accounted for in sizing of piping and
channels. Minimum flow should be included in the hydraulic design to avoid low velocities and potential solids deposition in
pipes and channels.

2.3. Unit Process Liquid Levels


Each unit process should be hydraulically designed to prevent liquid from overtopping the walls of structures under all
conditions. The top-of-structure elevation is set so that the freeboard is maintained above the high-water elevation. Depending
on the degree of expected surface disturbance and the relative frequency of the condition, typical freeboards can range from an
extreme low of 150 mm (6 in.) to 1 m (3.3 ft) or more. However, the typical minimum freeboard for maximum water levels is
approximately 0.3 m (12 in.). Common hydraulic design allows for little or no submergence of the various weirs throughout the
main process flow path. Some submergence is permitted when the peak-flow condition is judged to be an infrequent
occurrence. Any effects on flow distribution as it affects treatment capacities should be considered when including
submergence in the facility's hydraulic design. In addition to considering the maximum water surface elevation in relation to
structure walls, a designer should address conditions such as surface waves, foaming, splashing (such as occurs with
hydraulic jumps), and aeration (air bulking of liquid).

2.4. Unit Process Redundancy


The design treatment capacity of a facility should include unit process redundancy. Redundancy generally means that each of
the unit processes, depending on the mechanical equipment, can have one or more tanks out of service at average flow without
adversely affecting effluent quality. However, this may affect the hydraulic design. With one tank out of service, the flow to the
unit process may exceed the maximum flow for which it was designed and cause a hydraulic overload. In addition to average-
flow conditions, peak-flow conditions with individual unit processes out of service should be evaluated. A common approach is
to provide hydraulic capacity to accommodate having one of each primary process unit out of service during peak flow, along
the worst-case flow path relative to distance and flow. The effect on treatment process during such conditions should be
reviewed. However, temporary overload of unit processes may be acceptable for short periods or during an emergency; in any
case, the structure walls should not be overtopped.

2.5. Flow Distribution

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
To ensure proper treatment, it is essential to achieve equal flow distribution to each of the tanks that make up a unit process.
Because weirs provide good flow distribution without the need for mechanical equipment, such as rate-controlling valves, and
they are less subject to clogging, the -best technical solution for flow distribution is to provide an upstream weir control
structure to split flow evenly to the tanks of each unit process. Figure 3.4a illustrates this type of layout for two clarifiers. The
use of symmetry alone will not ensure equal flow distribution because flow over long weirs is particularly sensitive to head. For
example, a small change in head on the order of 13 mm (0.5 in.) is often enough to adversely affect flow distribution to a group
of processes having long effluent weirs as the means of level control. Flow distribution to circular clarifiers is a case in point.
Because of the long weirs associated with circular clarifiers, even slight differential settling of the clarifier structure will impair
flow distribution. Figure 3.4b illustrates an example of flow control structures for facilities with asymmetrical layouts. All weirs
must be adjustable after construction.

Figure 3.4 Typical design examples of equal flow distribution among clarifiers: (a) symmetrical
layout and (b) asymmetrical layout (in. × 25.4 = mm).

An inlet header or distribution channel or pipe can also be used to distribute flow to a unit process. However, in such cases the
inlet ports or gates to each tank should be designed with adequate head loss to ensure good distribution. This is described in
more detail later in this chapter.

As with flow distribution, equal distribution of sludge to treatment units should be maintained. Unless provided for in the
design, equal sludge distribution may not occur coincidentally with equal flow distribution. This is especially common where
flow is distributed by use of channels. Where channels are used, such as upstream of grit tanks, the wastewater flow should be
well mixed to ensure that the sludge distributes evenly with the flow.

2.6. Facility Head Loss


© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
2.6. Facility Head Loss
Head loss through facilities varies widely. The specific facility designs are a function of the available head needed for pumping
and site conditions. Equal distribution of flows to treatment units requires the expenditure of head (head losses); therefore,
facilities with lower-than-normal hydraulic head may be an indication that equal flow distribution is not occurring. Although
facilities may be designed to function well hydraulically outside typical norms, the total heads commonly found for secondary
treatment facilities range from 4.3 to 5.5 m (14 to 18 ft). This range of total head applies to WRRFs with facilities for flow
measurement, pretreatment, and disinfection. Given the variability of this information, the hydraulics of each facility in relation
to site and configuration must be evaluated, and assumptions based solely on what are thought to be typical head losses
associated with various types of unit processes should be avoided.

2.7. Handling Wet-Weather Flows


Wet-weather flow conditions may put at risk effluent quality at a WRRF. Therefore, at the start of the hydraulic design process,
the design engineer must have in mind various strategies to deal with this condition.

The first question that needs to be answered is related to the estimated magnitude of the peak factor (Qpeak/Qaverage) to be
expected, both for hourly and daily peaks. Knowing the frequency at which these peaks are expected to occur will also be
beneficial for facility hydraulics design.

Ideally, 10 years of flow data taken from the collection network, plus an understanding of the planning process of that network,
are valuable information. If the collection system does not receive rainwater, and if infiltration and inflow are considered to be
low (generally the case in new developments), peak factors may be assumed to be relatively low. On the contrary, old collection
systems that combine sewage with rainwater present the worst-case scenario in terms of peak flows (i.e., high peak flows).

The design engineer will have to be cautious about which factor to use. Using the maximum hourly flow ever recorded as a
peak flow will result in extremely expensive (and in most cases, seldom-used) hydraulic structures. On other hand, taking a
peak factor out of a textbook (e.g., 2) without understanding the expected behavior of the collection system may result in a
nonrealistic design.

For example, if flow records of reasonable extension (ideally, 10 years) and accuracy (e.g., ±10%) were available, the designer
may opt to pick the 95% daily flow as the peak flow for hydraulic design. This decision will leave 5% of the daily flows (about 15
days a year) to be dealt with outside the hydraulic structure, like a bypass structure (see discussion on previous section on
Flowrates).

Flows in excess of the design peak flow may be completely bypassed (i.e., diverted at the headworks) or partly treated (e.g.,
receiving preliminary or primary treatment) and then bypassed. Some secondary facilities step-feed wet-weather flows into the
activated sludge portion of the facility as a strategy to provide biological treatment to high flows.

The decision on which level of treatment excess flowrates receive is directly related to the effluent discharge permit
considerations and cost-benefit analyses.

2.8. Minimum and Maximum Velocity


The appropriate velocity in the pipe or channel must be maintained. A minimum velocity of 0.6 to 0.76 m/s (2 to 2.5 ft/sec) for
raw wastewater is required to prevent solids deposition in channels and pipelines. Flushing velocities for raw wastewater are
1.5 to 1.8 m/s (5 to 6 ft/sec). In some cases, the minimum velocity cannot be maintained because of specific process
requirements. For example, the entrance velocity to a clarifier is reduced to 0.3 m/s (1 ft/sec) to improve the settling
characteristics and prevent disturbance of the sludge blanket. For channels where low velocities occur and sludge settling can
be a problem, mixing or aeration is provided.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
3.0. Fundamentals of Hydraulic Engineering
3.1. Hydraulic Head
Hydraulic head calculations can be classified in three types: static head, friction (including turbulence), and velocity head. Static
head represents the potential energy measured by the difference in elevation between the static levels (i.e., the level that would
be achieved if no flow were to occur) in process units and receiving waters. For pumping applications, the static head
represents the difference in elevation between the water surface elevation in the wet well and the water level at the discharge of
the force main. Friction and turbulence head losses are defined as the loss in energy and the subsequent loss in hydraulic
grade required to convey liquid through pipes, channels, fittings, inlets, outlets, reducers or increasers, or other items that create
resistance to flow. Velocity head represents the dynamic or kinetic energy resulting from the movement of the liquid. When the
flow velocity is reduced or stopped, this energy converts to static head. When the flow velocity increases, a portion of the static
head converts to velocity head. The relationship of static head and velocity head is expressed by the Bernoulli equation for
open-channel flow, as follows:

(3.1)

Where,

Y = water depth, m (ft);

Z = baseline elevation (e.g., channel invert), m (ft);

v = average velocity, m/s (ft/sec); and

g = acceleration due to gravity, 9.8 m/s2 (32.2 ft/sec2 ).

At any hydraulic section, the total energy represented by the terms in the Bernoulli equation is constant and consists of the
water elevation (Y 1 + Z 1 ) plus the velocity head (v1 2 /2g). The water elevation is associated with the hydraulic grade line, or
hydraulic profile, whereas the total of the terms in the equation is associated with the energy grade line. Accounting for head
loss occurring between downstream and upstream sections (1 and 2, respectively), the energy equation is

(3.2)

where hL = head loss between sections 1 and 2, m (ft).

Along any channel, the energy equation is used to calculate the upstream depth (Y ) given the other parameters. The head loss
(hL) is expressed as a direct value or in terms of velocity head for downstream and upstream sections (v1 2 /2g or v2 2 /2g,
respectively). Where negligible or no velocity occurs, such as in tanks, the hydraulic grade line elevation equals the energy
grade line elevation.

