You are on page 1of 51

Source: HYDRAULIC DESIGN HANDBOOK

CHAPTER 21

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES


John A. Replogle and Albert J. Clemmens
USDAARS Water Conservation Laboratory Phoenix, Arizona

Clifford A. Pugh
US Bureau of Reclamation, Denver, Colorado

21.1 INTRODUCTION
Experienced water providers and users can use this chapter as a quick review of hydraulic principles related to water measurement and its relation to hydraulic design for environmental considerations. The hydraulic design of flow measuring structures usually confronts the engineer with two opportunities. One is the design of measurement structures in a retrofit situation and the other is in original project design. The retrofit mode is usually difficult and requires much innovation just to obtain passable function within the space and sizing limitations and other constraints usually imposed. Because of the increasing emphasis on quantifying flow rates and volumes in most aspects of water resource planning and management, the retrofit applications currently dominate the design problems. Most textbooks deal with recommending ideal installation situations and retrofit projects appear to be unable to comply without great economic impact. This too frequently can lead to arbitrary compromises that produce poor measurement performance. Even new installations may be limited by space requirements. This may force design decisions into the final construction that compromise accuracy. This chapter will strive to show the design concepts available, particularly those useful for designing both new and retrofit installations, and will point out measurement behaviors to be expected from various compromises. This chapter suggests those deviations that cause least impact and guides the designer to choices that may be hydraulically acceptable and still meet structural goals. Of the numerous flowmetering methods available to the hydraulic engineer, most are based on well-established hydraulic principles and are amenable to design manipulations of size, shape, and response. While this aspect of flow measurement is documented in several handbooks and texts, the design and retrofit of sites to accommodate and facilitate measurement is not as well described or is described in a scattered assortment of books and articles. Pipeline flows of water are usually less complicated to measure than open-channel flows, most obviously because the flow area does not change significantly with flow rate.
21.1 Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.2

Chapter Twenty-One

Consequently, many applications of pipeline flows are held to stricter accuracy standards than channel flows can reasonably achieve. Thus, channel flows and their measurement are usually limited to large delivery volumes and to accuracies acceptable to the related activities, such as sewer flows and irrigation deliveries. The purpose of this chapter is to consolidate design information for evaluating a flow measurement site, selecting a flow measuring system, and adapting the measuring site to optimize measuring and other functions that may be desired from the site. Emphasis will be on open-channel flow measurements because that is a likely need of the hydraulic engineer. Pipe flowmeters in water supply will also be discussed in lesser detail because the major application of the many types of pipe flowmeters is well covered in the chemical and petroleum industry literature. Experienced readers may wish to further investigate and seek more advanced references in hydraulics and fluid mechanics. Extensive information on fluid meter theory and detailed material for determining coefficients for tube-type meters is given in American Society, of Mechanical Engineers (ASME) (1959, 1971) and revisited with modern updates in books by Spitzer (1990) and Miller (1996). Brater and King (1982) have a thorough discussion of general critical depth relations and detailed relationships for most common hydraulic flow section shapes in open channels. Bos (1989) covers a broad segment of open channel water measurement devices.

21.2 HYDRAULIC CONCEPTS RELATED TO WATER MEASUREMENT


21.2.1 Basic Concepts for Pipe and Channel Flows Flow can be classified into closed conduit flow and open-channel flow. Open-channel flow conditions occur whenever the flowing stream has a free or unconstrained surface that is open to the atmosphere. Flows in canals or in vented pipelines that are not flowing full are typical examples. In hydraulics, a pipe is any closed conduit that carries water under pressure. The filled conduit may be square, rectangular, or any other shape, but is usually round. If flow is occurring in a conduit but does not completely fill it, the flow is not considered pipe or closed conduit flow, but is classified as open-channel flow. Flow rate in a pipeline responds mainly to the pressure gradient or head difference that exists between two points along the pipeline, modified by the frictional resistance to flow caused by pipe length, pipe roughness, bends, restrictions, changes in conduit shape and size, the nature of the fluid flowing, and the cross-sectional area of the pipe. In open-channel flows, the pressure gradient, or energy grade line, is controlled mainly by the force due to gravity, which is influenced by the channel slope, resistance from the channel wall roughness, the channel shape, and the flow area. The fluid is usually water. Basic flow metering in both pipe flow and open channels depends on determining an average flow velocity by some means and combining it with the flow cross-sectional area. For open channels, a common means involves current meter measurements where metered point velocities are applied to their applicable subareas and summed over a flow cross section. Exceptions include tracer-dilution techniques that do not require flow area or velocity. The uses of tracer techniques are applicable to special pump calibrations and some difficult channel flows (mountain streams). They are avoided for most city water distribution systems, sewer flows and irrigation applications because of the general expense with han-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.3

dling the equipment and doing the analysis. The most used techniques applicable to openchannel systems, including sewer flows and irrigation canal flow measurements, depend on exploiting the special velocity properties of critical flow, as discussed in a section 21.2.3. Continuity equation. The first basic equation for water flowing in either pipes or channels is the continuity equation, which simply states that discharge rate (volumetric flow per unit time), Q, is equal to flow cross-sectional area, A, times flow mean velocity, V, V through the flow cross section, or Q AV (21.1)

Bernoulli energy equation. Another basic equation involves energy relations and is also applicable to both pipe and channel flows. The most familiar form is for closed pipe flow, wherein the basic energy principles are described by the Bernoulli energy equation. For two locations along a pipe at stations 1 and 2,(Fig. 21.1), the Bernoulli equation can be expressed as z1 h1
2 V1 2g

z2

h2

2 V2 2g

constant

(21.2)

where the terms are expressed in length dimensions as z the height from an arbitrary reference plane (datum) h the pressure head V average velocity through the pipe cross-section at the designated location V2/2g the velocity head g the gravitational constant 1 ,2 subscripts denoting the respective locations along the pipeline. This equation is based on uniform velocity across the conduit area and no energy losses. However, in real fluid flows, nonuniform velocities exist and friction causes energy conversion to heat. Typically, these velocities are zero at the walls and reach a maximum profile velocity near the center of the flow. If the flow is viscous flow in a round pipe, the flow profile is parabolic, that is, bullet-shaped. If the velocity is fully turbulent, the bullet-shape is much flattened, with steep velocity gradients near the wall and nearly uniform profile across the remainder of the pipe. These idealized profiles can be skewed drastically by regulating valves, structures, conduit bends and other flow obstructions. Therefore, application of these equations depends on knowing, or controlling, the velocity profile so that the average velocity in the conduit cross section can be inferred.

FIGURE 21.1 Energy balance in pipe flow.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.4

Chapter Twenty-One

Equation (2.2) requires some adjustments to convert it to the energy equation, which is useful in analyzing flows in pipes or open channels with a small slope (Chow, 1959). First we introduce correction factors, 1 and 2, called the velocity distribution coefficients, to account for the computational expediency of using the average velocities, V1 and V2, to compute the kinetic energy term, V2/2g, at the respective locations 1 and 2 along a channel. These values for the usual range of turbulent flows in water usually range from about 1.01 to 1.05, although for thick petroleum products in pipe flows and low velocity flows, the value can approach a value of 2. Second, a term, hf, for the loss of energy between the two points is included. The result is z1 21.2.2 Pipe Hydraulics Reynolds number. The behavior of flow in pipes is governed primarily by the viscosity of the fluid. In pipeline flows, the ratio between the dynamic forces and the viscus forces is important for defining the limits between laminar and turbulent flows and other functions of pipe flow. This ratio is called the Reynolds number, Rn, and is defined as Rn where V the velocity of the flow, Lc ter, D and v the kinematic viscosity. VLc (21.4) v characteristic length, typically the pipe diameh1
1 2 V1 2g

z2

h2

2 V2 2g

hf

(21.3)

Headloss characteristics in pipes. The Reynolds number, Rn, defined above, represents the effect of viscosity relative to inertia and is used to define appropriate flow ranges for headloss equations in pipe flow. For example, headloss is proportional to the square of the velocity, when the velocities and pipe size combinations defined by a pipe-diameterbased Reynolds number, Rn, greater than about 1000. Most of the flows of interest in general hydraulic engineering have Reynolds numbers greater than 1000. Some exceptions are found in drip or trickle irrigation systems common in agricultural and urban landscape settings. The headloss, hf, for Rn greater than the minimum value of about 1000 is traditionally expressed in terms of a friction factor, f, the pipe diameter, D, pipe length, L, and the velocf ity head, V2/2g, where g is the gravitational constant, and V is the average velocity, as hf
2 f L V D 2g

(21.5)

The value for f is usually obtained from a Moody diagram which is a graphical representation of the f value in terms of the Reynolds number, the roughness height of the pipe wall material, , and the pipe diameter, D. The Moody diagram is a graphical solution of the Colebrook function 1 f 2log /D 3.7

2.51 Rn f

(21.6)

The values range from 0.0000015 m for smooth plastic pipe to 0.00026 m for cast iron pipe. Concrete pipe ranges from about 0.0003 m to 0.003 m (Daugherty and Ingersoll, 1954). The equation can be readily solved by iteration techniques using a computer spreadsheet.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.5

21.2.3 Channel Hydraulics Hydraulic mean depth. The hydraulic mean depth, Dm [U.S. Bureau of Reclamation (USBR), 1997] is the flow cross-sectional area, A, divided by the flow surface width, T, or T Dm A T (21.7)

For conduits such as pipes flowing nearly full, the surface flow width may be narrow, and Dm may be a larger value than the physical water depth. For the usual natural channels and most canals, Dm is interchangeable with average depth. Sometimes it is simply called the hydraulic depth (Chow, 1959). Froude number. Open-channel flow behavior is governed primarily by gravity forces. The ratio of the inertial forces to the gravity forces is called the Froude number, Fr, and is defined by Fr where V the velocity of the flow, g mean depth. V gDm the gravitational constant, Dm (21.8) the hydraulic

The Froude number applies to most open channel flows and is used for defining model scale ratios and estimating stable flow characteristics in open channels. Specific energy. It is useful to define the energy equation in terms of the local channel bottom instead of an arbitrary datum. This is called the specific energy, E, and is given by: V2 (21.9) 2g That is, the specific energy is equal to the sum of the depth of flow y and the velocity head (Fig. 21.2). E y Critical flow and critical depth. In open channels a flow phenomenon occurs that does not happen in closed pipe flows. The process is called critical flow. Critical flow is defined for open channel flows as the maximum discharge for the minimum specific energy, that is, critical flow represents the minimum combination of potential energy (depth of flow, y) and kinetic energy (velocity head, V2/2g) for the given discharge (Chow, 1959). The depth of flow then is the critical depth. By virtue of the continuity equation, for a constant discharge at critical flow, an increase in depth must necessarily be accompanied by a decrease in velocity, which is called subcritical velocity. Conversely, a decrease in depth for the same flow rate necessarily requires an increase in velocity, which is called supercritical velocity. When critical flow occurs in an open channel it can be shown (Chow, 1959) that V2 Dm c 2g 2 where Vc mean flow velocity, g gravitational constant, Dm and velocity distribution coefficient (Chow, 1959). (21.10) hydraulic mean depth,

This can further be combined with the continuity equation, Eq. (21.4), to express the critical flow discharge rate, Qc, as

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.6

Chapter Twenty-One

FIGURE 21.2 Specific energy balance.

Qc

Dm g

(21.11)

In practice, the water surface slope in a contraction is relatively steep and the precise plane of the critical flow section is not easily or reliably located. Thus, the data for accurately evaluating the hydraulic depth, Dm, is not readily obtained. For critical flow flumes, the flow depth is therefore not measured at this critical section, but instead a depth is measured in the upstream channel, where the velocity head is computable or is minimal. The critical depth is then mathematically derived based on energy principles described by Bos et al. (1991). These flumes, sometimes referred to as the computable flumes that rely on critical flow theory, will be discussed in more detail in Section 21.7. For maximum discharge for minimum energy, the condition described above for critical flow in open channels, it can be shown that Vc2 2g A 2T Dm 2 (21.12)

where: A = the channel flow area, T = the top width of the channel flow, Dm = the hydraulic mean depth, and Vc = the critical velocity. Thus, the velocity head at critical flow is equal to half the hydraulic mean depth, sometimes called hydraulic depth, Dm A/T (Chow, 1959). From the above, Vc 1 Fr (21.13) gD/ where Fr is the Froude number defined above. Thus, at critical flow the Froude number is unity. Also note that the Froude number can be defined by Fr V/Vc for velocities other V than critical. Normal depth. Yet another depth is associated with open channel flows, the normal depth. When the flow in an open channel does not change from station to station, the flow is said to be uniform and the bottom slope, the hydraulic grade line, and the energy grade line are all parallel to each other. Figure 21.2 shows the condition when the flow is not uniform. Modular limit. If the downstream depth in a channel is too deep, the backwater will prevent critical depth from occurring. The flow is considered to be submerged whenever the downstream water surface exceeds the crest elevation of a channel control, such as a

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.7

weir or flume. For flumes, particularly, this submergence has little effect on critical depth, and free flow exists until a certain limiting submergence for that particular flow module called the modular limit is reached. At some point of submergence, the upstream flow depth is affected. and the modular limit is exceeded, and free flow does not occur. The modular limit is defined as that limiting submergence ratio, and is based on the ratio of the downstream depth to upstream depth. The modular limit occurs when the downstream backwater causes more than1 percent change in the calculated discharge in a particular flow module, or device (Bos, 1989). When the modular limit is exceeded, the flow is called nonmodular. 21.2.4 Energy Balance Relationships in Channels Hydraulic problems concerning fluid flow are commonly described in terms of conserving kinetic and potential energy, and are conveniently expressed using the classical Bernoulli equation in combination with the Continuity equation. The applications of these equations are generally well documented, particularly for pipe flows, in texts and handbooks and are not repeated here (Brater and King, 1982; Miller, 1996). The case for open channels is less complete, but is given considerable treatment in Brater and King (1982), Chow (1959), and Herschy (1985). The computational uncertainties evolve from the effects of friction and viscosity that distort the classic assumptions of a uniform velocity profile across the fluid stream. When accountings for friction and flow profile are successfully applied, the results for discharge computations are usually good to excellent for both pipes and open channels (Bos et al., 1991). Headloss characteristics in channels. In terms of frictional headlosses, the wetted perimeter, Pw, of the flow is important. Hydraulic radius, Rh, is defined as the area of the flow section, A, divided by the wetted perimeter, Pw, or A (21.14) Pw Conversely, the wetted perimeter times the hydraulic radius is equal to the area of an irregular flow section. The hydraulic radius of a channel can be compared to the radius of a pipe, r, with a cross-sectional area A r2 and a circumference or wetted perimeter Pw r 2 r. Under these conditions, the hydraulic radius compares to the pipe radius, and to the pipe diameter, D, as r D Rh (21.15) 2 4 The Mannings formula. Canal and stream discharge rates are usually estimated with use of the Mannings formula. Many open-channel flow equations have been proposed, but the most used is the Mannings formula. This expression is partly rational and uses an empirical coefficient, n, that is used in both the SI and American unit systems. In general form it becomes Cm 2/3 1/2 V R S (21.16) n h e where V average velocity, n the Mannings roughness coefficient, Rh the hydraulic radius, Se the energy-line slope, and Cm conversion of units: 1.0 for metric units and 1.486 for American units. Rh The factor Se is the slope of the energy line. Note that the bed slope of the channel, So, and the slope of the water surface, Sw, are not to be used. These parameters are, however, equal to Se when uniform flow, with the resulting normal depth occurs. As defined above,
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.8

Chapter Twenty-One

normal depth occurs when a channel flow approaches uniformity from station to station along the channel (Chow, 1959). For design purposes, the n value for concrete lined canals is usually about 0.014. A good finish can lower it to 0.012, while concrete in poor condition and channels constructed with shot crete or gunite, usually have n values from 0.016 to 0.018. In some instances, concrete lined canals, with significant algae growth, have experienced n values as high as 0.032. This latter value approaches the values usually experienced with unlined channels, 0.030.04. Thus, for reliable application, the use of Mannings formula requires field experience and on-site inspection of the channel being computed.

21.2.5 Modeling Characteristics for Open Channels For flowing water in open channels, fluid friction is a factor as well as gravity and inertia. This would seem to present a problem for hydraulic scale modeling, because both dynamic and kinematic similarity are difficult to achieve simultaneously. Fortunately for most open-channel flows, there is usually fully developed turbulence. Thus, the fluid friction losses are nearly proportional to V 2, and are nearly independent of Reynolds number, Rn, with rare exceptions. This means that in openchannel flows, inertia and gravity forces dominate over viscous forces (associated with pipe flows) and are a function of the Froude number, Fn, alone. Geometric similarity between a model and a prototype then provides kinematic similarity. For kinematic similarity the ratios of the respective velocities are everywhere the same. The velocity ratio, Vr, is the velocity in the prototype, Vp, divided by the velocity in the model, Vm, or Vp Vr (21.17) Vm For Froude modeling, and from the definition for Fn, we note that V is proportional to the square-root of a length, L (for open channels we used the hydraulic depth, Dm) with the gravitational constant, g, assumed to be constant. Thus, the above equation can be written as Vp Lr (21.18) Vm where Lr = the length ratio between prototype and model dimensions, Lp: Lm Because the velocity varies as that Lr and the cross-sectional area as Lr2 it follows Qp : Qm or Qp Qm
Lp 5/2 Lm

L5/2 : 1 r

(21.19)

(21.20)

This equation is valid when all the physical structure dimensions and the heads are of the same ratio. For example, it can be used to convert a flume rating for one size to that

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.9

of a similar flume of another size. Scale modeling works best for determining calibrations in a range of Lp:Lm less than about 10:1, although ranges exceeding 50:1 have sometimes been used for studying special situations.

21.3 BASIC PRINCIPLES OF WATER MEASUREMENT


Flow is usually measured by determining an average flow velocity and using the flow area to compute the volume discharge. Flow meters then have the function of detecting this velocity and combining it with the physical information of the conduit to produce a useable readout. This is easily demonstrated for closed conduits. Propeller meters, ultrasonic meters, laser-Doppler velocimeters, electromagnetic meters, Venturi meters, and orifice meters all are based on inferring a basic velocity measurement applied to a flow area for a discharge rate. For open channels, many flumes depend on determining the velocity based on energy principles of critical flow. Weirs are usually described in terms of orifice flow integrated over the weir width and the crest depth. Again these are basically velocity expressions for flow through a defined area. Dilution techniques applicable to both closed pipe and open channel flows depend on detecting the amount of fluid added to a known starting amount of tracer material. The dilution ratio determines the discharge ratio, in the case of constant injection of a tracer. The tracer may be a chemical or even injected heat or heated fluid. Electromagnetic meters depend on generating voltages by flowing a conductive fluid, usually water, through a magnetic field to produce a velocity indication.

21.3.1 Water Meter Classification Flow measuring devices are commonly classified into those that are rate meters and measure discharge rate as the primary reported indication and those that are quantity meters and measure volume as the primary indication. The latter include weighing tanks and batch volume tanks and are used mostly in laboratory settings as flow rate standards. Devices in either of these broad classes can again be divided according to the physical principle that is used to detect that primary indication (ASME, 1959). The meter part that interacts with the flow to produce the primary indication is referred to as the primary device. This interaction exploits one or more of a few physical principles, such as pressure force, energy conversion, weight, electrical properties, mixing properties, sonic properties, and so on, to generate a signal. Primary devices are thus limited in number and variety. Secondary devices convert the primary interaction into useable readout. These secondary devices are numerous and relatively unlimited in configuration and variety. The function of one class can be converted into the response of the other with suitable secondary devices. Some water measuring devices particularly suitable to municipal water supply, wastewater treatment, agricultural irrigation, and drainage applications are the historical rate meters that are treated in most hydraulic text. These include (1) weirs, (2) flumes, (3) orifice meters, and (4) Venturi meters. Head, h, or upstream depth, commonly is used for the open channel devices such as flumes and weirs. Either pressure, p, head, h, or differential head, h, or differential pressure, p, is used with tube-type devices, such as Venturi meters and orifice meters.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.10

Chapter Twenty-One

Venturi meters in pipelines and long throated flumes in open-channel flows are examples where the energy principles and the flow accountings mentioned above give good to excellent computational results with minor dependency on empirical coefficients (Bos, 1989; Bos et al., 1991).

21.3.2 Installation Requirements Special difficulties arise in applying velocity profile and friction accountings when insufficient pipe or channel exists upstream from a flow measuring device. This is needed to ensure that predictable and acceptable velocity profiles are presented to the meter. Frequently pipe or channel lengths can be significantly shortened by special structural flow conditioners. These structural measures then become a design option. Some of these are discussed below and in section 21.3.3 Designs for pipe discharges are well described in textbooks and in standard handbooks. The design difficulties center around selecting appropriate metering candidates for accomplishing the measuring function and in providing an appropriate environment for economical, accurate, and serviceable operation. In the case of pipe flows, recommended straight pipe lengths, in terms of pipe diameter, are to be provided upstream of the meter to assure reasonable operating accuracy. These lengths depend on the flow pattern presented to the meter primarily caused by valves and pipe elbows upstream from the meter. The number and orientation of elbows greatly influence the circulation patterns and flow profile distortions presented to the meter. Open-channel flow water measurement generally requires that the Froude number of the approach flow be less than 0.5 to prevent wave action that would hinder or possibly prevent an accurate head determination. Energy concepts are used to describe Venturi meters in pipe flows based on the Bernoulli equation in which part of the pipe forms a contracted throat that necessarily changes the flow velocity and hence converts some of the static pressure to velocity head. The decrease in static pressure is the basis for flow detection. A similar concept can be applied to open channels. A historical version is the socalled Venturi flume (Brater and King, 1982) that detects the change in water surface elevation between an upstream station and in a contracted section. However, this small change is difficult to accurately detect, so the direct concept is not used. Rather, contractions are designed to be severe enough to force critical flow velocities in the contracted section. Thus, only an upstream head is needed to define the flow energy and flow area which can be converted to discharge rate. These are generally called critical-flow flumes. The flow condition where only one head measurement is needed is called free flow. The critical-flow flumes themselves consist of those called long-throated flumes that force parallel flow in the contracted, or control, section, called the throat, and those that have curvilinear flow in the throat and are called short-throated flumes. The limiting throat control section is the sharp-crested weir consisting of a thin plate. Thus, for flumes and weirs one unique head value exists for each discharge, simplifying the calibration procedure. However, if the downstream flow level submerges critical depth enough to affect the upstream reading, the modular limit is exceeded, and free flow does not occur. When exceeded, separate calibrations at many levels of submergence are then required, and two head measurements are needed to measure flow. This condition generally is to be avoided in meter site design because it reduces the accuracy of the measurement and increases the difficulty of flow determination. The modular limit for sharp-crested weirs, in practice, is less than zero, requiring full clearance of the overfall nappe of at least 3 cm, while short-throated flumes can usually tolerate 65 percent to 70 percent submergence. LongDownloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.11

throated flumes can tolerate from 70 percent to 90 percent depending on flow conditions and flume size. Designing flumes for submerged flow beyond the modular limit decreases the accuracy of the flow measurement. Sometimes flumes and weirs can be overly submerged unintentionally by poor design, construction errors, structural settling, attempts to supply increased delivery needs with increasing downstream heads, accumulated sediment deposits, or weed growths. Sometimes use of the submerged range beyond the modular limit is an economic compromise. Approach flow conditions for pipes. Water measurement devices are generally calibrated with certain approach flow conditions. The same approach conditions must be attained in field applications of measuring devices. Poor flow conditions in the area just upstream of the measuring device can cause large discharge indication errors. For open channels, the approaching flow should generally be subcritical. The flow should be fully developed, mild in slope, and free of curves, projections, and waves. Pipeline meters commonly require 10 or more diameters of straight pipe approach. Fittings and combinations of fittings, such as valves and bends, located upstream from a flow meter can increase the number of required approach diameters. Several references (ASME, 1971; ISO, 1991) give requirements for many pipeline configurations and meters. These are discussed in detail by Miller (1996). Flow conditioning options. Many installations, especially in retrofit situations, do not provide for sufficient lengths of straight pipe to remove velocity profile distortions and swirl to an acceptable level. Therefore, the designer may need to use flow conditioners in combination with straight pipe lengths. Swirl sensitivity varies widely. Some meters are particularly sensitive to swirl, such as the propeller and turbine meters. Magnetic flow meters are somewhat less sensitive to radial velocities than single-path ultrasonic flow meters. Venturi meters are less sensitive than orifice meters. For a swirl angle of 20 , the discharge coefficient changes by about 1 percent for a Venturi meter with 0.32 ( is the ratio of meter throat diameter to the pipe diameter) and about 10 percent for a similar orifice. Thus a swirl can increase the discharge through an orifice for the same differential head reading (Miller, 1996). In pipeline flows, contractions can produce a central jet and also increase an incoming swirl, while expansions tend to slow swirls and produce enough secondary flow to restore flow profiles to some semblance of acceptability. These characteristics can modify the straight pipe lengths needed or the type of flow conditioner to recommend (Miller, 1996). Rough pipes also tend to reduce a swirl. For flows, such as that encountered in sewage discharges and irrigation pipeline deliveries that originate from open channels, many of the tube-bundle types of flow conditioners can gather trash and cause maintenance problems. Many meter providers in these situations use fins or vanes that protrude from the wall and have sloped upstream edges that shed trash. The vanes protrude about one-fourth of the pipe diameter into the flow, leaving the center core of the flow open. While these vanes can vary in number and length, the logic being that the fewer the vanes the longer they should be in the direction of flow, common configurations are four vanes that are about two or three pipe diameters long. Vanes in themselves do not condition wall jets well. Field experience, has shown that troublesome flow profiles can be conditioned significantly by inserting an orifice into the pipe. The orifice diameter is about 90 percent of the pipe diameter and is used to control wall jets and force them to mix with the general flow. The orifice in itself tends to cross-mix

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.12

Chapter Twenty-One

the jets and would appear to reduce spin. However, if the jets are symmetrical and an initial swirl exists, orifices tend to increase the swirl. Inserting an orifice appears to be supported by recent recommendations of Miller (1996) where it is stated: To achieve a fully developed profile, it is important that the flow be blocked or restricted close to the wall, with the central core having the larger flow area. The addition of vanes when space permits is recommended. Because orifices, in general, tend to force the flow to the pipe center while increasing spin, it appears best to place the vanes upstream from the orifice. If they are placed downstream, the spin not only may be increased, but the spinning central flow may not be touched by the vanes.