3.2. Pipe Flow


Pipe flow, or closed-conduit flow, occurs when the fluid completely fills the pipe and places it under pressure. If the pipe is not
full, no pressure is exerted on the pipe and the flow is characterized as open-channel flow, discussed later. This section is
specifically about pipes that are full and under pressure.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
For closed-conduit flow, the pressure head, P/ρg, is substituted for the depth, Y, in the Bernoulli equation, where P is pressure in
kilopascals or kilonewtons per square meter (pounds per square foot) and ρ is the fluid density in kilograms per cubic meter (lb
sec2 /ft4 ). For the English system, ρg is more familiarly expressed as the specific weight, γ(N/m3 [in. lb/cu ft]), and the Bernoulli
equations is thus

(3.3)

where Z becomes the elevation. These variables are illustrated inFigure 3.5.

Figure 3.5 Bernoulli relationship for a pipe.

Head loss in closed-conduit flow occurs as a result of friction between the fluid and the pipe walls. The head loss can be
measured empirically by running the fluid through a pipe and measuring the pressure at points 1 and 2. If both ends of the pipe
are at the same elevation, Z 1 = Z 2 , and because the velocity does not change, (v2 /2g) 1 = (v2 /2g) 2 , then

(3.4)

(3.5)

The relationship between hL and ΔP is not a simple one. At very low velocities, the flow in the pipe will belaminar (i.e.,
nonturbulent), exhibiting special flow characteristics and resulting in a laminar relationship between flow and pressure drop.
Then, as the velocity increases, the flow will suddenly jump to turbulent flow. In addition, the type of fluid used and the
roughness of the pipe will affect the pressure drop. In short, predicting the pressure drop from head loss in a pipe is not an easy
task and requires a semiempirical approach.

The most widely used equation for estimating the head loss between two points in a closed conduit under pressure is the
Darcy–Weisbach equation. This equation has the advantage of being dimensionally homogeneous and being applicable to all
fluids and all pipes. The Darcy–Weisbach equation is written as

(3.6)

Where,

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
hL = head loss due to friction, m (ft) of pipe;

L = length of pipe, m (ft);

f = friction factor, dimensionless;

d = inside diameter of pipe, m (ft);

v = average velocity, m/s (ft/sec); and

g = acceleration due to gravity, 9.8 m/s2 (32.2 ft/sec2 ).

All of the variables in the Darcy–Weisbach equation can be controlled or estimated except the friction factor,f. This term is
estimated using a plot of f versus the Reynolds number, defined as

(3.7)

Where,

d = pipe diameter, m (ft);

v = velocity, m/s (ft/sec);

ρ = fluid density, kg/m3 (lb-sec2 /ft4 ); and

μ = fluid viscosity, kg·m/s (lb-sec/sq ft).

The plot of the friction factor versus the Reynolds number is commonly called theMoody diagram, found in many basic
hydraulics and fluid mechanics books.

Because the velocity, v, appears in both the Reynolds number and the Darcy–Weisbach equation, a solution using the Moody
diagram is often trial and error. The calculation is simplified in the case of water. The density and viscosity of water are known.
Solving the Darcy–Weisbach equation for velocity, as

(3.8)

(3.9)

Where,

d = pipe diameter, m (ft);

K = a constant, function of the pipe roughness; and

S = slope of the hydraulic gradient, or hL/L.

The most popular equation used in U.S. engineering practice for estimating head losses in closed conduits is the Hazen–
Williams equation, which is based directly on the above modification of the Darcy–Weisbach equation. The Hazen–Williams
equation was derived by conducting numerous experiments with different pipes and back-calculating the constant K. The
Hazen–Williams equation for U.S. customary units is written as

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(3.10)

Where,

v = velocity (ft/sec);

R = hydraulic radius (ft);

= d/4 for circular pipes, where d = pipe inside diameter (ft);

S = slope of the energy gradient, (hL/L);

L = pipe length (ft); and

C = the Hazen–Williams friction coefficient.

Note that the powers 0.63 and 0.54 in eq 3.8 are very close to 0.5 as in the Darcy–Weisbach equation.

Solving the Hazen–Williams equation for head loss

For international system of units (SI):

(3.11)

For U.S. customary units:

Where,

hL = friction head loss, m (ft);

v = mean velocity of flow, m/s (ft/sec);

R = hydraulic radius, area/wetted perimeter, m (ft);

C = surface-roughness coefficient, dimensionless; and

L = pipe length, m (ft).

The surface-roughness coefficient varies with the pipe's type, age, and condition. Older ductile or cast iron pipes can have a
surface-roughness coefficient value of 100; new polyvinyl chloride pipes can have a surface-roughness coefficient value of 140.
When using new pipes and determining pump heads, the designer should consider two surface-roughness coefficient values:
one corresponding to the new condition of the pipe and the other to the design life of the pipe.

A useful relationship can be derived from the Hazen–Williams equation by recognizing first that the equation can be solved for
flowrate if the continuity equation, Q = Av, is used where Q is the flowrate (m3 /s or cu ft/sec) and A is the area through which the
flow occurs (m2 or sq ft), which is equal to πd2 /4. The Hazen–Williams equation (U.S. customary units) can then be written as

(3.12)

For a given pipe, C (roughness coefficient) and d (diameter) are the same. The relationship between any one flowrate,Q1 , and
any other flowrate, Q2 , through the same pipe is then

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(3.13)

and because L, the length of pipe, is the same, and S = hL/L,

(3.14)

This equation is especially useful when system head curves are to be constructed for pumped systems being designed. This is
discussed further below.

Head losses caused by pipe fittings such as valves, bends, entrances, exits, and sudden enlargements or constrictions, are
caused by friction and turbulence as flow passes through the devices. Almost all of the head loss through fittings, however, is
attributed to turbulence as the flow strives to return to hydraulic equilibrium after being distorted. It is better to calculate fitting
losses using a constant, K, for the type of fitting, multiplied by the velocity head pertaining to the fitting. However, many
engineers use the equivalent-length method to calculate head losses associated with fittings. Either method will produce
satisfactory results; however, for fluids with viscosities different from that of water, the equivalent-length approach will be
incorrect. The K values for commonly used fittings can be found in hydraulics handbooks. Using theK-value method, head loss
for pipe fittings can be calculated using the following equation:

(3.15)

Where,

hL = head loss, m (ft);

K = constant for type of fitting, dimensionless;

v = velocity in pipe, m/s (ft/sec);

g = acceleration due to gravity, 9.8 m/s2 (32.2 ft/sec2 ).

3.3. Open-Channel Flow


© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
3.3. Open-Channel Flow
When there is a free surface in a pipe or channel, the flow is open-channel flow. As the name implies, open-channel flow is
associated with channels. Open-channel flow may also occur in unsurcharged influent sanitary sewers or in effluent outfall
lines. The Manning equation is commonly used to calculate head loss for open-channel flow; in SI units, it is

(3.16)

And, in U.S. customary units, it is

(3.17)

Where,

hL = head loss, m (ft);

R = hydraulic radius (cross-sectional area divided by wetted perimeter), m (ft);

n = roughness coefficient, dimensionless;

v = velocity, m/s (ft/sec); and

L = channel length, m (ft).

The roughness-coefficient values depend on the roughness of the conveyance channel and range from 0.009 for glass to more
than 0.06 for natural channels. The most commonly used materials for channels in WRRFs are concrete and steel. New
concrete and steel can have roughness-coefficient values as low as 0.010; as the materials deteriorate, roughness-coefficient
values ranging from 0.013 to 0.016 can be expected. Gravity Sanitary Sewer Design and Construction (ASCE et al., 2007)
provides a listing of widely accepted roughness-coefficient values.

In addition to head loss in channels, pipes, fittings, and valves, head loss also occurs with in-line equipment, flow meters and
flumes, and weirs. In-line equipment and flow meter head loss data can be obtained from equipment manufacturers. Flumes,
such as Parshall or Palmer–Bowlus flumes, are used for open-channel flow measurement. Head loss for flow through weirs,
launders, ports, and distribution headers is described in the following sections.

3.4. Weir Control


Although weirs are sometimes used for flow measurement in facilities, they more commonly serve as control devices to
maintain a required water surface elevation in a unit process. Weirs are classified in accordance with the shape of the notch.
Weir types include rectangular, V-notch, trapezoidal, proportional, and parabolic. The upper edge of the weir or weir plate is
referred to as the crest of the weir. The depth of the water over the crest is thehead. The head, or depth, over the weir is
measured as the difference between the upstream water surface elevation above the point where weir drawdown occurs and
the weir crest elevation (see Figure 3.6).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.6 Weir sections: (a) sharp-crested weir and (b) broad-crested weir.

Weirs are either sharp-crested or broad-crested, as shown in Figure 3.6. The most commonly used weirs in WRRFs are
rectangular, straight-edge, and V-notch. Weir plates are bolted to a wall or trough so that they can be leveled independently of
the structure. V-notch weirs are best suited for long weirs, such as those found in circular settling tanks.

V-notch weir angles range from 22.5 to 120 deg. The 90-deg V-notch is the most commonly used notched weir, particularly for
circular settling tanks. Under free-flow (unsubmerged) conditions, as shown in Figure 3.7a, the head over a sharp-crested V-
notch weir can be calculated with the following -equation:

(3.18)

Where,

Q = flow, m3 /s (cu ft/sec);

φ = angle of the notch, deg

H = head over the crest, m (ft); and

C = weir coefficient, 1.38 for SI units (2.5 for U.S. customary units), for a 90-deg weir.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.7 Weir flow conditions: (a) free flow and (b) submerged flow.

Head over a sharp-crested rectangular weir under free-flow conditions can be calculated using the following equation:

(3.19)

Where,

Q = flow, m3 /s (cu ft/sec);

L = length of weir, m (ft);

H = head over the crest, m (ft); and

Cw = coefficient that accounts for the approach velocity in terms of the ratio of the weir plate depth to the head over the
crest. The values range widely depending on the approach velocity. The most commonly used value, 1.82 for SI units (3.3
for U.S. customary units), accurately represents deepwater upstream, such as that found in tanks, with an extremely low
approach velocity toward a sharp-crested rectangular weir.

Often, a rectangular weir has end contractions (i.e., the ends of the weir project inward from the sides of the channel). End
contractions cause the nappe to be aerated, thereby preventing pulsations of the upstream water surface. The shape of V-notch
weirs also permits nappe aeration. The effects of the end contractions for rectangular weirs are accounted for by subtracting
0.1H, where H is the weir head over the crest, from L, the length of the weir, for each end contraction. The equation for a
rectangular weir with a contraction on each end is as follows:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(3.20)

These equations apply under free-flow conditions. When the downstream water elevation rises above the weir crest, the weir
becomes submerged, as shown in Figure 3.7b. Submerged weirs are not used for flow measurement because the weir
equations do not directly apply to submerged conditions, and thus produce sizable errors. The head over the crest can be
calculated from the equations; however, the discharge flow must be corrected by using curves that have been developed
experimentally (King and Brater, 1963). The head created upstream of a submerged weir should be adjusted accordingly.

3.5. Effluent Launders


Effluent launders are found at the outlet end of unit processes, such as settling tanks. These launders, or channels, serve to
collect flow along its length as the flow moves downstream toward the outlet end. Conveying flow along the launders requires
head to overcome the friction loss along the channel and the exchange of momentum as the water falls perpendicular to the
flow stream and accelerates along the direction of flow. The side-overflow formula, described in detail in Open-Channel
Hydraulics (Chow, 1959), is a differential equation for flow with increasing discharge. Because of the low average value of the
Reynolds number along the entire length of the launder, use of a higher-than-normal friction factor is required (e.g., 50% higher).

3.6. Flow Distribution (Channel and Pipe Headers)


Ports, or gates, are sometimes used to equally distribute flow from influent channels or pipe headers to unit processes, such as
shown in Figure 3.8. The sensitivity of the process dictates the need for equal flow distribution to each parallel tank and also
within each tank. The head loss required to equally distribute the flow must overcome the hydraulic gradient along the channel
or pipe. All things being equal, the flow distribution scheme shown in Figure 3.4 provides a much more effective distribution
than the one shown in Figure 3.8. With feed coming from one end of the channel, there is a tendency for the first channel in the
series to receive more flow than the next, and so on. The channel thus must be sized so as to minimize head loss along its
length. In spite of their disadvantages, flow distribution structures like the one shown in Figure 3.8 are still used in WRRFs and,
for that reason, the editors include a section on it in this textbook.

Figure 3.8 A manifold or distribution channel.

The hydraulic relationship between the header and gates in which the outflow is continuous or nearly continuous along the
length of the header is (Camp and Graber, 1968; Fair et al., 1968)

(3.21)

Where,

hLp = head loss through port, m (ft);

Δh = hydraulic grade differential along the channel or pipe, m (ft); and

m = ratio of the lowest flow and highest flow through the respective ports.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Energy in the form of head loss is required to cause the equal distribution of flow from the header to the process tank. This
equation is used to determine the head loss that must be induced by the inlet gates or ports to ensure uniform distribution to all
tanks from a common inlet header or distribution channel. For example, to keep the flow to port 1 in Figure 3.8 within 5% (q1 /q5
< 1.05) of that to any other port such as port 5, the head loss through the gate should be approximately 10 times the hydraulic
grade level differential over the entire length of the distribution channel or pipe header.

To determine the actual head loss through a gate or port, the following equation applies:

(3.22)

Where,

hLg = head loss through the gate, m (ft);

Q = flow, m3 /s (cu ft/sec);

C = gate or orifice coefficient, dimensionless;

A = area of gate or port opening, m2 (sq ft); and

g = acceleration due to gravity, 9.8 m/s2 (32.2 ft/sec2 ).

Values of the gate, or orifice, coefficient for calculating head loss through ports under varying conditions can be found in
Handbook of Applied Hydraulics (Davis and Sorensen, 1969).

The hydraulic gradient differential is influenced by velocity head and head loss along the upstream header pipe (Camp and
Graber, 1968)

(3.23)

Where,

Δh = hydraulic grade differential, m (ft);

v0 = header inlet velocity, m/s (ft/sec);

g = acceleration due to gravity, 9.8 m/s2 (32.2 ft/sec2 ); and

hLh = header head loss, m (ft).

When the outlet flow from a pressurized header approximates a uniform, continuous outflow along the length of the header, the
head loss along the header can be estimated by calculating the head loss as if the inlet flow to the header were to be conveyed
along the entire length of header and dividing the result by 3. This estimation procedure is valid for the ratio of the lowest flow
and highest flow through the respective ports greater than 0.9.

4.0. Unit Process Hydraulics and Other Hydraulic Elements

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Within each unit operation, devices are used to distribute flow, maintain a certain water depth, and control the flow. Typical
devices include shutoff gates, weir gates, valves, ports, weirs, baffles, orifices, launders, and underdrains. Each of these devices
imposes a head loss on the system and should be considered in the hydraulic calculations. For example, Figure 3.9 shows that
head loss must be calculated between the manhole and clarifier, the side-weir equation is needed for calculating the water
elevation in the clarifier effluent launder, and the V-notch weir equation is needed for calculating the head over the weir crest.

Figure 3.9 Hydraulics for a typical clarifier.

If the flow is to be equally distributed to multiple tanks or within an individual tank, distribution boxes, channels, and header
pipes can be used for this purpose. Although modulating ports, gates, or valves can provide equal distribution, they require
control systems that share the disadvantages inherent to any electrical and mechanical system (e.g., failure and high
maintenance). Where possible to install, static devices are always better suited for the distribution of flow.

4.1. Screens
© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
4.1. Screens
Some processes can impose varying heads. One such process is the influent screen. The wastewater flowing through the bars
of the screen creates head loss. The head loss increases if the screen becomes partially or wholly clogged by debris. As the
screen is cleaned by mechanical rakes or other means, the head loss decreases. The typical head loss for bar screens ranges
from 0.15 to 0.76 m (0.5 to 2.5 ft). The equation for calculating head losses through screens is as follows:

(3.24)

Where,

hLs = head loss through the screen, m (ft);

vB = velocity through bar screen, m/s (ft/sec);

v = velocity upstream of bar screen, m/s (ft/sec);

g = gravitational acceleration, 9.8 m/s2 (32.2 ft/sec2 ); and

k = head loss coefficient (1/0.7).

Expressing the upstream velocity in terms of multiples of the velocity through the screen (clean [full flow area] or partially
clogged [decreased flow area]) simplifies the calculation of head loss. Knowing the downstream water depth, the Bernoulli
equation is used to determine the upstream water depth. The highest head loss through the screens is based on the maximum
clogging of the screen that can be tolerated. For mechanically cleaned screens, the head loss or change of head through the
screens is regulated by a timer or, in some cases, a differential level sensor. Best design is to use a timer with an upstream
head override. Head differential override is also used by many engineers.

4.2. Grit Tanks


Regardless of the type of grit tank used (gravity, aerated, etc.), equal distribution of flow and grit to the tanks is important.
Although the flow may be equally distributed, equal distribution of grit may not occur. This is especially true when inlet header
channels are used. Flow separation points (such as those found at sharp corners) along the inlet header channels can cause
dead zones where grit drops out. Once settled, the grit tends to eventually migrate to the closest tank or tanks. Other than
eliminating the possibility of dead zones, which is sometimes a difficult task, the grit-laden inlet flow can be aerated to keep the
flow well mixed and the grit in suspension. The use of air to keep the grit suspended and well mixed may require odor-control
facilities to treat the offgases.

4.3. Primary Settling Tanks


© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
4.3. Primary Settling Tanks
Much of the head loss found in WRRFs is associated with inlet piping to primary settling tanks. One of the largest losses can be
caused by the inlet ports to circular settling tanks. The size of the ports at the top of the inlet columns, and the associated head
losses, often cannot be predicted because the required information depends on the tank manufacturer that is selected. To cover
this uncertainty, tank manufacturers should be consulted for the appropriate head loss value during the design phase of a
WRRF project. In addition, the head loss caused by the inlet ports must be determined. The inlet column and inlet port head loss
at the maximum design flowrate should be included in the equipment specifications for the settling tank.

In addition to the equal flow splitting required for each settling tank, equal distribution of flow is required within each tank itself.
A rectangular settling tank, for example, requires good distribution of flow along its inlet conduit and through its ports or inlet
gates. Best design seems to be the use of ports with diffusers or baffles.

4.4. Aeration Tanks


As with any unit process, equal distribution of flow to individual aeration tanks and within each tank is important. Head loss is
needed to ensure equal distribution of flow. This generally occurs in the form of head loss caused by inlet gates. Inlet weirs can
also be used to distribute flows, although they are not as effective hydraulically. The inlet gates should be sized to create
sufficient head loss to overcome the head differential along the inlet header. To reduce excessive head losses through the inlet
gates, large header channels are used to keep the head differential along the channel low. The channels may be aerated to keep
sludge in suspension. Aeration of the channels is not an option in cases where the treatment process would be affected (i.e.,
anoxic tanks).