21.3.3 Examples of Flow Conditioning in Field Situations Flow conditioning in an irrigation delivery pipeline. As mentioned previously, measuring devices frequently must be installed in flow situations that are less than optimal. A field example occurred in Arizona where a large pipe was used as an outlet to a secondary canal and a single-path ultrasonic meter placed in it was subjected to flow profile distortions. The pipe was about 0.75 m in diameter and delivered approximately 400 L/s. The flow rate readout was unstable, with fluctuations varying by about 15 percent. The problem appeared to be caused by slowly spiraling flow induced by the bottom jet from a partly open pipe o inlet gate and a 45 elbow. This is similar to two closely spaced pipe elbows that are not in the same plane, which can cause a spiral flow pattern (ASME, 1971). A successful attempt to modify the jet and cause it to cross mix so that the jet effects and the strength of the spiral flow were reduced, was accomplished by inserting a large ratio orifice in the pipe (Fig. 21.3). This consisted of an annular metal ring with the outside radius approximately that of the pipe and an inside diameter about 10 percent less, or an orifice with 90 percent. The orifice was installed about three diameters downstream from the elbow. The slight increase in headloss was compensated by increasing the upstream gate opening. The orifice can be constructed by cutting notches from an appropriately sized piece of angle iron or aluminum and bending it to a polygon that approximates the circle diameter of the pipe interior. Some leakage around the ring is acceptable. For propeller meters, additional vanes projecting from the walls may be needed to further

FIGURE 21.3 An orifice plate with a large opening is used to condition a flow profile.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.13

reduce spiral flow. These vanes would be placed upstream from the orifice. In this installation, the fluctuation was reduced to within about 3 percent. Flow conditioning in channels. By analogy and using a minimum of 10 pipe diameters of a straight approach channel, open channel flow would require 40 hydraulic radii of straight, unobstructed, unaltered approach, based on the calculation of hydraulic radius for circular pipes being equal to one-fourth the pipe diameter, (Eq. 21.15). This would translate for very wide channels into approximately 40 times the flow depth. For narrow channels that are as deep as they are wide, this would compute to be about 13 channel depths or top widths. Other recommendations on approach channel criteria are presented by Bos (1989) and USBR (1997). Major features of that criteria follow: If the control width is greater than 50 percent of the approach channel width, 10 average approach flow widths of straight, unobstructed approach are required. If the control width is less than 50 percent of the approach width, 20 control widths of straight, unobstructed approach are required. If upstream flow is below critical depth, a jump should be forced to occur. In this case, 30 measuring heads of straight, unobstructed approach after the jump should be provided. If baffles are used to correct and smooth out approach flow, then 10 measuring heads (10 h1) should be placed between the baffles and the measuring station. Approach flow conditions should be continually checked for deviation from these conditions as described in Bos (1989) and USBR (1997). The baffles described above can become unacceptable maintenance problems in open channels. Some field expediencies are therefore described that have been found to work in specific instances, but have not been studied for assured design generalizations. Nevertheless, these constructions are but small extensions to currently accepted practices in pipe flows. Applications for open channel flow conditioners include abrupt channel turns, sluice gate outflows, and channels downstream from a hydraulic jump. The abrupt turns may benefit from floor and wall mounted vanes or fins. Based on pipe flow experience, and assuming the channel is half of a closed conduit, these fins or vanes would probably be about 10 percent to 15 percent of the channel depth. As in pipe flow, wall jets that can develop downstream from sluice gates appear to need treatment. This can be in the form of a structural angle bolted on the channel floor and up the walls. Suggested size, based on the pipe flow analogy, is for the angle to be about 5 percent into the channel flow depth. Whether the sidewalls need larger angles when the channels are wide has not been tested.

21.3.4 Wave Suppression Of special concern in open channels is wave suppression downstream from a sluice gate, hydraulic jump, or an abrupt turn. Thus, the flow conditioners in channels have the additional task not present in pipe flows of surface wave suppression. Excessive waves in irrigation canals make reading sidewall gages difficult. These waves are usually caused by a jet entry from a sluice gate or by a waterfall situation. The unstable surface can be 1020 cm high and extend for tens of meters downstream. Wave suppression in canals. A surface wave suppressor was tested by Schuster as reported in USBR (1997). It basically was a constructed roof over the canal for a distance
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.14

Chapter Twenty-One

FIGURE 21.4 Wave suppresor design (From USBR, 1997).

equal to about four times the flow depth. The roof structure is inserted into the flow about one-third the flow depth. All flow is forced to pass under the structure. Wave suppression is between 60 percent and 93 percent (Fig. 21.4). For canals that usually flow at one level, this wave-suppression method is appropriate. The wave suppressor shown in Figure 21.4 has been successfully used in both large and small channels (USBR, 1997). An important aspect is that the structure is fixed and not allowed to float. Floating suppressors are not effective. Successful field applications of wave suppressors include some installations in trapezoidal irrigation channels, with 1:1 side slopes and 60-cm bottom width. They were flowing about 400 L/S at about 45 cm deep. While the velocity was not high, about 0.8 m/s, the agitation from a flow entry gate was producing waves about 15 cm high. The suppressor roof was only about 60 cm in the direction of flow, and penetrated the flow by about 15 percent. Another version that has worked in small channels is illustrated in Figure 21.5. This can work with a single crossmember if the flow is usually at a fixed discharge rate and becomes similar to the suppressor described above. In severe jet cases an additional floor sill, about 10 percent of the flow depth in height, has been used successfully. The length of the roof in the flow direction has not been well studied, but field observations seem to support a length greater than two lengths of the surface wave, if that can be estimated, otherwise, use two to four times the maximum flow depth as described above. To suppress waves in canals that do not always flow at the same depth, a staggered set of baffles may help (Replogle, 1997). Because these will be submerged part of the time, they must have a thickness that overlaps slightly to accommodate the vertical depth of interest. To avoid obstructing the channel severely, these baffles probably should not obstruct more than about 20 percent of the channel at any particular location. Staggering them as shown in Fig. 21.5 would accomplish this without excessive obstruction. Rounding the upstream edges will help shed trash, but may be less effective in suppress-

FIGURE 21.5 Wave suppressor for variable-depth flows in a canal. (From Replogle, 1997)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.15

ing waves. Observe in the sequence of drawings in Fig. 21.5 that the staggering is upward in the downstream direction. Note that the next baffle slightly overlaps the horizontal flow lines so that flow passing over the top of one baffle is not allowed to free-fall and start another wave. Fig. 21.5 ac illustrate, the general behavior as the flow becomes less deep.

21.4 MEASUREMENT ACCURACY


Accurate application of water measuring devices generally depends upon standard designs or careful selection of devices, careful fabrication and installation, good calibration data and adequate analysis. Also needed is proper user operation with appropriate inspection and maintenance procedures. During operation, accuracy requires continual verification that all measuring systems, including the operators, are functioning properly. Thus, good training and supervision are required to attain measurements within prescribed accuracy bounds. Accuracy is the degree of conformance of a measurement to a standard or true value. The standards are selected by users, providers, governments, or compacts between these entities. All parts of a measuring system, including the user, need to be considered in accessing the system's total accuracy. As mentioned above, a measurement system usually consists of a primary element, which is that part of the system that creates what is sensed, and is measured by a secondary element. For example, weirs and flumes are primary elements. A staff gage is a secondary element. Designers, purchasers, and users of water measurement devices generally rely on standard designs and manufacturers to provide calibrations and assurances of accuracy. A few water users and providers have the facilities to check the condition and accuracy of flow measuring devices. These facilities have comparison flow meters and/or volumetric tanks for checking their flow meters. These test systems are used to check devices for compliance with specification and to determine maintenance needs. However, maintaining facilities such as these is not generally practical. Various disciplines and organizations do not fully agree on some of the definitions related to measuring device specifications, calibration, and error analysis. Therefore, it is important to verify that a clear and mutual understanding of the specifications, calibration terminology, and the error analysis processes is established when discussing these topics with others.

21.4.1 Definitions of Terms Related to Accuracy Error. Error is the deviation of a measurement, observation, or calculation from the truth. The deviation can be small and inherent in the structure and functioning of the system and be within the bounds or limits specified. Lack of care and mistakes during fabrication, installation, and use can often cause large errors well outside expected performance bounds. Because the true value is seldom known, some investigators prefer to use the term uncertainty. Uncertainty describes the possible error or range of error which may exist. Investigators often classify errors and uncertainties into spurious, systematic, and random types. Precision. Precision is the ability to produce the same measurement value within given accuracy bounds when successive readings of a specific quantity are measured. Precision represents the maximum departure of all readings from the mean value of the readings. Thus, a single observation of a measurement cannot be more accurate than

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.16

Chapter Twenty-One

the inherent precision of the combined primary and secondary precision. It is possible to have good precision of an inaccurate reading. Thus, precision and accuracy differ. Spurious errors. Spurious errors are commonly caused by accident, resulting in false data. Misreading and intermittent mechanical malfunctions can cause discharge readings well outside of expected random statistical distribution about the mean. Spurious errors can be minimized by good supervision, maintenance, inspection, and training. Experienced, well-trained operators are more likely to recognize readings that are significantly out of the expected range of deviation. Unexpected spiral flow and blockages of flow in the approach or in the device itself can cause spurious errors. Repeating measurements does not provide information on spurious error unless repetitions occur before and after the introduction of the error. On a statistical basis, spurious errors confound evaluation of accuracy performance. Systematic errors. Systematic errors are errors that persist and cannot be considered random. Systematic errors are caused by deviations from standard device dimensions, anomalies to the particular installation, and possible bias in the calibration. Systematic errors cannot be removed or detected by repeated measurements. They usually cause persistent error on one side of the true value. The value of a particular systematic error for a particular device may sometimes be considered as a random error. For example, an installation error in the zero setting for a flume might be + 1 mm for one flume and 2 mm for another. For each flume the error is systematic, but for a number of flumes it would be a random error. Random errors. Random errors are caused by such things as the estimating required between the smallest division on a head measurement device and water surface waves at a head measuring device. Loose linkages between parts of flowmeters provide room for random movement of parts relative to each other, causing subsequent random output errors. Repeated readings decrease the average expected error resulting from random errors by a factor of the square root of the number of readings. Total error. Total error of a measurement is the result of systematic and random errors caused by component parts and factors related to the entire system. Sometimes, error limits of all component factors are well known. In this case, total limits of simpler systems can be determined by computation (Bos et al., 1991). In more complicated cases, it may be difficult to confidently combine the limits. In this case, a thorough calibration of the entire system as a unit can resolve the difference. In any case, it is better to do error analysis with data where entire system parts are operating simultaneously and compare discharge measurement against an adequate discharge comparison standard. Expression of errors. Instrument errors are usually expressed by manufacturers as either a percent of reading or a percent of full scale. The secondary devices based on electronic outputs are more frequently expressed in terms of percent full scale. The designer must be aware that a probable error value of say 1 percent full-scale can exceed 10 percent for small value readings on the output device. When used with weirs, for example, the head reading of h1.5 in the weir equation can increase this 10 percent head measurement error to a 15 percent flow measurement error.

21.4.2 Terms Related to Measurement Capability Linearity. Linearity usually means the maximum deviation in tracking a linearly varying quantity, such as measuring head, and is generally expressed as percent of full scale.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.17

Discrimination. Discrimination is the number of decimals to which the measuring system can be read. Precision is no better than the discrimination. Repeatability. Repeatability is the ability to reproduce the same reading for the same quantities. Thus, it is related to precision. Sensitivity. Sensitivity is the ratio of the change of a secondary measurement, such as head, to the corresponding change of discharge. Range and Rangeability. Range is fully defined by the lowest and highest value that the device can measure without damage and comply within a specified accuracy. The upper and lower range bounds may be the result of mechanical limitations, such as friction at the lower end of the range and possible overdriving damage at the higher end of the range. Range can be designated in other ways: (1) as a simple difference between maximum discharge (Qmax) and minimum discharge (Qmin), (2) as the ratio (Qmax/Qmin), called rangeabilix ty, and (3) as a ratio expressed as 1:(Qmin/Qmax). Neither the difference nor the ratios fully define range without knowledge of either the minimum or maximum discharge. Additional terms (hysteresis, response, lag, rise time). Additional terms related more to dynamic variability might be important when continuous records are needed or if the measurements are being sensed for automatic control of canals and irrigation. Hysteresis is the maximum difference between measurement readings of a quantity established by the same mechanical set point when set from a value above and reset from a value below. Hysteresis can continually get worse as wear of parts increases friction or as linkage freedom increases. Response has several definitions in the instrumentation and measurement field. For water measurement, one definition for response is the smallest change that can be sensed and displayed as a significant measurement. Lag is the time difference of an output reading when tracking a continuously changing quantity. Rise time is often expressed in the form of the time constant, defined as the time for an output of the secondary element to achieve 63 percent of a step change of the input quantity from the primary element.

21.4.3 Comparison Standards Water providers may want, or may be required, to have well-developed measurement programs that are highly managed and standardized. If so, water delivery managers may wish to consult American Society for Testing Materials Standards (ASTM, 1988), Bos (1989), International Organization for Standardization (ISO, 1983: ISO, 1991), and the National Handbook of Recommended Methods for Water Data Acquisition (USGS, 1980). Research laboratories, organizations, and manufacturers that certify measurement devices may need to trace accuracy of measurement through a hierarchy of increasingly rigid standards. The lowest standards in the entire hierarchy of physical comparison standards are called working standards, which are shop or field standards used to control quality of production and measurement. These standards might be gage blocks or rules used to ensure proper dimensions of flumes during manufacturing or devices carried by water providers and users to check the condition of water measurement devices and the quality of their output. Other possible working standards are weights, volume containers, and stopwatches. More complicated devices are used, such as surveyors levels, to check weir staff gage zeros. Dead weight testers and electronic standards are needed to check and maintain more sophisticated and complicated measuring devices, such as acoustic flow meters and devices that use pressure cells to measure head.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.18

Chapter Twenty-One

For further measurement assurance and periodic checking, water users and organizations may keep secondary standards. Secondary standards are used to maintain integrity and performance of working standards. These secondary standards can be sent to government laboratories, one of which is the National Bureau of Standards in Washington, D.C., to be periodically certified after calibration or comparison with accurate replicas of primary standards. Primary standards are defined by international agreement and maintained at the International Bureau of Weights and Measurements in Paris, France. Depending on accuracy needs, each organization should trace their measurement performance up to and through the appropriate level of standards. For example, turbine acceptance testing, such as in the petroleum industry, might justify tracing to the primary standards level.

21.5 SELECTION OF PRIMARY ELEMENTS OF WATER MEASURING DEVICES


21.5.1 General Requirements Design considerations involve the selection of the proper water measurement device for a particular site or situation. Site-specific factors and variables must be considered in extended detail. Each system has unique operational requirements and installation concerns. Knowledge of the immediate measurement needs and reliable estimates on future demands of the proposed system is advantageous. Possible selection constraints may be imposed by laws and compact agreements and should be consulted before selecting a measurement device. Contractual agreements for the purchase of pumps, turbines, and water measuring devices for water supply, sewage and drainage districts often dictate the measurement system required for compliance prior to payment. These constraints may be in terms of accuracy, specific comparison devices, and procedures. Bos (1989) provides an extensive and practical discussion on the selection of open channel water measurement devices. Miller (1996) provides a recent compilation of selection criteria for pipe flow-meters suited to liquids and steam and other gas flows. Bos (1989) provides a selection flow chart and a table of water measurement device properties to guide the selection process for the open channel devices. Miller (1996) describes each meter in detail for the pipe systems, but is more general in leaving the selection to the designer. Because the design engineers for civil engineering projects are most likely to be dealing with irrigation water supply, waste water, or drainage and flood flows, the emphasis is placed on the measuring systems deemed most appropriate to these processes. Large closed-pipe systems for water supplies are frequently encountered, so installation situations appropriate to these will also be included. Gas flows, including steam, are more likely to be encountered by mechanical and chemical engineers and those readers are referred to Miller (1996) and ASME (1959, 1971).

21.5.2 Types of Measuring Devices System operators for water supply, drainage, and waste water commonly use many types of standard water measuring devices, usually in open channels with limited applications in closed conduits. Particularly prominent uses of open channel devices are found in irrigation delivery systems and farm distribution systems, although these measuring devices

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.19

are frequently used for sewer flows and even flood flows. However, the latter two areas of application are frequently more difficult because of the likelihood of heavy bed loads and floating debris. In pipe flowmeters, the most commonly installed devices in industry are the orifice meters, accounting for up to 80 percent of all industrial meters (Miller, 1996). Venturi meters and flow tubes provide much of the remainder. In absolute numbers, the household meters, based on various technologies from nutating disks to paddle-wheel turbines, dominate. For openchannel flows, weirs, flumes, submerged and free orifices, and current meters dominate the flow measuring methods. Pipe flow meters, propeller and turbine, acoustic, magnetic, and vortex-shedding meters are used on large water supply wells such as those used in irrigation and municipal water supply. Differential head meters, such as orifice meters, Venturi meters, and flow tubes, are also used in these applications. The meters considered herein are 1. Open-channel flow devices a. b. c. d. e. f. a. b. c. d. e. f. Current metering (cup, propeller, and electromagnetic probes) Weirs Flumes Acoustic (transonic and Doppler) Tracers Miscellaneous Differential head meters Acoustic (transonic and Doppler) Tracers Turbine/propeller/other insert mechanical Vortex-shedding Miscellaneous

2. Pipe flow devices

The main factors which influence the selection of a measuring device include (USBR, 1997): a. Accuracy requirements b. Cost c. Legal constraints d. Range of flow rates e. Head loss f. Adaptability to site conditions g. Adaptability to variable operating conditions h. Type of measurements and records needed I. Operating requirements j. Ability to pass sediment and debris k. Longevity of device for given environment l. Maintenance requirements m. Construction and installation requirements

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.20

Chapter Twenty-One

n. o. p. q. r.

Device standardization and calibration Field verification, troubleshooting, and repair User acceptance of new methods Vandalism potential Impact on environment

Accuracy requirements. The desired accuracy of the measurement system is an important consideration in the selection of a measurement method. Most water measurement installations, including the primary and secondary devices, can produce accuracies of 5 percent. Some systems are capable of 1 percent under laboratory settings. However, in the field, maintaining such accuracies usually requires considerable expense or special effort in terms of construction, secondary equipment, calibration in-place, and stringent maintenance. Selecting a device that is not appropriate for the site conditions can result in a nonstandard installation of reduced accuracy, sometimes exceeding 10 percent. Accuracies are frequently reported that relate only to the primary measurement method or device. However, many methods require secondary measurement equipment that produces the actual readout. This readout equipment typically increases the overall error of the measurement. Cost. The cost of the measurement method includes the cost of the device itself, the installation, secondary devices, operation, and maintenance. Measurement methods vary widely in their cost and in their serviceable life span. Measurement methods are often selected based on the initial cost of the primary device with insufficient regard for the additional costs associated with providing the desired records of flows over an extended period of time. Legal constraints. Governmental or administrative water board requirements may dictate the water measurement devices or methods. Water measurement devices that become a standard in one geographic area may not necessarily be accepted as a standard elsewhere. In this sense, the term standard does not necessarily signify accuracy or broad legal acceptance. Many water agencies require certain water measurement devices used within their jurisdiction to conform to their standard for the purpose of simplifying operation, employee training, and maintenance. Flow range. Many measurement methods have a limited range of flow conditions for which they are applicable. This range is usually related to the need for certain prescribed flow conditions which are assumed in the development of calibrations. Large errors in measurement can occur when the flow is not within this range. For example, using a bucket and stopwatch for large flows that engulf the bucket is not very accurate. Similarly, sharp-edged devices, such as sharp crested weirs, typically do not yield good results with large channel flows. These are measured better with large flumes or broad-crested weirs, which in turn are not appropriate for trickle flows. Certain applications have typical flow ranges. Irrigation supply monitoring seldom demands a low-flow-to-high-flow range above about 30, while this range on natural stream flows may exceed 1000. In some cases, secondary devices can limit the practical range of flow rates. For example, with devices requiring a head measurement, the accuracy of the head measurement from a visually read wall gage may limit the measurement of low flow rates. For some devices, accuracy is based on percent of the full-scale value. While the resulting error may be well within acceptable limits for full flow, at low flows, the resulting error may become excessive, limiting the usefulness of such measurements. Generally, the device should be

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.21

selected to cover the desired range. Choosing a device that can handle an unnecessary large flow rate may result in compromising measurement capability at low flow rates, and vice versa. This choice depends on the objective of the measurement. For example, in irrigation practice, usually choose a device that can measure the most common flow range at the expense of poorly measuring extremes, such as flood flows. For urban drainage, the flood peak may be important. For practical reasons, different accuracy requirements for high and low flows may be chosen. This is reasonable when an annual total is the primary goal and the low flows contribute a small percentage to that total. Also, if the inaccurate low flow readings are truly a random error then this error approaches zero with large accumulations of readings. Thus the designer needs to know if management decisions are made from individual readings or from long term averages. Headloss. Most water measurement devices require a drop in head. On retrofit installations, for example, to an existing irrigation project, such additional head may not be available, especially in areas that have relatively flat topography. On new projects, incorporating additional headloss into the design can usually be accomplished at reasonable cost. However, a tradeoff usually exists between the cost of the device and the amount of headloss. For example, acoustic flow meters are expensive but require little headloss. Sharp-crested weirs are inexpensive but require a relatively large headloss. The head loss required for a particular measuring device usually varies over the range of discharges. In some cases, head needed by a flow measuring device can reduce the capacity of the channel at that point. Adaptability to site conditions. The selection of a flow measuring device must address the site of the proposed measurement. Several potential sites may be available for obtaining a flow measurement. The particular site chosen may influence the selection of a measuring device. For example, discharge in a canal system can be measured within a reach of the channel or at a structure such as a culvert or check structure. A different device would typically be selected for each site. The device selected ideally should not alter site hydraulics so as to interfere with normal operation and maintenance. Also, the shape of the cross-sectional flow area may favor particular devices. Adaptability to variable operating conditions. Flow demands for most water delivery systems usually vary over a range of flows and flow conditions. The selected device must accommodate the flow range and changes in operating conditions, such as variations in upstream and downstream head. Weirs or flumes should be avoided if downstream water levels can, under some conditions, cause excessive submergence. Also, the information provided by the measuring device should be conveniently useful for the operators performing their duties. Devices that are difficult and time consuming to operate are less likely to be used and are more likely to be used incorrectly. In some cases, water measurement and water level or flow control are desired at the same site. A few devices are available for accomplishing both (e.g., constant-head orifice, vertically movable weirs, and Neyrpic flow module; Bos, 1989). However, separate measurement and control devices are typically linked for this purpose and usually can exceed the performance of combined devices in terms of accuracy and level control, if care is exercised to assure that the separate devices are compatible and achieve both functions when used as a system. Type of measurements and records needed. An accurate measure of instantaneous flow rate is useful for system operators in setting and verifying flow rate. However, because flow rates change over time, a single (instantaneous) reading may not accurately reflect the total volume of water delivered. Where accounting for water volume is desired,

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.22

Chapter Twenty-One

a method of accumulated individual flow measurements is needed. Where flows are steady, daily measurements may be sufficient to infer total volume. Most deliveries, however, require more frequent measurements. Meters that accumulate total delivered volume are desirable where water users take water on demand. Totalizing and automatic recording devices are available for many measuring devices. For large structures, the cost for water-level sensing and recording hardware is small relative to the structure cost. For small structures, these hardware costs remain about the same and thus become a major part of the measurement cost, and may often exceed the cost of the primary structure itself. Many water measuring methods are suitable for making temporary measurements (flow surveys) or performing occasional verification checks of other devices. The method chosen for such a measurement might be quite different from that chosen for continuous monitoring. Although many of these flow survey methods are suited for temporary operation, the focus here is on methods for permanent installations. Operating Requirements. Some measurement methods require manual labor to obtain a measurement. Current metering requires a trained staff with specialized equipment. Pen-and-ink style water-stage recorders need operators to change paper, add ink, and verify proper functioning. Manual recording of flows may require printed forms to be manually completed and data to be accumulated for accounting purposes. Devices with manometers require special care and attention to assure correct differential-head readings. Automated devices, such as ultrasonic flowmeters and other systems that use transducers and electronics, require operator training to set up, adjust, and troubleshoot. Setting gatecontrolled flow rates by simple canal level references or by current metering commonly requires several hours of waiting between gate changes for the downstream canal to fill and stabilize. However, if a flume or weir is installed near the control gate, that portion of the canal can be brought to the stable, desired flow level and measured flow rate in a few minutes, and the canal downstream of the flume or weir can then fill to the correct level over a longer time without further gate adjustments. Thus, the requirements of the operating personnel in using the devices and techniques for their desired purposes must be considered in meter selection. Some measuring devices may inherently serve an additional function applicable to the operation of a water supply system. For example, weirs and flumes serve to hydraulically isolate upstream parts of a canal system from the influence of downstream parts. This occurs for free overfall weirs and flumes flowing below their modular limits. Acoustic, propeller, magnetic, and vortex-shedding flowmeters do not provide this function without additional structural measures such as a downstream overfall. If these meters are used, and the isolation function is desired, then the designer should be made aware of the requirement and provide a free overfall. Isolating the influence of upstream changes from affecting downstream channels, is less easily accomplished. However, it can be partly implemented with orifices that have a differential head that is large compared to the upstream fluctuations. The designer should be aware that a sharp-crested weir overfall requires a relatively high head drop and may need to be excessively wide to provide the isolation function with low absolute head drop. While a long board can be used downstream from a propeller meter to provide the necessary width of flow that will pass a required quantity of water at small head, that small head, and the crude board would not be well suited for measuring flow rate. The designer may wish to take advantage of broad-crested weir behavior and provide a thick crest that can withstand in excess of about 80 percent submergence, which usually translates into low absolute head loss. When used with a propeller meter, for example, the broad-crested weir need not be well defined and can be economically installed (Replogle, 1997).
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.23