The mode of operation for the aeration tank should be addressed during calculation of the hydraulic profile. Aeration tanks can
be operated as completely mixed flow, plug flow, or stepped flow (feed) and its variations. The calculation of the hydraulic
grade line should anticipate unplanned changes to the operating mode because of future operational requirements. Regardless
of the mode, the equal distribution of flow to each tank and its cells should be adequately addressed during the hydraulics
design. To prevent excessive head losses, multiple ports of different sizes—each dedicated to a particular operating mode—
may be required. In facilities that experience large diurnal flow variations, the parallel gates in the distribution channels must be
frequently adjusted if the flow to various tanks is to be equalized. The effluent collector channel for aeration tanks can have
zones of low velocity. To prevent sludge deposition, however, aeration of the effluent channel may be necessary.

4.5. Secondary Settling Tanks


Head loss caused by the inlet column and ports should be estimated and included in the equipment specifications for circular
settling tanks. Often, return activated sludge is drawn directly from the settling tank hopper bottom. For some activated sludge
treatment facilities, the recirculated activated sludge is drawn off via a pipe situated within the inlet column. For such piping
configurations, the return activated sludge is withdrawn from above the inlet ports to the settling tank. Head losses affected by
the obstructing return activated sludge withdrawal pipe should be included in the calculation of the column and port head
losses.

4.6. Disinfection Tanks


© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
4.6. Disinfection Tanks
Equal distribution of flow to and within disinfection or contact tanks is important. Because sludge settling is not as much of a
concern, the tanks are designed to operate at low velocities. This permits head loss to be low. Head loss is kept at a minimum
and is concentrated at the inlet to cause equal flow distribution and at the effluent weir to maintain proper water depth. To
minimize weir head and to allow for proper operation of scum-removal equipment intolerant to wide water-level fluctuations,
finger weirs are commonly used. Finger weirs consist of multiple channels fitted with weirs that extend out into the tank at its
outlet end. Baffles are often used to minimize short circuiting.

4.7. Other Unit Processes


Other unit treatment processes, such as trickling filters, rotating biological contactors, membrane bioreactors (MBRs), and
gravity filters, have hydraulic requirements, some of which are peculiar to their design. For trickling filters to operate, sufficient
water surface elevation must exist at the inlet structure to push the water through the distributor and keep it rotating over the
media. Rotating biological contactors have minor head losses because water only has to flow through the trough. For effluent
gravity filters, sufficient water surface elevation must exist over the filter media to convey the liquid through the media and the
underdrains. Therefore, the manufacturers of these and other treatment units should be consulted to gather information on the
water depth, velocity, and head loss requirements of the respective equipment.

4.8. Yard Piping


Yard piping is the system of pipes connecting the various processes. The construction material for yard pipe is almost always
ductile iron pipe because of its strength and ease of disassembly and repair. Joints are commonly flanged for ease of repair.
Piping should never be buried under concrete slabs or other tanks.

4.9. Outfalls
For ocean discharge, the higher density of seawater (1027 kg/m3 [1.99 lb-sec2 /ft4 ]) must be taken into account when
computing the static head at the outfall outlet. (Within the U.S. customary unit system, the more familiar specific weight of 64
lb/cu ft is used to calculate the static head of seawater.) Although distribution along an outfall's diffuser outlets is based on the
same basic principles as those discussed previously in this chapter, proper distribution of effluent through the outlet diffusers
depends on many other complicated parameters and interactions. Often, a detailed analysis of the currents in receiving waters
is required before the hydraulic design of the outfall can be initiated. Outfalls should be sloped slightly upward, toward the
effluent end, and not be level or be sloped downward. This allows bubbles to flow downstream along the crown and avoids
bubble buildup and air binding in the pipe.

Additional considerations for the proper hydraulic design of outfalls include seawater purging for deepwater tunneled diffusers
with long riser pipes; air in the outfall as it affects flow capacity, buoyancy, and related structural failure; and surging and related
water hammer (Grace, 1978; Gunnerson and French, 1996).

5.0. Pumps and Pumping


5.1. Pump Characteristics

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Pumps are used to add energy to fluids such as water and wastewater. In closed conduits, centrifugal pumps with various
types of impellers are the most popular because they are efficient and dependable. In some cases, for example in the pumping
of viscous materials such as sludge, positive displacement pumps are useful. Dewatering devices are often fed by progressing
cavity pumps that squeeze the sludge through pipes without surges. In open channels, Archimedes screws are the most
popular devices for efficiently and effectively lifting wastewater moderate elevations. Sludge pumps are discussed further in
Chapter 12. The discussion that follows applies to rotating impeller pumps, the most popular pumps in WRRFs.

The ability of a pump to function in a given application is described by its characteristic curves, defined as the head developed
by the pump versus the flowrate produced. The head (H) versus flowrate (Q) curve is constructed for a given pump using a
setup such as pictured in Figure 3.10 (generally, a second valve is installed upstream of the pump to isolate the pump from the
positive head to allow maintenance/replacement—this valve is not shown in Figure 3.10). A pump is installed in the line from a
reservoir with positive suction pressure (the water level in the reservoir is above the elevation of the pump). The valve at the
pump-discharge side is closed. Pressure in the pipeline can be measured immediately before and after the pump.

Figure 3.10 Pump setup for establishing its characteristic curve.

If the valve is closed and the pump is not on, the pressure at both points (shown as manometer tubes inFigure 3.10) are at the
same level as the water in the reservoir. If the pump is now turned on, the pressure at the intake side remains the same, but the
pump causes the pressure in the discharge side to increase by some value HO, measured in meters (or feet) of water in the
manometer tube. This value is plotted on an H versus Q plot as shown in Figure 3.11.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.11 The pump characteristic curve.

If the valve on the discharge side is now opened slightly, some flow will result, but total energy has to be equal so that the height
of water in the manometer tube will drop slightly, indicating a lower pressure in the discharge line, the pressure head being
converted to velocity head. These two values of H and its corresponding Q, call them H1 and Q1 , are then plotted in Figure 3.11.
As the valve is opened wider and wider, H continues to drop as Q increases, resulting in a plot known as the characteristic curve
for this pump. Often a given style of pump (rotational speed, impeller width, angle of impellers, etc.) is manufactured in different
diameters. If each pump is tested, a family of characteristic curves will result: each curve is specific to that pump impeller.

The amount of energy used to move the water and develop the pressure head is known as thebrake horsepower, or BHP. The
energy output from the pump is known as water horsepower, WHP, and defined as

(3.25)

Where,

Q = flowrate, m3 /s (cu ft/sec);

H = head developed, N/m2 (lb/sq ft); and

WHP = water horsepower, N-m/s (lb-ft/sec).

Thus the units of WHP are N·m/s, or power.

The efficiency of a pump is measured as the ratio of WHP (energy in) over BHP (energy out), as

(3.26)

For any run with the pump as discussed above, the energy input to the pump is measured as the kilowatts used, and the
resulting output is calculated by multiplying the flowrate at the corresponding head produced. Thus, for each flowrate, an
energy efficiency can be calculated, and this can be plotted on the H versus Q plot. If the family of pump impellers of a given
design is similarly tested and the efficiencies are calculated, a series of isoefficiency curves will result, as shown in a typical plot
in Figure 3.12. This plot is then a complete picture of the characteristics of this pump and is used in the selection of a pump for
a given application.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.12 A series of characteristic curve and associated efficiencies for a family of pampas (d =
impeller diameter).

5.2. System Head Curves


An application of a pump requires that the job the pump is supposed to do be defined. This is done by drawing asystem head
curve.

Suppose a specific application requires that water be pumped from the elevation in tank A to the elevation in tank B, as shown
in Figure 3.13. The length of the pipe and all of its appurtenances such as valves, elbows, and other sources of minor losses,
are well defined. The static head, or the difference in elevations, is HO and this is plotted on an H versus Q plot, as shown in
Figure 3.14.

Figure 3.13 Pumping from reservoir A to reservoir B.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.14 System head curve.

An arbitrary flow is selected and the head loss at that flow is calculated, using equations such as Darcy–Weisbach or Hazen–
Williams. For some flowrate, Q1 , the head loss is then calculated as H1 , including losses due to valves, pipe fittings, meters, and
so on. This value of H1 is then plotted at the corresponding Q1 . To complete the curve, other values of Q must be chosen and
the corresponding head losses calculated. This is, however, a tedious task and the usefulness of the equation

(3.27)

now becomes obvious. If the head loss is known for any flow, the head loss for any other flow can be calculated using this
simple ratio. The full system head curve is then drawn.

5.3. Pump Selection


For a given application, as described by the system head curve, a pump is selected by superimposing the system head curve on
the characteristic curves and selecting a pump that has a high efficiency. Figure 3.15 shows how the system head curve from
Figure 3.14 is superimposed on the pump characteristic curve, Figure 3.12. The point at which these two curves, the
characteristic curve and the system head curve, intersect, is known as the operating point. If this pump is applied to the system
of pipes and appurtenances for which the system head curve was drawn, and the pump is turned on, it should discharge the
amount of water defined by the operating point.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.15 System head curve superimposed on pump characteristic curves to determine
operating point.