Ability to pass sediment and debris. Canal systems often carry a significant amount of sediment in the water. Removal of all suspended solids from the water is usually prohibitively expensive. Thus, some sediment will likely be deposited anywhere the velocities are reduced, which typically occurs near flow measuring structures. Whether this sediment causes a problem depends on the specific structure and the volume of sediment in the water. In some cases, this problem simply requires routine maintenance to remove accumulated sediment; in others, the accumulation can make the flow measurement inaccurate or the device inoperative. Sediment deposits can affect approach conditions and increase approach velocity in front of weirs, flumes, and orifices. Floating and suspended debris such as aquatic plants, washed out bank plants, and fallen tree leaves and twigs can plug some flow measurement devices and cause significant flow measurement problems. Many of the measurement devices which are successfully used in closed conduits (e.g., orifices, propeller meters, and so on) are not usable in culverts or inverted siphons because of debris in the water. Attempting to remove this debris at the entrance to culverts is an additional maintenance problem. Flumes, especially long-throated flumes, can be designed to resist sedimentation. The design options available are to select a structure shape that will maintain velocities that assure erosion of sediments, or at least continued movement of incoming sediments through the flume, at important flow rates. In large broad-crested weirs (a class of longthroated flumes) for capacities greater than 1 m3/s per m of flume width, velocities greater than 1 m/s can be achieved for the upper 75 percent of the flow range, and is usually erosive enough to maintain flume function even for high-sediment bed loads. At the lower flow ranges and for heavy sediment bed loads, deposition is likely and frequent maintenance may be required. Trapezoidal sections tend to retain low velocities into the upper ranges of flow and are less sediment worthy. Long-throated flumes with flat bottoms throughout and side contractions maintain a high velocity for 0.5 m3/s per m width, and higher, but must have throat lengths that are 2 to 3 times the throat width in order to be accurately computable. The sediment worthiness of a flume design depends more on these absolute velocities than on whether the flume floor is flat throughout or raised as in a broadcrested weir. This prompts the designer to select shapes that can provide these velocities. One suggestion for broad crested weirs in a fixed sized channel is to construct a false floor in the head gage area to increase the velocity there and prevent changes in area of flow there. Also, sediments can accumulate in the upstream channel to a depth of the false floor without affecting the function of the flume. This can extend the time between mandatory channel cleaning. Device environment. Any measurement device with moving parts or sensors is subject to failure if it is not compatible with the site environment. Achieving proper operation and longevity of devices is an important selection factor. Very cold weather can shrink moving and fixed parts differentially and solidify oil and grease in bearings. Water can freeze around parts and plug pressure ports and passageways. Acidity and alkalinity in water can corrode metal parts. Water contaminants such as waste solvents can damage lubricants, protective coatings, and plastic parts. Mineral encrustation and biological growths can impair moving parts and plug pressure transmitting ports. Sediment can abrade parts or consolidate tightly in bearing and runner spaces in devices such as propeller meters. Measurement of wastewater and high sediment transport flow may preclude the use of devices that require pressure taps, intrusive sensors, or depend upon clear transmission of

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.24

Chapter Twenty-One

sound through the flow. Water measurement devices that depend on electronic devices and transducers must have appropriate protective housings for harsh environments. Improper protection against the site environment can cause equipment failure or loss of accuracy. Maintenance requirements. The type and amount of maintenance varies widely with different measurement methods. For example, current metering requires periodic maintenance of the current meter itself and maintenance of the meter site to assure that is has a known cross section and velocity distribution. When the flow carries sediment or debris, most weirs, flumes, and orifices require periodic cleaning of the approach channel. As mentioned above, design and meter selection can mitigate the maintenance problems with sediments, but are not likely to eliminate them. Electronic sensors need occasional maintenance to ensure that they are performing properly. Regular maintenance programs are recommended to ensure prolonged measurement quality for all types of devices. Construction and installation requirements. In addition to installation costs, the difficulty of installation and the need to retrofit parts of the existing conveyance system can complicate the selection of water measurement devices. Clearly, devices that can be easily retrofitted into the existing canal system are much preferred because they generally require less down time, and usually present fewer unforeseen problems. Device standardization and calibration. A standard water measurement device infers a documented history of performance based on theory, controlled calibration, and use. A truly standard device has been fully described, accurately calibrated, correctly constructed, properly installed, and sufficiently maintained to fulfill the original installation requirements and flow condition limitations. Discharge equations and tables for standard devices should provide accurate calibration. Maintaining a standard device usually only involves a visual check and measurement of a few specified items or dimensions to ensure that the measuring device has not departed from the standard. Many standard devices have a long history of use and calibration, and thus are potentially more reliable. Commercial availability of a device does not necessarily guarantee that it satisfies the requirements of a standard device. When measuring devices are fabricated onsite or are poorly installed, small deviations from the specified dimensions can occur. These deviations may or may not affect the calibration. The difficulty is that unless an as-built calibration is performed, the degree to which these errors affect the accuracy of the measurements is largely unknown. All too frequently, design deviations are made under the misconception that current metering can be used to provide an accurate field calibration. In practice, calibration by current metering to within 2 percent is difficult to attain. An adequate calibration for free-flow conditions requires many current meter measurements at several discharges. Changing and maintaining a constant discharge for calibration purposes is often difficult under field conditions. Field verification, troubleshooting, and repair. After construction or installation of a device, some verification of the calibration is generally recommended. Usually, the methods used to verify a permanent device (e.g., current metering) are less accurate than the device itself. However, this verification simply serves as a check against gross errors in construction or calibration. For some devices, errors occur as components wear and the calibration slowly drifts away from the original. Other devices have components that simply fail, that is, you get the correct reading or no reading at all. The latter is clearly preferred. However, for many devices, occasional checking is required to ensure that they are still performing as intended. Selection of devices may depend on how they fail and how easy it is to verify that they are performing properly.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.25

User acceptance of new methods. Selection of a water measurement method must also consider the past history of the practice at the site. When improved water measurement methods are needed, proposing changes that build on established practice are generally easier to institute than radical changes. It can be beneficial to select a new method that allows conversion to take place in stages to provide educational examples and demonstrations of the new devices and procedures. Vandalism potential. Instrumentation located near public access is a prime target for vandalism. Where vandalism is a problem, measurement devices with less instrumentation, or instrumentation that can be easily protected, are preferred. When needed, instrumentation can be placed in a buried vault to minimize visibility. Impact on environment. During the selection of a water measurement device, consideration must be given to potential environmental impacts. Water measurement devices vary greatly in the amount of disruption to existing conditions that is needed for installation, operation, and maintenance. For example, installing a weir or flume constricts the channel, slows upstream flow, and accelerates flow within the structure. These changes in the flow conditions can alter local channel erosion, local flooding, public safety, local aquatic habitat, and movement of fish up and down the channel. These factors may alter the cost and selection of a measurement device.

21.5.3 Selection Guidelines Selection of a water measurement method can be a difficult, time-consuming design process if one were to formally evaluate all the factors discussed above for each measuring device. This difficulty is one reason that standardization of measurement devices within water agency jurisdictions is often encouraged by internal administrators. However, useful devices are sometimes overlooked when devices similar to previous purchases are automatically selected. Therefore, some preliminary guidance on selection is offered to the designer so that the number of choices can be narrowed down before a more thorough design analysis of the tradeoffs between alternatives is performed. Short list of devices based on application. The list of practical choices for a water measuring device is quickly narrowed by site conditions because most devices are applicable to a limited range of channel or conduit conditions. Economics also limits applicable devices. For example, few irrigation deliveries to farms can justify expensive acoustic meters. Likewise, using current meters for manual flow measurement in a channel is appropriate for intermittent information but is usually too labor intensive for use on a continuous basis. Table 21.1 provides a list of commonly used measurement methods that are considered appropriate for each of several applications. Table 21.2 provides an abbreviated table of selection criteria and general compliance for categories of water measurement devices. The symbols ( ), (0), and ( ) are used to indicate relative compliance for each selection criterion. The ( ) symbol indicates positive features that might make the device attractive from the standpoint of the associated selection criteria. A ( ) symbol indicates negative aspects that might limit the usefulness of this method based on that criterion. A (0) indicates no strong positive or negative aspects in general. A (V) means that the suitV ability varies widely for this class of devices. The letters NA mean that the device is not applicable for the stated conditions. A single negative value for a device does not mean that the device is not useful or appropriate, but other devices would be preferred for those selection criteria. Vortex-shedding flow meters are not specifically rated in this grouping. They are expected to compete with orifice meter applications. They generally offer less head loss
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.26

Chapter Twenty-One ApplicationBased Selection of Water Measurement Devices

TABLE 21.1

1. Openchannel conveyance system a. Natural channels (see Herschy, 1985) (1) Rivers Periodic current metering of a control section to establish stage discharge relation Broadcrested weirs Longthroated flumes Shortcrested weirs t Acoustic velocity maters (AVMtransient time) Acoustic Doppler velocity profiles Floatvelocity/area method t Slopearea methods (2) Intermediatesized and small streams Current metering/control section Broadcrested weirs Longthroated flumes Shortcrested weirs t Shortthroated flumes t Acoustic velocity meters (AVMtransient time) Floatvelocity/area method t b. Regulated channels (see USBR, 1997) (1.) Spil ways (a) Gated Sluice gates Radial gates (b) Ungated Broadcrested weirs (including special crest shapes, Ogee crest, etc.) Shortcrested weirs t (2) Large canals (a) Control structures Check gates Sluice gates Radial gates Overshot gates (b) Other Longthroated flumes Broadcrested weirs Shortthroated flumes t Acoustic velocity meters

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.27 TABLE 21.1 (Continue)

(3.) Small canals (including openchannel fluid conduit flow) Longthroated flumes Broadcrested weirs Shortthroated flumes t Sharpcrested weirs Rated flow control structures (check gates, radial gates, sluice gates, overshot gates) Acoustic velocity meters (c) Other Floatvelocity/area methods t (4) Irrigation delivery to farm turnout (a) Pipe turnouts (short inverted siphons, submerged culverts, etc.) Metergates Current meter Weirs Shortthroated flumes Longthroated flumes (b) Other Constant head orifice Rated sluice gates Movable weirs 2. Closed conduit conveyance systems (see Brater and King, 1982; Miller, 1996) (a.) Large pipes Venturi meters, venturi tubes, nozzles Rated control gates (orifice) Acoustic velocity meters (transit time) (b.) Small and intermediatesized pipelines Venturi meters, Venturi tubes, nozzles Orifices (inline, endcap, shunt meters, etc.) Propeller and turbine meters Magnetic meters Acoustic meters (transittime and Doppler) Pitot tubes Elbow meters Vortex shedding Trajectory methods (e. g., full-pipe trajectory; California pipe method for part full pipe) Other commercially available meters (household types)
SOURCE: From USBR (1997).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

TABLE 21.2
Submerg Orifice
Head meters* Head meters

Selection Guide for Water Measuring Devices.


Current Metering Vel Meter and Sluice Meters meters Doppler Transonic Acoustic Radial Propeller Differ. Mechan. Magnetic Acoustic Acoustic Acoustic Transon. Pipe (Channel) 0 0 0 0 0 V 0
NA NA NA NA

21.28
(Channel) 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 (Channel) Gates (Pipe Exit) (Pipes) (Pipes) (Pipes) (Pipes) (Pipe, 1 Path) (Multipath) 0 0 0 0
NA NA NA NA NA

Devise

Sharp

Broad

Long

Short

Selection

Crested

Crested

Throated

Throated

Criteria

Weirs

Weirs

Flumes

Flumes

Accuracy

Cost

Flows

5 m3/s

Flows

0.25 m3/s

Flow span

0 0 0 0 0 0

Headloss Site Condition Lined canal 0


NA NA NA NA

NA NA

NA NA

NA NA

NA NA

Unliner canal

Short, full pipe

NA

NA

NA

Closed conduit Measurement Type Rate 0 0 0 0 0 V 0 0 0 0 0 0 0 0 0 0 0 0 0 0

NA

NA

NA

V V 0 0 0 0 0 0 0 0 0

0 0 0 0 0 2 0 0 0 0 0 0 0 0 0 0

Volume Sediment Sediment pass

Debris pass Longivity Moving parts

Electr. requir.

Maintenance

Construction

Field verify

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Standarization

Source: Adapted from USBR (1997). Symbols 1, 0, 2 are used as relative indicators comparing application of the listed water measuring device to the listed criteria Symbol V denotes that situability varies widely. Symbol NA denotes not applicable to criteria

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

Venturi, orifice, pilot tube, etc. Propeller meters, turbine meters, paddle wheel meters, etc.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.29

that orifice meters and can cover a wider discharge range for a particular installation. Although they have been around for many years, they have only recently been offered in a configuration that makes them competitive with orifice meters, which they are generally expected to replace because they can produce less pipe head loss. Open-channel applications for vortex-shedding meters are not considered practical.

21.6 SELECTION OF SECONDARY DEVICES FOR DISCHARGE READOUT AND CONTROL


While the emphasis of this chapter is on hydraulic design, it is important that the output of the design be translated into useful information for the user.

21.6.1 Intended Uses The secondary device that is used with a flowmeter depends strongly on the use of the information. Immediate management decisions, such as adjusting a valve or canal gate, require nearly instant feedback to the operator at an accuracy and precision that fully utilizes the available accuracy of the primary device. As discussed above, random errors over many measurements tend to cancel, and long-term totals can often absorb large random detection errors that would not be tolerable for decisions depending on a single reading. The designer should be cognizant of the user needs and be prepared to provide an appropriate output.

21.6.2 Quality Assurance Secondary devices provide a variety of functions, primary of which are data recording and data quality assurance. The secondary devices necessary for these may not be the same. It is good practice to provide manual, instantaneous flow rate output at the meter site so that the servicing personnel can quickly know that the main secondary instrument is functioning. For weirs and flumes wall gages that show flow rates directly are recommended as a quick visual check of secondary instrumentation. However, a wall gage that shows a head reading that is converted to a discharge rate by the technician using a table or equation will usually suffice. Sometimes it is practical to provide field check capability to a secondary device by special treatment of the installation. For example, if a pressure transducer is used to detect head on a flume, the transducer can be mounted in a stilling well attached to the flume. If it is further mounted on a movable rack, it can allow servicing personnel to raise and lower the transducer a prescribed amount to verify that the output signal correctly reports the change (Replogle, 1997). In pipe flows, differential head meters, such as Venturi and orifice meters should be fitted with manometer ports that can be easily accessed by servicing personnel to verify the detection and transmission of data by electronic devices. Propeller, turbine, acoustic, and magnetic meters are among meters that are usually closely integrated with their secondary devices so that separate verification of the primary device function is more challenging. These checks of the secondary devices do not, however, necessarily detect a malfunction of the primary device, such as scale growth on Venturi surfaces, worn orifice plates,

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.30

Chapter Twenty-One

or sediment filled weirs and flumes. It is sometimes advisable to provide measurement in a main line and in both lines of a bifurcation, thus providing redundancy that can help detect meter malfunctions. Regular inspection is a necessary operational requirement and an important design and selection feature is often the ease of inspection.

21.7 APPLICATIONS OF LONG-THROATED FLUMES


Long-throated flumes are replacing the older open-channel devices, such as Parshall flumes, because of their design flexibility in terms of size, shape, accuracy and economics. Versions for large canals and small irrigation furrows are in use. Shapes vary from rectangular to trapezoidal, with circular versions useful in partly full pipes and complex versions for exceptionally wide flow ranges (Bos et al., 1991; Clemmens et al., 1993; USBR, 1997; Wahl and Clemmens, 1998). Long-throated flumes and broad-crested weirs operate by using a channel contraction to cause critical flow. If there is not enough contraction, critical flow does not occur. Flow is then nonmodular and gauge readings become meaningless. If there is too much contraction, the water surface upstream may be raised excessively and cause canal over-

TABLE 21.3 Choices of Broad Crested Weir Sizes and Rating Tables for Lined Canals, Metric Units.*, Canal Shape Max Side Bottom Slope Width z1 b1 (m) (1) (2) 1.0 0.25 Canal Depth d (m) (4) 0.70 Range of Canal Capacities Lower Upper (m3/s) (m3/s) (4) (5) 0.08 0.09 0.10 0.11 0.12 0.13 0.09 0.10 0.11 0.12 0.13 0.16 0.11 0.12 0.12 0.16 0.18 0.20 0.12 0.13 0.14 0.24|| 0.38 0.43 0.37 0.32 0.21 0.34 0.52 0.52 0.44 0.31 0.33|| 0.52|| 0.68|| 0.64 0.46 0.29 0.39|| 0.62
||

Selection Table 21.4 (6) Am Bm Cm Dm1 Em1 Fm1 Bm Cm Dm1 Em1 Fm1 Gm1 Dm2 Em1Em2 Fm1Fm2 Gm1 Hm Im Em2 Fm2

Weir Weir Shape Crest Sill Width Ht. bc p1 (m) (m) (7) (8) 0.5 0.6 0.7 0.8 0.9 0.1 0.6 0.7 0.8 0.9 1.0 1.2 0.8 0.9 1.0 1.2 1.4 1.6 0.9 1.0 0.125 0.175 0.225 0.275 0.325 0.375 0.15 0.20 0.25 0.30 0.35 0.45 0.15 0.20 0.25 0.35 0.45 0.55 0.15 0.20

Min Head Loss H (m) (9) 0.015 0.018 0.022 0.026 0.030 0.033 0.017 0.021 0.025 0.029 0.033 0.039 0.019 0.024 0.029 0.037 0.043 0.048 0.021 0.025

1.0

0.30

0.75

1.0

0.50

0.80

1.0

0.60

0.90

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.31 TABLE 21.3 (Continued) Canal Depth d (m) (4) 0.16 Range of Canal Capacities Lower Upper (m3/s) (m3/s) (4) (5) 1.09 0.18 0.20 0.22 0.16 0.18 0.20 0.22 0.20 0.24 0.27 0.29 0.32 0.35 0.24 0.27 0.29 0.32 0.35 0.38 0.29 0.32 0.35 0.38 0.43 0.32 0.35 0.38 0.43 0.49 0.55 0.35 0.38 0.43 0.49 0.55 Gm1 0.86 0.64 0.43 0.91 1.51 1.22 0.94 1.3|| 2.1|| 2.5 2.2 1.8 1.4 1.8 2.8 3.9|| 3.5 3.1 2.6 3.4|| 4.7 5.7 5.1 3.9 4.1|| 5.6|| 7.2 5.9 4.5 3.3 4.8|| 6.5|| 8.1|| 6.6 5.1 Selection Table 21.4 (6) 1.2 Hm Im Jm Gm2 Hm Im Jm Km Lm Mm Nm Pm Qm Lm Mm Nm Pm Qm Rm Nm Pm Qm Rm Sm Pm Qm Rm Sm Tm Um Qm Rm Sm Tm Um Weir Weir Shape Crest Sill Width Ht. bc p1 (m) (m) (7) (8) 0.30 1.4 1.6 1.8 1.2 1.4 1.6 1.8 1.50 1.75 2.00 2.25 2.50 2.75 1.75 2.00 2.25 2.50 2.75 3.00 2.25 2.50 2.75 3.00 3.50 2.50 2.75 3.00 3.50 4.00 4.50 2.75 3.00 3.50 4.00 4.50 0.035 0.40 0.50 0.60 0.225 0.325 0.425 0.525 0.300 0.383 0.467 0.550 0.633 0.717 0.333 0.417 0.500 0.583 0.667 0.750 0.417 0.500 0.583 0.667 0.833 0.417 0.500 0.583 0.750 0.917 1.083 0.417 0.500 0.667 0.833 1.000 Min Head Loss H (m) (9) 0.043 0.050 0.049 0.030 0.038 0.047 0.053 0.031 0.38 0.044 0.050 0.056 0.059 0.036 0.042 0.049 0.055 0.062 0.066 0.046 0.052 0.059 0.065 0.081 0.048 0.055 0.061 0.074 0.084 0.089 0.051 0.058 0.071 0.083 0.092

Canal Shape Max Side Bottom Slope Width z1 b1 (m) (1) (2)

1.0

0.75

1.0

1.5

0.60

1.2

1.5

0.75

1.4

1.5

1.00

1.6

1.5

1.25

1.7

1.5

1.50

1.8

Source: Adapted From Replogte et al., (1990). a La Hmax:Lb 2 to 3p1:x La Lb 2 to 3Hmax 3 H L 1.5H1max, but within range given in Table 25.4 d 1.2 h1max p1; H .0.1H1 Limited by sensitivity Maximum recomended canal depth || Limited by Froude number, otherwise limited by canal depth

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.32

Chapter Twenty-One

TABLE 21.4 Rating Equations and Ranges of Application for Trapezoidal Broadcrested Weirs* (Adapted from Replogle, Clemmens and Bos, 1990) Discharge equation: Q C1(h1 C2)U Units: (h1, m; Q, m3/s) Coef for Eq.
0.23

Weir Am
L 0.34 0.30

Weir Bm
L 0.42 0.35

Weir Cm
L 0.58 040

Weir Dm1
L 0.58 0.30

Weir Dm2
L 0.45 0.38

Weir Em1
L 0.56 0.38

Weir Em2
L 0.56 0.42

Weir Fm1
L 0.61

C1 C2 U h1min h1max Q1min Q1max Coef. for Eq.

2.145 0.0067 1.8667 0.030 00.220 0.005 0.140 Weir Fm2


0.42 L

2.365 0.0079 1.8599 0.030 0.280 0.005 0.240 Weir Gm1

2.276 0.0045 1.7597 0.040 0.340 0.010 0.380 Weir Gm2

2.837 0.0125 1.8765 0.040 0.390 0.010 0.520 Weir Hm

2.913 0.0101 1.8555 0.040 0.300 0.010 0.250 Weir Im

2.921 0.0090 1.8152 0.030 0.370 0.010 0.520 Weir Jm

3.037 0.0089 1.8252 0.030 0.370 0.010 0.520 Weir Km

3.122 0.010 1.813 0.030 0.370 0.010 0.680 Weir Lm

0.61 0.50

0.75 0.45

0.68 0.56,

0.84 0.48

0.71

0.40

60

0.48

0.72 0.58

0.87

C1 C2 U h1min h1max Q1min Q1max Coef. for Eq.

3.319 0.0107 1.8376 0.030 0.410 0.010 0.680 Weir Mm


0.65 L

3.601 0.0130 1.8271 0.040 0.500 0.020 1.100 Weir Nm

3.695 0.0108 1.8141 0.040 0.450 0.020 0.920 Weir Pm

3.854 0.0067 1.7522 0.040 0.560 0.020 1.500 Weir Qm

4.135 0.0096 1.7438 0.040 0.480 0.020 1.200 Weir Rm

4.281 0.0048 1.6708 0.030 0.400 0.020 0.940 Weir Sm

4.369 0.0170 1.8324 0.060 0.470 0.040 1.300 Weir Tm

5.389 0.0151 1.8581 0.060 0.580 0.050 2.100 Weir Um

0.97 0.75

1.10 0.80

1.20 0.85

1.28 0.95

1.40 0.95

1.40 0.85

1.20 0.68

1.00

C1 C2 U h1min h1max Q1min Q1max

5.831 0.0169 1.8528 0.060 0.650 0.050 2.800

6.284 0.0268 1.8986 0.090 0.750 0.100 3.900

6.713 0.0190 1.8397 0.080 0.790 0.100 4.700

7.115 0.023 1.8331 0.070 0.850 0.100 5.700

7.563 0.0243 1.8477 0.070 0.930 0.100 7.200

8.397 0.0209 1.8043 0.060 0.950 0.100 8.200

9.144 0.0129 1.7301 0.060 0.810 0.100 6.600

9.789 0.0080 1.6711 0.060 0.660 0.100 5.100

Source: Adapted from Replogle et al. (1993). Calibrations developed with computer model (Clemmens et al., 1993).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.33

topping or other problems. The problem facing the designer is to select the shape of the control section, or throat, so that critical flow occurs throughout the full range of discharges to be measured. Also, the designer must provide acceptable sensitivity and accuracy while not causing too much disruption in upstream flow conditions (e.g., sediment deposition, canal overtopping). This appears to be a difficult task, but existing design aids and rating tables make this task more manageable. A selection of these aids and tables are presented herein. A multitude of possible designs have been sorted, based on practical experience and theory, to a relatively few selected structures from which the designer may choose. Stage-discharge equations or direct rating tables for three types of channel conditions are presented in this chapter. These are: (l) lined trapezoidal channels, (2) earthen canals or lined rectangular channels, and (3) circular pipes and conduits not flowing full. Some offerings for natural streams are presented in Bos et al., (1991) and Replogle et al., (1990). Trapezoidal broad-crested weirs (i.e., having only a bottom contraction) were used for the development of selected standard sizes for lined trapezoidal channels. Calibration equations developed from tables computed with the computer model (Clemmens et al., 1993) are presented in the following series of tables (Tables 21.3 and 21.4). Trapezoidal flumes with side contractions were not selected for general use in standard irrigation situations because they are usually more difficult to construct and require more head loss. A design aid was also developed to help ensure sufficient sensitivity, limit the Froude number, and otherwise assist in selection. The designer needs to only select a weir width with its corresponding sill height. Rectangular broad-crested weirs were again chosen for unlined canals. However, for these weirs the designer must select a channel (and throat) width as well as a sill height, and must be more aware of the other design considerations. For lined rectangular channels, only the sill height must be selected. These tables and equations can also be used to determine the rating for side-contracted, rectangular flumes by appropriate adjustments to handle changing velocity-of-approach problems (Bos et al., 1991). Circular pipes can also be accommodated with broad-crested weirs extending across the pipe. Convenient sizes have been built using sill heights ranging from 0.2 to 0.5 of the pipe diameter. The designer selects a sill height and pipe diameter to handle the desired flow rate and backwater conditions. For natural streams, V shaped flumes were chosen because of the wide range of discharges that must be measured. The only design choice here is the throat side slope. The designer has the option of designing a flume shape or size, not presentad here, by using the theoretically based computer program (Clemmens et al., 1993). A new version of this program for the Windows environment is currently in the beta-test phase (Wahl and Clemmens, 1998).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.34

Chapter Twenty-One

FIGURE 21.6. Broad-crested weir in trapezoidal, concrete-lined canal.