Sometimes the head to be developed is quite high, and a single pump will not be sufficient to produce the necessary energy. In
that case, two pumps can be placed in series, as shown in Figure 3.16. The H versus Q curves are added (vertically) to produce
the characteristic curve for the two pumps in series. In other cases, any single pump will be inadequate to provide the flowrate,
at a modest head, and several pumps are then placed in parallel, as shown in Figure 3.17. In this case the system head curves
are added horizontally. Such an arrangement is used most frequently where a significant variation in flow is expected, such as
influent pump stations.

Figure 3.16 Pumps in series.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.17 Pumps in parallel.

5.4. Wet-Well Design


No single method for sizing wet wells applies to all design situations. Proper wet-well sizing considers four critical factors—
detention time, mixing, pump cycle time, and turbulence at the pump intake.

As good practice, wet-well detention times generally should not exceed 30 minutes for average flow, to minimize generation of
unpleasant odors. In colder climates, longer detention times may be acceptable. Where such detention time limitations would
be impractical, odor mitigation must be accounted for in wet-well design. Odor mitigation practice ranges from providing gas-
tight covers to chemical feed or offgas scrubbing systems.

Pump cycle time refers to the elapsed time between successive motor starts (i.e., the time to fill the wet well plus the time to
empty it). Excessive motor wear and shortened service life result from cycle times less than the manufacturer's
recommendation. Minimum cycle times range from approximately 5 minutes for 4-kW (5-hp) motors to over 30 minutes for
150-kW (200-hp) motors. The motor manufacturer should be consulted for minimum cycle time recommendations or special
motor designs.

Minimum wet-well volume can be determined from the following equation:

(3.28)

Where,

V = wet well volume, m3 (gal);

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
C = cycle time, minutes; and

Q = pump capacity, m3 /min (gal/min).

For constant-speed pumps, the minimum cycle time results if the influent flow equals 50% of the rated pump capacity. This
limitation often determines wet-well volumes for both single- and multiple-pump installations. For multiple-pump installations,
alternating the lead pump after each pumping cycle effectively doubles the cycle time and reduces wet-well volumes
accordingly. Wet-well volumes also can be optimized with strategic pump "on" and "off" settings. In multiple constant-speed
installations, the required wet-well capacity represents the sum of the wet-well capacities required for the individual pumps.
Such an allowance will prevent cycling when lag or standby pump units enter service.

Some designers use detention capacity in the influent sewer line to minimize wet-well volume by setting the pump "off" level
above the invert elevation of the influent sewer. This practice may be acceptable in large installations, where influent
wastewater velocities are sufficient to minimize solids deposition in the influent sewer.

Wet wells for variable-speed pumping systems can be significantly smaller than for comparably sized constant-speed stations.
In determining the volume required for variable-speed pump applications, two factors need to be considered—(1) the capacity of
the pump at minimum speed, and (2) pump operation at constant speed. The latter would occur if the variable-speed drive
control failed and the pump would be "forced" to run in a constant-speed mode. If sufficient standby capacity is available, this
may not be of concern. The capacity at minimum speed will vary with each installation and can be determined from the system
head curve analysis presented previously. Equation 3.28 can be used, but the Q value is the pumping capacity at minimum
speed. If the pump has the potential to operate at constant speed for long periods of time (condition 2 above), consideration
should be given to increasing the wet-well volume available to that of a constant-speed pump. This is determined from eq 3.28,
using Q equal to the maximum pumping capacity.

In multiple variable-speed installations, the wet-well capacity required is the sum of the wet-well capacities required for the
individual pumps. Such an allowance, similar to multiple constant-speed installations, will prevent cycling when lag or standby
pump units enter service.

When determining the pump operating levels in the wet well, the design engineer needs to consider the net positive suction
head (NPSH) requirements of the pump. Wet-well designs should allow adequate submergence and clearance between pump
intakes, to prevent turbulent currents and vortexes that could otherwise reduce pump efficiency or capacity. These
requirements may dictate longer detention times than those necessary to meet pump cycling requirements. The pump
manufacturer's recommendations and the Hydraulic Institute Standards for Centrifugal, Rotary, and Reciprocating Pumps
should be consulted for sizing sumps and configuring wet wells. Adherence to these suggestions, based on testing by several
pump manufacturers, will help ensure proper pumping station design and avoid costly future modifications. For large, complex
wet-well systems and very large pumps, where established design practice would be impractical, hydraulic (physical) modeling
will be necessary to ensure acceptable wet-well performance.

6.0. Common Problems Encountered in Facility Hydraulics


© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
6.0. Common Problems Encountered in Facility Hydraulics
Two of the most common problems encountered in facility hydraulics are inaccurate flow measurement and uneven flow
splitting. These two problems not only affect facility hydraulics per se, but also impact the overall performance of the facility.

Accurate flow measurement is needed for several reasons, among them

1. To keep records to be used in future facility expansions;

2. To calculate contaminant loadings to the receiving water body; and

3. To be able to dose chemicals for various treatment purposes.

Inaccurate flow measurement may be caused by faulty design and/or installation of flow metering devices (e.g., wrong
equipment selection, inadequate minimum distances, variations from original design not taken into account when installing
flow meters). Designs and construction engineers must put emphasis on ensuring that flow metering meets expected ranges
of accuracy.

Uneven flow splitting, that is, not distributing flow to facility trains in equal parts, results in facility trains operating under
different hydraulic and contaminant loadings, stressing one or more trains, while underloading the remaining train(s). Uneven
loading does not average out to result in similar effluent quality than that expected in a facility with even flow splitting. In some
cases, particularly under higher flows, uneven flow splitting may result in not meeting effluent standard requirements.

Uneven flow splitting is generally caused by faulty design of flow division structures. Experience shows that, in general, flow
splitting should be done using overflow structures (weirs), as opposed to underflow sluice gates. The hydraulic designer must
decide on the most adequate flow splitting structure for each particular case.

7.0. Design Examples


7.1. Partial Hydraulic Profile in International System of Units
The hydraulic profile for the primary and secondary processes for a WRRF should be based on the plan schematic shown in
Figure 3.18 and the profile shown in Figure 3.19. The invert elevation for the primary effluent channel and the elevation of the
notches for the primary clarifier effluent weir should be set.

Figure 3.18 Plan—design example—partial hydraulic profile.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.19 Profile—design example—partial hydraulic profile.

7.1.1. Input Parameters


Design peak flow = 630 m3 /h = 0.175 m3 /s.

RAS = 50% of the influent wastewater flow, and is returned to the junction box downstream of the primary clarifiers.

Primary and secondary clarifiers are circular, with 90-deg v-notches, 230 mm center to center; weir elevation shown is at the
bottom of the notch.

Primary clarifier effluent channel is rectangular in cross-section, 460 mm wide; the invert is horizontal (slope = zero).

Each aeration basin has an effluent weir, which is rectangular and has a 1.8-m crest length.

The mixed-liquor splitter box has two rectangular weirs, each 1.8-m crest length; there is one for each of the secondary
clarifiers.

7.1.2. Assumptions
All weirs are sharp-crested; rectangular weir coefficient = 1.82; v-notch = 1.38.

Gate and orifice coefficient = 0.60; sharp, not rounded.

Ignore end contractions at weirs.

Manning's n = 0.015 for channels and pipes.

Entrance loss coefficient, K = 0.5; exit loss coefficient, K = 1.0; 90-deg elbow loss, K = 0.4.

Ignore losses in the aeration basin influent channel and mixed-liquor effluent channel (this is done for expediency here.
These losses are likely to be very small).

All process units are in operation (i.e., normal operation) (hydraulic profile also should be calculated, assuming one process
unit is out of service, to check freeboard, etc.).

7.1.3. Calculations
Design flowrates:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Peak wastewater flow 0.175 m 3/s

RAS rate (maximum) 0.088 m 3/s

Peak wastewater flow + RAS 0.263 m 3/s

Water surface elevation (WS El) in the secondary clarifier:

Weir crest elevation = 518.262

Number of clarifiers = 2

Flow per clarifier = 0.175 m3 /s/2 = 0.088 m3 /s

Perimeter of weir = π × diameter = π × 20 m = 62.8 m

Number of 90-deg v-notches = 62.8 m/0.23 m center to center = 273

Flow per weir = (0.088 m3 /s)/273 = 0.00032 m3 /s

Head over the weir = (Q/1.38)1/2.5 = (0.00032 m3 /s/1.38)0.4 = 0.035 m from H = [Q/(C tan ϕ/2)]0.4, where Q = flow (m3 ), ϕ =
angle of the notch (deg), H = head over the crest (m), and C = weir coefficient (1.38 for a 90-deg weir)

WS El in secondary clarifier = 518.262 + 0.035 = 518.297

WS El in discharge side of mixed-liquor splitter box:

Select pipe size:

Number of clarifiers = 2

Flowrate per clarifier = (0.263 m3 /s)/2 = 0.132 m3 /s

Design for 0.75m/s

A = Q/V = (0.132 m3 /s)/0.75 m/s = 0.176 m2 , use 460-mm diameter, A = 0.167 m2

V = Q/A = (0.132 m3 /s)/0.167 m2 = 0.79 m/s

V2 /2g = (0.79 m/s)2 /(2 × 9.8 m/s2 ) = 0.032 m

Manning equation Q = (1/n) A R0.67 S 0.5; let K = (1/n) A R0.67, then S = (Q/K) 2

R = diameter/4 for pipes flowing full; R = (460 mm)/4 = 115 mm = 0.115 m

K = (1/0.015) × (0.167 m2 ) × (0.115 m)0.67

S = [(0.132)/2.614]2 = 0.0025 m/m

Pipe length = 30.5 m; friction loss = 30.5 m × 0.0025 m/m = 0.76 m

Minor losses, K: V2 /2g

Entrance loss, K = 0.5; 90-deg bend loss K = 0.4; exit loss K = 1.0

Total K = 0.5 + 2(0.4) + 1.0 = 2.3

Minor losses = 2.3 × 0.032 m = 0.074 m

WS El in discharge side of mixed-liquor splitter box = 518.297 + 0.076 m + 0.074 m = 518.447

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
WS El upstream of mixed-liquor splitter box weirs:

Design total flowrate = 0.263 m3 /s

Number of weirs = 2; one for each secondary clarifier

Flow per weir = (0.263 m3 /s)/2 = 0.132 m3 /s

Weir length 51.8 m; rectangular weir plate

Weir coefficient = 1.82

Head over the weir, H = (Q/C × L) 0.67; = [(0.132 m3 /s)/(1.82 × 1.8)]0.67 = 0.116 m

Weir crest El = 518.600

WS El upstream of mixed-liquor splitter box weirs = 518.600 + 0.116 m = 518.716

WS El in aeration basin mixed-liquor effluent channel:

Design total flowrate = 0.263 m3 /s

Select pipe size:

Design for 0.75 to 0.9 m/s

A = Q/V = (0.263 m3 /s)/0.75 m/s = 0.351 m2 , diameter = 0.68 m

A = Q/V = (0.263 m3 /s)/0.9 m/s = 0.292 m2 , diameter = 0.60 m

Use 600-mm diameter, A = 0.292 m2

V = Q/A = (0.263 m3 /s)/0.292 m2 = 0.90 m/s

V2 /2g = (0.90 m/s)2 /(2 × 9.8 m/s2 ) = 0.041 m

Manning equation Q = (1/n) A R0.67 S 0.5; let K = (1/n) A R0.67, then S = (Q/K) 2

R = diameter/4 for pipes flowing full; R = (600 mm)/4 = 150 mm = 0.150 m

K = (1/0.015) × (0.292 m2 ) × (0.15)0.67 = 5.46

S = [(0.263)/5.46]2 = 0.0023 ft/ft

Pipe length = 30.5 m; friction loss = 30.5 m × 0.0023 m/m = 0.070 m

Minor losses, K: V2 /2g

Entrance loss, K = 0.5; 90-deg bend loss K = 0.4; exit loss K = 1.0

Total K = 0.5 + 2(0.4) + 1.0 = 2.3

Minor losses = 2.3 × 0.041 m = 0.094 m

WS El in aeration basin mixed-liquor effluent channel = 518.716 + 0.070 m + 0.094 m = 518.880

WS El in aeration basin:

Design total flowrate = 0.263 m3 /s

Number of weirs = 2; one for each aeration basin

Flow per weir = (0.263 m3 /s)/2 = 0.132 m3 /s

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Weir length = 1.8 m; rectangular weir plate

Weir coefficient = 1.82

Head over the weir, H = (Q/C × L) 0.67; = [(0.132 m3 /s)/(1.82 ×1.8)]0.67 = 0.116 m

Weir crest El = 519.085

WS El in aeration basin = 519.085 + 0.116 m = 519.201

WS El in aeration basin influent channel:

Design total flowrate = 0.263 m3 /s

Number of gates = 4; 2 for each aeration basin

Flow per gate = (0.263 m3 /s)/4 = 0.066 m3 /s

Gate width = 380 mm

Gate height = 380 mm

Gate area, each = (0.38 m) × (0.38 m) = 0.144 m2

Gate orifice coefficient = 0.60

Gate head loss H = (1/2g) × (Q/C × A) 2 = [1/(2 × 9.8)] × [0.066/(0.6 × 0.114)]2 = 0.030 m

Check velocity through the gate V = Q/A = 0.066 m3 /s/0.144 m2 = 0.46 m/s OK

Head loss = 0.03 m; should give good flow distribution, as head loss in the influent channel itself likely will be very small.
Note that a weir or cut-throat flume could be used in lieu of the gate.

WS El in aeration basin influent channel = 519.201 + 0.030 m = 519.231

WS El in junction box:

Design total flowrate = 0.263 m3 /s; includes RAS at this point

Use 600-mm diameter pipe, as flow is same as from aeration basin

V = Q/A = (0.263 m3 /s)/0.292 m2 = 0.90 m/s

V2 /2g = (0.90 m/s)2 /(2 × 9.8 m/s2 ) = 0.041 m

Manning equation Q = (1/n) A R0.67 S 0.5; let K = (1/n) A R0.67, then S = (Q/K) 2

R = diameter/4 for pipes flowing full; R = (600 mm)/4 = 150 mm = 0.15 m

K = (1/0.015) × (0.292 m2 ) × (0.15 m)0.67 = 5.46

S = [(0.263)/5.46]2 = 0.0023 m/m

Pipe length = 30.5 m; Friction loss = 30.5 m × 0.0023 m/m = 0.070 m

Minor losses, K: V2 /2g

Entrance loss, K = 0.5; exit loss K = 1.0

Total K = 0.5 + 1.0 = 1.5

Minor losses = 1.5 × 0.041 m = 0.062 m

WS El in junction box = 519.231 + 0.070 m + 0.062 m = 519.363

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
WS El at primary clarifier effluent trough outlet:

Design total flowrate = 0.175 m3 /s

Number of primary clarifiers = 2

Flow to each primary clarifier = (0.175 m3 /s)/2 = 0.088 m3 /s

Select pipe size:

Design for 0.9 m/s

A = Q/V = (0.088 m3 /s)/0.9 m/s = 0.098 m2 , diameter = 350 mm

Use 300-mm diameter, A = 0.071 m2 (note that 350 mm could be used; designer preference)

V = Q/A = (0.088 m3 /s)/0.071 m2 = 1.23 m/s

V2 /2g = (1.23 m/s)2 /(2 × 9.8 m/s2 ) = 0.077 m

Manning equation Q = (1/n) A R0.67 S 0.5; let K = (1/n) A R0.67, then S = (Q/K) 2

R = diameter/4 for pipes flowing full; R = (300)/4 = 75 mm

K = (1/0.015) × (0.071 m2 ) × (0.075 m)0.67 = 0.83

S = [(0.088)/0.83]2 = 0.011 m/m

Pipe length = 30.5 m; friction loss = 30.5 m × 0.011 m/m = 0.336 m

Minor losses, K: V2 /2g

Entrance loss, K = 0.5; 90-deg bend loss K = 0.4; exit loss K = 1.0

For a junction box, assume total energy loss on entrance—a conservative assumption. As an alternative, the confluence
pressure and momentum equation could have been used (eq 3.30). Also, a slightly larger pipe diameter could have been
used to reduce the head loss.

Total K = 0.5 + 1(0.4) + 1.0 = 1.9

Minor losses = 1.9 × 0.077 m = 0.146 m

WS El at primary clarifier effluent trough outlet = 519.363 + 0.336 m + 0.146 m = 519.845

Maximum WS El in primary clarifier effluent channel:

Design total flowrate = 0.175 m3 /s

Number of primary clarifiers = 2

Flow to each primary clarifier = (0.175 m3 /s)/2 = 0.088 m3 /s


3
Flow splits two ways in each effluent channel in circular clarifiers = 0.088 m/s/2 = 0.044 m3 /s

Primary clarifier channel width = 460 mm

Critical depth = Dc = [Q2 /(B2 × g)]0.33 = {[(0.044)2 /(0.46) 2 ]/9.8}0.33 = 0.10 m

Minimum invert of primary clarifier effluent channel for free flow = 519.845 + 0.10 m = 519.945

Set weir trough invert elevation at 519.97 (allows for a small freefall at end of channel)

Maximum water depth in primary clarifier effluent channel = 1.73 ×Dc (because it is free flow)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Maximum water depth in primary clarifier effluent channel = 1.73 × 0.10 m = 0.173 m

Maximum WS El in primary clarifier effluent channel = 519.970 + 0.173 m = 520.143

Set primary clarifier effluent weir crest at 520.200 (allows for a small drop)

WS El in primary clarifiers:

Weir crest elevation = 520.200

Total design flow = 0.175 m3 /s

Number of clarifiers = 2

Flow per clarifier = 0.175 m3 /s/2 = 0.088 m3 /s

Perimeter of weir = π × diameter = π ×15 m = 47.1 m

Number of 90-deg v-notches = 47.1/0.23 center to center = 205

Flow per weir = (0.088 m3 /s)/205 = 0.00043 m3 /s

Head over the weir = (Q/1.38)1/2.5 = (0.00043 m3 /s/1.38)0.4 = 0.04 m (eq 3.29)

WS El in primary clarifier = 520.20 + 0.04 = 520.24

See Figure 3.20 for the completed hydraulic profile with calculated elevations.

Figure 3.20 Profile—design example—partial hydraulic profile with calculated elevations.