Sediment carrying capabilities are sometimes important design characteristic. As mentioned in a previous section, flumes that are V shaped do not pass bedoad sediment well. The rating tables are developed with the assumption of a particular known approach channel cross section (or flow area). However, any particular control section size and shape can be used with any approach section size and shape. The discharge can be adjusted with an approach velocity coefficient, as discussed in detail in Bos et al., (1991), or custom calibrated with the flume program (Clemmens et al., 1993; Wahl and Clemmens, 1998). The rating tables given here automatically limit the Froude number. If smaller approach areas are used, the designer must determine that the Froude number remains less than about 0.45. Frequently, the site conditions may call for flumes that would have dimensions beyond the ranges provided by the ratings in this chapter. To extend beyond these limits and for further information refer to Ackers et al., (1978), Bos (1989); Bos et al., (1991); Ciemmens et al., (1984, 1993); and Wahl and Clemmens, (1998).

21.7.1 Structures for Lined Trapezoidal Canals Standard broad-crested weir sizes are recommended for use in slip-formed canals of convenient metric dimensions. In lined canals, the canal itself furnishes the entrance channel section La, and the tailwater section (Fig. 21.6). Basic requirements become a converging transition to the throat section and the throat section itself. Calibrations for these standard sizes are included herein so that the subsequent designer may select one of these weirs to be built into an existing lined channel as shown in Fig. 21.6. In selecting these standardsized canals and the related flow rates, consideration was given to proposals by the International Commission on Irrigation and Drainage, to the construction practices of the U.S. Bureau of Reclamation, U.S. Department of the Interior (USDI), and to design criteria for small canals used by the Natural Resources Conservation Service, U.S. Department of Agriculture (USDA) (Bos et al., 1991). Present practice leans or tends toward side slopes of 1:1 horizontal to vertical for small, monolithic, concrete-lined canals with bottom widths less than about 0.8 m and
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.35

depths less than about 1 m. Deeper and wider canals tend toward side slopes of 1.5:1. When the widths and depths are greater than about 3 m, the trend is more toward 2:1 side slopes. This is particularly observed if canal operating procedures may allow rapid dewatering of the canal. In some soil conditions this can cause hydrostatic pressures on the underside of the canal walls that lead to wall failure. Most of the lined canals used in a tertiary irrigation unit or on large farms are of the smaller size. They have 0.3- to 0.6-m bottom widths, 1:1 side slopes, and capacities below 1 m3/s (35 ft3/s). In Table 21.3 precomputed broad-crested weir selections are given for canals with bottom widths at quarter-meter increments, with special insertions for 0.3 m (approximately 1 ft) and 0.6 m (approximately 2 ft). The offering of many precomputed sizes will aid in retrofitting older canal systems and yet not prevent the adoption of standard sized canals as proposed by the International Commission on Irrigation and Drainage (ICID). Canal sizes with bottom widths in excess of 1.5 m or 5 ft, respectively, are avoided in the precomputed tables on the assumption that these sizes deserve special design consideration. Table 21.3 shows a number of precomputed weirs in SI units that may be used for the various combinations of bottom widths and sidewall slopes as given in the first two columns. American units are available in USBR (1997). The third column lists recommended values of maximum canal depth, d, for each side-slope, bottom-width combination. For each canal size, several standard weirs can be used (Column 6). Columns 4 and 5 give the limits on canal capacity for each canal-weir combination. These limits on canal capacity originate from three sources: 1. The Froude number in the approach channel is limited to a maximum of 0.45 to ensure water surface stability. 2. The canal freeboard Fb upstream from the weir should be greater than 20 percent of the upstream sill-referenced head, h1. In terms of canal depth this limit becomes d 1.2 h1 P1. 3. The sensitivity of the weir at maximum flow should be such that a 0.01 m change in the value of the sill-referenced head h1 causes less than 10 percent change in discharge. Also indicated in the last column of Table 21.3 is a minimum headloss H that the weir must provide. Excessive downstream water levels may prevent this minimum headloss, which means that the weir exceeds its modular limit and no longer functions as an accurate measuring device. The required head losses for the various broad-crested weirs were evaluated by the method described earlier and, for design purposes, listed for each weir size with the restriction that the computed modular limit will not exceed 0.90. Thus, the design headloss is either 0.1 H1 or the listed value for H, whichever is greater. For these calculations, it was assumed that the weir was placed in a continuous channel with a constant cross section (e.g., p1 p2, b1 b2, and z1 z2) and that the diverging transition was omitted (abrupt expansion). Technically, the modular limit is based on the drop in total energy head through the weir (i.e., including velocity head). In the above continuous channels, the velocity head component is usually similar in magnitude upstream and downstream from the structure when p1 is approximately equal to p2. This means that h may be substituted for H in most cases. Table 21.3 is primarily intended for the selection of these standard weirs. It is also useful for the selection of canal sizes. The Froude number in the canal is automatically limited to 0.45. Selecting the smallest canal for a given capacity will give a reasonably effi-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.36

Chapter Twenty-One

cient section. For instance, if the design capacity of the canal is to be 1.0 m3/s, the smallest canal that can be incorporated with a measuring structure has b1 0.60 m, z1 1.0, and d 0.90 m. Larger canals can also be used. The hydraulic grade line of the channel should also be checked to ensure an adequate design. Each standard weir can be used for a range of bottom widths. This is possible because the change in flow area upstream from the weir causes only a small change in velocity of approach and a corresponding small change in energy head. The width ranges have been selected so that the error in discharge caused by the change in flow area is less than 1 percent. This is a systematic error for any particular approach channel size, and the extent of this error varies with discharge. A weir suitable for several of the listed canal bottom widths is also suitable for any intermediate width. For example, in Table 21.3, Weir Gm1 can be used in canals with bottom widths of 0.30, 0.50, and 0.60 m, or any intermediate width, for example, b1 0.45 m. The user will need to calculate the sill height, headloss, and maximum design discharge for these intermediate sizes. The rating equations for the weirs are given in Table 21.4 and will reproduce the values presented in the original tables produced by computer modeling (Bos et al., 1991) to within about 1 percent. The original tables were computed using the following criteria (Fig. 21.6): 1. Each weir has a constant bottom width bc and a sill height p1 that varies with the canal dimensions. 2. The ramp length can be chosen such that it is between 2 and 3 times the sill height. The 3:1 ramp slope is preferable. 3. The gage is located a distance at least H1max upstream from the start of the ramp. In addition, it should be located a distance of roughly two to three times H1max from the entrance to the throat. 4. The throat length should be greater than 1.5 times the maximum expected sill-referenced depth h1max, but should be within the limits indicated in Table 21.4. 5. The canal depth must be greater than the sum of ( p1 freeboard requirement, or roughly 0.2 times h1max. h1max Fb), where Fb is the

Occasionally, a weir cannot be found from these tables that will work satisfactorily. The user must then judge between several options, for example: 1. Find a new site for the flume with more vertical head available. 2. Add to the canal wall height upstream from the site so that more backwater effect can be created. 3. Try one of the other weir shapes. 4. Use the tables to interpolate and get a rating for an intermediate width, probably with some sacrifice in accuracy. 5. Produce a special design using the computer model. Example. Given: An existing canal has a bottom width of b1 0.30 m, side slopes z1 1, and a total depth d 0.55 m. The discharge depth relationship was estimated for one flow rate by using surface flows, and using Mannings formula (Eq. 16) to estimate the depth at the maximum expected flow. For this example the maximum flow was estimated to be Qmax 0.15 m3/s at a canal flow depth y2 0.43 m prior to installation of the measuring device. Required: Select a broad-crested weir structure from Table 21.4.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.37

Procedure: The basic strategy for selecting a flume for an existing canal is to cause enough rise in the existing flowing water surface to create the necessary water surface drop needed to allow the flume to function properly. Because the rise in water surface is nearly equal to the required energy loss, H, this value is used to approximate the water surface drop, or, h H. This minimum water surface rise usually must be provided for the maximum expected flow rate, and is then safe for lesser flow rates for this type of installation in a trapezoidal channel. (Note: Placing a rectangular structure in a trapezoidal channel may require checking the low flow also, because as flow rate decreases, the depth drops faster in a rectangular channel than in a trapezoidal channel, and discharge may be nonmodular at low flow rates.) Use Table 21.3 and find the rows for z1 1 and b1 0.30 m. For a flow rate of 0.15 m3/s, weirs Bm through Fm1 can be tried because Qmax falls within the discharge ranges for each of these weirs. Weir Bm is tried first because it has the lowest sill height, p1 0.15 m, and should be the most economical to build. The minimum required head loss listed in column 9, H, for this weir is 0.017 m. The actual headloss through the weir should be the greater of the listed H and 0.1h1. To find the head on the flume, h1, use the discharge equation for long-throated flumes given at the top of Table 21.4: Q C1(h1 C1 C2 U Solving for h1 and with Q h1 0.150 m3/s:
Q 1 u C1

C2)U 2.365 0.0079 1.8599

From Table 21.4 the appropriate coefficients for the discharge equation for flume Bm are

C2 h1

0.15 1.8599 2.365

0.0079

0.219 m

In this case, 0.1 h1 0.0219 m exceeds the listed value of 0.017 m, and the larger value of 0.0219 m is used as the needed drop in water surface. For the flume to function properly, the difference in the upstream water depth, h1 p1, and the downstream channel depth, y2, should be greater than the required water surface drop, or (h1 p1 ) y2 (0.219 0.150) 0.430 0.061 m 0.0219 m

The computed value is actually negative, which means this flume would be completely submerged Now examine the next sized weir. However, because the next size can be estimated to raise the water surface by something slightly less than the sill increase of 0.05 m, it is unlikely that raising the sill by only 0.05 m will work because the previous sill produced no drop, by virtue of the negative value calculated. Thus, move on to weir Dm1 with a p1 0.21 m. In Table 21.4, we find ,with the corresponding values from Table 21.4 for C1,C2 and U, C that h1 0.196 and H = 0.025 m in Table 21.3. However, 0.1 h1 0.020 m, so use the listed value for H 0.025 m as the needed drop. Now again check the difference in the upstream depth, h1 p1, and downstream channel depth, y2 to see if (h1 p1 ) y2 H 0.025 m (the required drop in water surface).
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.38

Chapter Twenty-One

This time, (h1 p1) y2 (0.196 0.250) 0.430 0.016 m, which still does not provide the needed drop of 0.025 m. Next try weir Em1 with p1 0.30 m. This time, h1 is calculated to be 0.186 m and, from Table 21.3, H 0.029 m. Then (h1 p1) y2 (0.186 0.30) 0.430 0.056 m, which exceeds the needed 0.029 m value. Thus, this weir meets the primary criterion and can be checked further.
Weir C1 C2 U h1 0.1 h1 Table H Req. Drop h1+p1 Adequate y2

Bm Dm1 Em1

2.365 2.913 2.921

0.0079 1.8599 0.0101 1.8555 0.0090 1.8152

0.219 0.196 0.186

0.0219 0.217 0.0219 20.061 0.0219 0.025 0.0186 0.029 0.025 0.029 0.016 0.056

No No Yes

The minimum required canal depth, d, including freeboard, can be calculated from dmin 1.2 h1 P1 1.2 (0.186) 0.30 0.523 m, which is less than the 0.550 m canal depth given in this example. This weir is acceptable, and the remainder of the hydraulic dimensions can be obtained from Tables 21.3 and 21.4. They are The length from the gauge to the ramp La = H1max = 0.37 m The ramp length Lb 3 p1 3 (0.3 m) 0.9 m The throat length L 3 p1 1.5 (0.186) 0.28 m However, from Table 21.4, L 0.38 m, so use L 0.38 m, to use the given calibration values. A downstream expansion is not necessary. The flume width bc 0.90 m. This example is modified from Bos et al. (1991). Other examples and detailed discussion of this example are presented therein.