7.2. Design Example—Partial Hydraulic Profile in United States


Customary Units
The hydraulic profile for the primary and secondary processes for a WRRF should be based on the plan schematic shown
previously in Figure 3.18 and the profile shown in Figure 3.21. The invert elevation for the primary effluent channel and the
elevation of the notches for the primary clarifier effluent weir should be set.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 3.21 Profile—design example—partial hydraulic profile.

7.2.1. Input Parameters


Design peak flow = 4 mgd.

RAS = 50% of the influent wastewater flow and is returned to the junction box downstream of the primary clarifiers.

Primary and secondary clarifiers are circular with 90-deg, v-notches, 9 in. center to center; weir elevation shown is at the
bottom of the notch.

Primary clarifier effluent channel is rectangular in cross-section, 18 in. wide; the invert is horizontal (slope = zero).

Each aeration basin has an effluent weir, which is rectangular and 6 ft in crest length.

The mixed-liquor splitter box has two rectangular weirs, each 6 ft in crest length—one for each of the secondary clarifiers.

7.2.2. Assumptions
All weirs are sharp-crested; rectangular weir coefficient = 3.3.

Gate and orifice coefficient = 0.60; sharp, not rounded.

Ignore end contractions at weirs.

Manning's n = 0.015 for channels and pipes.

Entrance loss coefficient, K = 0.5; exit loss coefficient, K = 1.0; 90-deg elbow loss, K = 0.4.

Ignore losses in the aeration basin influent channel and mixed-liquor effluent channel (this is done for expediency here.
These losses are likely to be very small).

All process units are in operation (i.e., normal operation) (hydraulic profile also should be calculated, assuming one process
unit is out of service, to check freeboard, etc.).

7.2.3. Calculations
Design flowrates:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Peak wastewater flow 4 mgd (6.19 cu ft/s)

RAS rate (maximum) 2 mgd (3.09 cu ft/s)

Peak wastewater flow + RAS 6 mgd (9.28 cu ft/s)

Water surface elevation (WS El) in the secondary clarifier:

Weir crest elevation = 1699.90

Number of clarifiers = 2

Flow per clarifier = 6.19 cu ft/s/2 = 3.09 cu ft/s

Perimeter of weir = π × diameter = π × 65 ft = 204.2 ft

Number of 90-deg v-notches = 204.2/0.75 ft center to center = 272

Flow per weir = (3.09 cu ft/s)/272 = 0.0114 cu ft/s

Head over the weir = (Q/2.5)1/2.5 = (0.0114 cu ft/s/2.5)0.4 = 0.12 ft (eq 3.29)

WS El in secondary clarifier = 1699.90 + 0.12 = 1700.02

WS El in discharge side of mixed-liquor splitter box:

Select pipe size:

Number of clarifiers = 2

Flowrate per clarifier = (9.28 cu ft/s)/2 = 4.64 cu ft/s

Design for 2.5 ft/s

A = Q/V = (4.64 cu ft/s)/2.5 ft/s = 1.85 sq ft, use 18-in. diameter, A = 1.76 sq ft

V = Q/A = (4.64 cu ft/s)/1.76 sq ft = 2.63 ft/s

V2 /2g = (2.63 ft/s)2 /(2 × 32.2 ft/s2 ) = 0.11 ft

Manning equation Q = (1.49/n) A R0.67 S 0.5; let K = (1.49/n) A R0.67, then S = (Q/K) 2

R = diameter/4 for pipes flowing full; R = (18 in. × 1 ft/12 in.)/4 = 0.375 ft

K = (1.49/0.015) × (1.76 sq ft) × (0.375 ft)0.67 = 90.96

S = [(4.64)/90.96]2 = 0.0026 ft/ft

Pipe length = 100 ft; friction loss = 100 ft × 0.0026 ft/ft = 0.26 ft

Minor losses, K: V2 /2g

Entrance loss, K = 0.5; 90-deg bend loss K = 0.4; exit loss K = 1.0

Total K = 0.5 + 2(0.4) + 1.0 = 2.3

Minor losses = 2.3 × 0.11 ft = 0.25 ft

WS El in discharge side of mixed-liquor splitter box = 1700.02 + 0.26 ft + 0.25 ft = 1700.52

WS El upstream of mixed-liquor splitter box weirs:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Design total flowrate = 9.28 cu ft/s

Number of weirs = 2; one for each secondary clarifier

Flow per weir = (9.28 cu ft/s)/2 = 4.64 cu ft/s

Weir length = 6 ft; rectangular weir plate

Weir coefficient = 3.33

Head over the weir, H = (Q/C × L) 0.67; = [(4.64 cu ft/s)/(3.336)]0.67 = 0.38 ft

Weir crest El = 1701.00

WS El upstream of mixed-liquor splitter box weirs = 1701.00 + 0.38 ft = 1701.38

WS El in aeration basin mixed-liquor effluent channel:

Design total flowrate = 9.28 cu ft/s

Select pipe size:

Design for 2.5 to 3 ft/s

A = Q/V = (9.28 cu ft/s)/2.5 ft/s = 3.71 sq ft, diameter = 2.2 ft

A = Q/V = (9.28 cu ft/s)/3 ft/s = 3.09 sq ft, diameter = 2.0 ft

Use 24-in. diameter, A = 3.14 sq ft

V = Q/A = (9.28 cu ft/s)/3.14 sq ft = 2.96 ft/s

V2 /2g = (2.96 ft/s)2 /(2 × 32.2 ft/s2 ) = 0.14 ft

Manning equation Q = (1.49/n) A R0.67 S 0.5; let K = (1.49/n) A R0.67, then S = (Q/K) 2

R = diameter/4 for pipes flowing full; R = (24 in. × 1 ft/12 in.)/4 = 0.50 ft

K = (1.49/0.015) × (3.14 sq ft) × (0.50 ft)0.67 = 195.92

S = [(9.28)/195.92]2 = 0.0022 ft/ft

Pipe length = 100 ft; friction loss = 100 ft × 0.0022 ft/ft = 0.22 ft

Minor losses, K: V2 /2g

Entrance loss, K = 0.5; 90-deg bend loss K = 0.4; exit loss K = 1.0

Total K = 0.5 + 2(0.4) + 1.0 = 2.3

Minor losses = 2.3 × 0.14 ft = 0.32 ft

WS El in aeration basin mixed-liquor effluent channel = 1701.38 + 0.22 ft + 0.32 ft = 1701.92

WS El in aeration basin:

Design total flowrate = 9.28 cu ft/s

Number of weirs = 2; one for each aeration basin

Flow per weir = (9.28 cu ft/s)/2 = 4.64 cu ft/s

Weir length = 6 ft; rectangular weir plate

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Weir coefficient = 3.33

Head over the weir, H = (Q/C × L) 0.67; = [(4.64 cu ft/s)/(3.336)]0.67 = 0.38 ft

Weir crest El = 1702.60

WS El in aeration basin = 1702.60 + 0.38 ft = 1702.98

WS El in aeration basin influent channel:

Design total flowrate = 9.28 cu ft/s

Number of gates = 4; 2 for each aeration basin

Flow per gate = (9.28 cu ft/s)/4 = 2.32 cu ft/s

Gate width = 1.25 ft

Gate height = 1.25 ft

Gate area, each = (1.25 ft) = (1.25 ft) = 1.56 sq ft

Gate orifice coefficient = 0.60

Gate head loss H = (1/2g) × (Q/C × A) 2 = [1/(2 × 32.2)] × [2.32/(0.6 × 1.56)]2 = 0.10 ft

Check velocity through the gate V = Q/A = 2.32 cu ft/s/1.56 sq ft = 1.5 ft/s OK

Head loss = 0.10 ft should give good flow distribution, as head loss in the influent channel itself likely will be very small.
Note that a weir or cut-throat flume could be used in lieu of the gate.

WS El in aeration basin influent channel = 1702.98 + 0.10 = 1703.08

WS El in junction box:

Design total flowrate = 9.28 cu ft/s; includes RAS at this point

Use 24-in. diameter pipe, as flow is same as from aeration basin

V = Q/A = (9.28 cu ft/s)/3.14 sq ft = 2.96 ft/s

V2 /2g = (2.96 ft/s)2 /(2 × 32.2 ft/s2 ) = 0.14 ft

Manning equation Q = (1.49/n) A R0.67 S 0.5; let K = (1.49/n) A R0.67, then S = (Q/K) 2

R = diameter/4 for pipes flowing full; R = (24 in. × 1 ft/12 in.)/4 = 0.50 ft

K = (1.49/0.015) × (3.14 sq ft) × (0.50 ft)0.67 = 195.92

S = [(9.28)/195.92]2 = 0.0022 ft/ft

Pipe length = 100 ft; friction loss = 100 ft × 0.0022 ft/ft = 0.22 ft

Minor losses, K: V2 /2g

Entrance loss, K = 0.5; exit loss K = 1.0

Total K = 0.5 + 1.0 = 1.5

Minor losses = 1.5 × 0.14 ft = 0.20 ft

WS El in junction box = 1703.08 + 0.22 + 0.20 ft = 1703.50

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
WS El at primary clarifier effluent channel outlet:

Design total flowrate = 6.19 cu ft/s

Number of primary clarifiers = 2

Flow to each primary clarifier = (6.19 cu ft/s)/2 = 3.09 cu ft/s

Select pipe size:

Design for 3 ft/s

A = Q/V = (3.09 cu ft/s)/3 ft/s = 1.03 sq ft, diameter = 1.1 ft

Use 12-in. diameter, A = 0.79 sq ft

V = Q/A = (3.09 cu ft/s)/0.79 sq ft = 3.94 ft/s

V2 /2g = (3.94 ft/s)2 /(2 × 32.2 ft/s2 ) = 0.24 ft

Manning equation Q = (1.49/n) A R0.67 S 0.5; let K = (1.49/n) A R0.67, then S = (Q/K) 2

R = diameter/4 for pipes flowing full; R = (12 in. × 1 ft/12 in.)/4 = 0.25 ft

K = (1.49/0.015) × (0.79 sq ft) × (0.25 ft)0.67 = 30.85

S = [(9.28)/30.85]2 = 0.0101 ft/ft

Pipe length = 100 ft; friction loss = 100 ft × 0.0101 ft/ft =1.01 ft

Minor losses, K: V2 /2g

Entrance loss, K = 0.5; 90-deg bend loss K = 0.4; exit loss K = 1.0

For a junction box, assume total energy loss on entrance—a conservative assumption. As an alternative, the confluence
pressure and momentum equation could have been used (eq 3.30). Also, a slightly larger pipe diameter could have been
used to reduce the head loss.

Total K = 0.5 + 1(0.4) + 1.0 = 1.9

Minor losses = 1.9 × 0.24 ft = 0.46 ft

WS El at primary clarifier effluent channel outlet = 1703.50 + 1.01 ft + 0.46 ft = 1704.97

Maximum WS El in primary clarifier effluent channel:

Design total flowrate = 6.19 cu ft/s

Number of primary clarifiers = 2

Flow to each primary clarifier = (6.19 cu ft/s)/2

Flow splits two ways in each effluent channel in circular clarifiers = 3.09 cu ft/s/2 = 1.55 cu ft/s

Primary clarifier channel width = 1.5 ft

Critical depth = Dc = [Q2 /(B2 × g)]0.33 = {[(1.55)2 /(1.5)2 ]/32.2}0.33 = 0.32 ft

Minimum invert of primary clarifier effluent channel for free flow = 1704.97 + 0.32 ft = 1705.29

Set weir trough invert elevation at 1705.40 (provides a small drop to ensure free discharge)

Maximum water depth in primary clarifier effluent channel = 1.73 ×Dc (because it is free flow)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Maximum water depth in primary clarifier effluent channel = 1.73 × 0.32 ft = 0.55 ft

Maximum WS El in primary clarifier effluent channel = 1705.40 + 0.55 = 1705.95

Set primary clarifier effluent weir crest (notch elevation) at 1705.85 + 0.25 = 1706.20 (allows for 0.25-ft drop)

WS El in primary clarifiers:

Weir crest elevation = 1706.20

Total design flow = 6.19 cu ft/s

Number of clarifiers = 2

Flow per clarifier = 6.19 cu ft/s/2 = 3.09 cu ft/s

Perimeter of weir = π × diameter = π × 50 ft = 157.08 ft

Number of 90-deg v-notches = 157.08/0.75 ft center to center = 209

Flow per weir = (3.09 cu ft/s)/209 = 0.0148 cu ft/s

Head over the weir = (Q/2.5)1/2.5 = (0.0148 cu ft/s/2.5)0.4 = 0.13 ft from H = [Q/(C tan ϕ/2)]0.4, where Q = flow (cu ft/s), ϕ =
angle of the notch (deg), H = head over the crest (ft), and C = weir coefficient (2.5 for a 90-deg weir)

WS El in primary clarifier = 1706.20 + 0.13 = 1706.33

See Figure 3.22 for the completed hydraulic profile with calculated elevations.

Figure 3.22 Profile—design example—partial hydraulic profile with calculated elevations.

8.0. Conclusions
© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
8.0. Conclusions
Humans have an inexplicable affinity for water. There is something mighty, clean, and beautiful about water in nature, and we
are drawn to it, be it a small puddle in the street or the magnificent Niagara Falls. This is all very Darwinian, of course. Our very
survival depends on the availability of water for drinking, transport, power production, manufacturing, recreation, and yes, for
the disposal of wastes. Being able to understand how water behaves in the macro sense is one of the great pleasures of
engineering. Everyone knows that water flows downhill, but only engineers know how to make water run uphill. Hydraulic
engineering, the topic of this chapter, is fun. But it is also serious. In the design of WRRFs, hydraulics can make the difference
between success and failure. Overflowing clarifiers, pumps that do not have enough suction head, poorly operating distribution
boxes—these and thousands of other hydraulic elements in a WRRF—deserve careful engineering design and consideration.
Because of space limitations, this chapter covered only the bare essentials. There is much more serious business, and fun, to
be found in hydraulic engineering.

9.0. References
© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
9.0. References
American Society of Civil Engineers/Environmental and Water Resources Institute; Water Environment Federation (2007)
Gravity Sanitary Sewer Design and Construction, 2nd ed.; ASCE Manuals and Reports on Engineering Practice No. 60; WEF
Manual of Practice No. FD-5; American Society of Civil Engineers: Reston, Virginia.

Camp, T. R.; Graber, S. D. (1968) Dispersion Conduits. J. Sanit. Eng. Div., Proc. Am. Soc. Civ. Eng., 9 (1), 31–40.

Chow, V. T. (1959) Open-Channel Hydraulics; McGraw-Hill: New York.

Davis, C.; Sorensen, K. (1969)Handbook of Applied Hydraulics; McGraw-Hill: New York.

Environment Canada and Ontario Ministry of the Environment and Energy (1995) Guidance Manual for Sewage Treatment
Plant Process Audits.

Fair, G. M.; Geyer, J. Ch.; Okun, D. A. (1968) Water and Wastewater Engineering: Water Purification and Wastewater Treatment
and Disposal; Vol. 2; Wiley & Sons: New York.

Grace, R. A. (1978) Marine Outfall Systems: Planning, Design and Construction; Prentice-Hall: Upper Saddle River, New
Jersey.

Gunnerson, C. G.; French, J. A. (1996)Wastewater Management for Coastal Cities: The Ocean Disposal Option; Springer-
Verlag: New York.

Hydraulic Institute (1979) Hydraulic Institute Engineering Data Book; The Hydraulic Institute: Cleveland, Ohio.

Hydraulic Institute (1983) Hydraulic Institute Standards for Centrifugal, Rotary, and Reciprocating Pumps, 14th ed.; Hydraulic
Institute: Parsippany, New York.

Jones, G. M. (Ed.) (2006) Pumping Station Design, 3rd ed.; Butterworth-Heinemann: Burlington, Massachusetts.

King, H. W.; Brater, E. F. (1963) Handbook of Hydraulics; McGraw-Hill: New York.

Nolasco, D. (1999) Optimization, Debottlenecking, and Capacity Assessment using the Process Audit Methodology: A Dozen
Years of Canadian Experience. In: Plant Operations: Maximizing the Performance of Small to Medium Size Wastewater
Treatment Plants, WEF Specialty Conference, June 6–9, 1999, Milwaukee, Wisconsin.

Takács, I., Patry, G. G.; Nolasco, D. (1991). A generalized dynamic model for the thickening/clarification process.Water
Research, 25 (10), 1263–1271.

Water Environment Federation (2007) Manual of Practice MOP-11, Operation of Municipal Wastewater Treatment Plants.

Water Environment Federation; American Society of Civil Engineers; Environmental & Water Resources Institute (2009)
Design of Municipal Wastewater Treatment Plants, 5th ed.; WEF Manual of Practice No. 8; ASCE Manuals and Reports on
Engineering Practice No. 76; McGraw-Hill: New York.

10.0. Symbols Used in This Chapter


A = area of gate or port opening, m2 (sq ft)

BHP = brake horsepower, N-m/sec (lb ft/sec)

C = gate or orifice coefficient, dimensionless

C = Hazen–Williams friction coefficient

C = surface-roughness coefficient, dimensionless

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
C = weir coefficient

d = inside diameter of pipe, m (ft)

Δh = hydraulic grade differential along the channel or pipe, m (ft)

f = friction factor, dimensionless

g = acceleration due to gravity, 9.8 m/s2 (32.2 ft/sec2 )

γ = specific weight, N/m3 (in. lb/cu ft)

hL = friction head loss, m (ft)

hLg = head loss through gate, m (ft)

hLp = head loss through port, m (ft)

hLs = head loss through screen, m (ft)

H = head developed, N/m2 (lb/sq ft)

H = head over crest, m (ft)

K = a constant, function of the pipe roughness

k = head loss coefficient

L = pipe or channel, m (ft)

m = ratio of the lowest flow and highest flow through the respective ports

μ = fluid viscosity, kg-m/s (lb-sec/sq ft)

n = roughness coefficient, dimensionless

P = pressure, N/m2 (lb/sq in.)

φ = angle of the notch, deg

Q = flow, m3 /s (sq ft/sec)

R = hydraulic radius (cross-sectional area divided by wetted perimeter), m (ft)

ρ = fluid density, kg/m3 (lb sec2 /ft4 )

S = slope of the energy gradient, hL/L

v = velocity, m/s (ft/sec)

v = velocity upstream of bar screen, m/s (ft/sec)

vB = velocity through bar screen, m/s (ft/sec)

WHP = water horsepower, N-m/s (lb ft/sec)

Y = water depth, m (ft)

Z = baseline elevation (e.g., channel invert), m (ft)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.

You might also like