FIGURE 21.7. Flow measuring structure for earthen channel with rectangular control section. (From Bos et al., 1991)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.39 TABLE 21.5 Rating equations and Ranges of Application for Rectangular Weirs*

Lb 2 to 3 times p1; La H1max; La Lb 2 to 3 times H1max H 0.1H1, or value listed H Discharge, q, is given im m3/s per meter width Coef. for Eq.
C1 C2 U h1min h1max qmin qmax H

0.10,bc,0.20m L 5 0.2m p1 5 p1 5 0.05m 8


2.449 0.0003 1.608 0.014 0.130 0.003 0.092 0.012 1.817 0.000 1.530 0.026 0.130 0.003 0.079

0.20,bc,0.30m L 5 0.35m p1 5 p1 5 0.1 8


2.271 0.0014 1.612 0.025 0.235 0.006 0.221 0.025 m 1.744 0.000 1.517 0.025 0.330 0.006 0.192

0.30,bc,0.50m p1 5 0.1
2.276 0.0013 1.615 0.035 0.330 0.011 0.381 0.027 m

p1 5 0.2
2.017 0.0007 1.574 0.035 0.330 0.011 0.353 0.044 m

L 5 0.5m p1 5 8
1.731 0.000 1.517 0.035 0.330 0.011 0.353

Coef. for Eq.


C1 C2 U h1min h1max qmin qmax H

p1 0.1 m
2.316 0.003 1.641 0.050 0.360 0.019 0.438 0.028 m

0.5 bc .0m L 0.75m p1 p1 2m 0.3 m


2.081 0.003 1.611 0.050 0.500 0.018 0.689 0.048 m 1.973 0.003 1.594 0.050 0.500 0.018 0.660 0.063 m

p1
1.709 0.000 1.516 0.050 0.500 0.018 0.595

p1 0.2 m
2.095 0.004 1.627 0.070 0.670 0.030 1.110 0.046 m

1.0 bc 2.0m L 1.0m p1 p1 0.3 m 0.4 m


1.976 0.0027 1.598 0.070 0.670 0.030 1.059 0,066 m 1.976 0.0027 1.599 0.070 0.670 0.030 1.028 0.086 m

p1
1.702 0.000 1.519 0.070 0.670 0.030 0.925

Coef. for Eq.


C1 C2 U h1min h1max qmin qmax H
*

p1 0.1m
2.098 0.0056 1.637 0.100 0.700 0.070 1.191 0.047 m

bc 2.0m L 1.0m p1 0.2m


1.925 0.006 1.609 0.100 1.000 0.051 1.960 0.87 m

p1 0.3m
1.848 0.004 1.581 0.100 1.000 0.051 1.877 0.124 m

p1
1.691 0.000 1.515 0.100 1.000 0.051 1.689

Source: Adapted from Replogle et al. (1990). Calibrations developed with computer model (Clemmens et al., 1993).

21.7.2 Rectangular Structures for Unlined Canals Weirs and flumes for earthen (unlined) channels require a structure that contains the following basic parts: entrance to approach channel, approach channel, converging transition, throat, diverging transition, stilling basin, and riprap protection. As illustrated in Fig. 21.7, the discharge measurement structure for an earthen channel is longer, and thus more expensive, than a structure in a concrete-lined canal (Fig. 21.6).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.40

Chapter Twenty-One

In the latter, the approach channel and sides of the control section already exist and the riprap is not needed. The approach canal of Fig. 21.7 provides a known flow area and velocity of approach. The rating equations for the rectangular weirs given in Table 21.5 assume that the approach section is rectangular and has the same width as the throat. If the upstream sillreferenced head is not measured in a rectangular approach canal of this same width, but instead is measured in the upstream earthen section, then these tables require correction to the discharge, Q, for the change in the approach velocity. Procedures for this are given by Bos (1989). The full-length structure of Fig. 21.7 can be further shortened by deleting the diverging transition or the rectangular tail water channel. The diverging transition may be deleted if the available head loss over the structure exceeds 0.4 H1, so that no velocity head needs to be recovered. The rectangular tail water channel may be deleted if, at maximum flow, the Froude number at the beginning of the exit channel is less than 1.7 (Bos et al.,1991). A rectangular broad-crested weir discharges nearly equal quantities of water over equal widths. The major differences are associated with the friction along the walls. Thus the flow is nearly two-dimensional over the weir, so that rating tables can provide the flow rate, q, in m3/s per meter width of sill for each value of h1. This allows a wide variety of sizes for rectangular broad-crested weirs. For each width, bc, of the weir, an accurate rating table for the total discharge, Q, can be developed by multiplying the table discharges by bc. Thus, Q bcq (21.21)

Table 21.5 presents rating equations for a series of rectangular broad-crested weirs that were developed from the computer-modeled tables given in Bos et al. (1991). These equations will reconstruct those computer-derived tables to within about 1.5 percent. This adds an additional uncertainty to the 3 percent error claimed for the original tables. That 3 percent includes the added error caused by averaging small groupings of weir widths. The weirs were selected to keep the error of side wall effects to within 1 percent. Ratings are given for several sill heights, pl, to aid in design. Interpolation between sill heights gives reasonable results. If the approach area, A1, is larger than that used to develop these rating tables, either because of a higher sill or a wider approach channel, the ratings must be adjusted for Cv. To simplify this process, the discharge over the weir for a Cv value of 1.0 is given in the far right column of each grouping. This column is labeled Pl , because that would cause the velocity of approach to be zero and Cv 1.0. This approximates a weir at the outlet of a reservoir or lake. The complete correction procedure is given in Bos et al., (1991) and Bos (1989). The design procedure for lined rectangular canals, is relatively straightforward. It consists of selecting a sill height, P1, that causes modular flow throughout the discharge range, and provides sufficient freeboard at the maximum discharge. For unlined canals, an appropriate width must be chosen. There are usually several widths that will work. Extremely wide, shallow flows are subject to measurement errors due to low head detection sensitivity. Extremely narrow, deep flows require long structures and large head losses. Because of the wide variety of shapes that can be encountered in earthen channels and in the range of discharges to be measured, it is rather complicated to determine the inter-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.41

related values of h1max, p1, and bc of the structure. While this makes the design process somewhat more complicated, it allows the designer greater flexibility and expands the applicability of the weirs. 1. The discharges to be measured (per meter width) must be within the range of discharges shown in the rating table for the selected weir if these flume dimensions and tables are to be used. 2. The allowable measurement error should not be exceeded. This allowable error may be different at different flow rates (see Bos et al., 1991). 3. Flow should be modular at all flow rates to be measured. 4. Placing a weir in the canal should not cause overtopping upstream. 5. The measuring structure should be placed in a straight section with a relatively uniform cross section for a distance of about 10 times the width of the channel. 6. The Froude number should not exceed 0.45 for a distance of at least 30 times h1 upstream from the weir. The following criteria should be considered by the designer. If these criteria are followed, the designer should obtain a satisfactory structure that will operate as intended. For a rectangular structure in an earthen canal, the rectangular section need not extend 10 times its width upstream from the flume if a gradual taper is used to guide the flow into the rectangular section. For the flumes given here, it is recommended that the rectangular section extend upstream from the head measurement location (gauging station), as shown in Fig. 21.7. It is also recommended that riprap be placed downstream from the structure for a distance of 4 times the maximum downstream water depth. A step should be provided at the transition between the rectangular section and the riprap section to avoid local erosion from floor jets. Sizing of riprap and filters is discussed by Bos (1989) and Bos et al. (1991). An analysis of head measurement errors is presented by Bos (1989) and Bos et al., (1991) and will not be repeated here. For lined channels, a freeboard criterion of 0.2 h1max has been used satisfactorily. For unlined channels it may be more appropriate to specify a

FIGURE 21.8 Long-throated flume in a partly filled pipe.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.42

Chapter Twenty-One

maximum water depth, y1max. Submergence or modular flow should be checked at both minimum and maximum expected discharges. If the channel is rectangular, or the length of the rectangular throated structure downstream from the weir sill is as in Fig. 21.7, then we can use the lower value of H 0.1 H1 or the H value given at the bottom of Table 21.5. If a shorter length in an earthen channel is used and the tailwater channel is significantly larger than the stilling basin would be, then considerably more headloss will probably be required. The designer could use the headloss value for the discharge into a lake or pool, H 0.4 H1. This may rep resent a drastic difference in the value of headloss. The designer may decide to use the shortened structure and calculate the actual modular limit by use of the computer model (Bos et al., 1993; Wahl and Clemmens, 1998). Another alternative is to build a prototype in the field and set the crest to the appropriate level by trial and error. For a more comprehensive background and design discussion for rectangular flumes, refer to Bos et al. (1991).

21.7.3 Structures for Circular Channels The broad-crested weir has been adapted for flow measurements in pipes with a free water surface because of its relatively simple construction, Fig. 21.8. Other shapes, such as side contractions and center line piers can be and have been used. The reader is referred to the computer program and design procedure for most calibrations (Clemmens et al., 1993; Wahl and Clemmens, 1998). There are two situations of particular interest for designing a weir in a pipe. The weir can be placed somewhere in the middle of a straight pipe, or at the end of a pipe, such as the entrance to a deep manhole, which is probably more common because the site is more accessible. When the weir is placed in the interior of a straight pipe, the presence of the weir must cause a rise in the upstream water surface to develop the required head loss. This increase in flow area upstream causes a proportional decrease in velocity and may cause subsequent sediment deposition. Where this is a problem, a downstream ramp should be considered to reduce the required head loss and thus accommodate a smaller sill height. Sediment problems are further aggravated at low flows because the free-flowing (normaldepth) water surface drops proportionally faster than the water surface upstream from the weir. Thus, the velocity difference becomes greater and greater with decreasing discharge. For situations where wide fluctuations in flow rates exist and sedimentation is a problem, an alternative weir shape or location should be considered. An alternative for the measuring location is the end of a pipe, particularly where there is a large drop in water surface. A weir can often be designed so that the water level upstream from the weir either matches, or is below, the normal depth of water in the pipe. This may avoid aggravating sedimentation problems. Portable flumes using plastic pipe. Portable flumes for flow rates up to about 50 L/S with trapezoidal or rectangular shape are described in Bos et al. (1991). More recently, calibrations have been presented for sills placed in circular pipe sections. Small versions are convenient to construct and use as portable flumes. Dimensionless ratings for average roughnesses and profile lengths were previously presented for partly full circular culverts and pipes fitted with similar sills (Clemmens et al., 1984). However, direct computation by the computer model (Clemmens et al., 1993) for the specific roughness of the construction materials provides a slightly more accurate rating. These portable versions can be used for measuring drain tile outlet flows or to measure irrigation furrow flows. An equation for these circular flumes was developed for a group with the following relative proportions:

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.43

La Lb

L 3 p1

3 D/4

where D the diameter of the pipe. The equation is limited to the following ranges of application: 0.2 0.1 m P1 D 0.5 10 m p1)

0.1 L h1max 0.9 (D The developed expression is Q(m3/s) where D2.5C1


h1 D

C2

(21.22)

p1 D p p 2 C2 0.018561 0.070370 1 0.065893 1 D D p U 2.0492 1.8092 1 1.5162 ( 1 /D)2 (p / The forgoing expressionsD were derived by computing a series of flume sills and ramps C1 2.6308 1.7823 (broad-crested weirs) in a pipe 1 m in diameter with sill heights between 0.2 D and 0.5 D. The resulting direct computed tables for each sill height were curve-fitted to generate values for C1, C2, and U for each sill height in the 1-m-diameter pipe. These several values for C1, C2, and u as functions of p1 were further curve-fitted to obtain the above expressions. The D2.5 factor represents the Froude-modeling length ratio applied to generalize the equation to a range that we recommend should not exceed about 10:1 (0.110 times the direct-computed size of D 1 m). For the smallest size, deviations at the low flow ranges can be as high as 3 percent compared to direct computer-modeled tables. For the larger sizes this deviation is less than 1 percent from the direct computed tables. Example. Because of the relative complexity of applying the above equation to various sizes, an example for a common size in Enghish units is presented. The diameter is 1 ft and the sill height is D/3: Convert American measurements to meters: D p1 Then C1 C2 0.018561 2.6308 0.07037 1.7823 0.1016 0.3048 0.065893

1 ft 0.3333 ft

0.3048 m 0.1016 m

2.0367
0.1016 2 0.3048

0.1016 0.3048

0.002426

U and

2.0492

1.8092

0.1016 0.3048

1.5162

0.1016 2 0.3048

1.6146

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.44

Chapter Twenty-One

5/2

0.30485/2

0.05129

Thus, Eq. (21.22), for this pipe size and sill height becomes Q or Q (0.1045)

(0.05129)(2.0367)

h1 0.3048

.002436

1.6146

1.6146

h1 0.3048

.002426

where h1 is expressed in meters and Q is in m3/s. ( Multiply Q by 35.31 to convert to ft3/s.) As with the previously discussed broad-crested weirs, the width, bc, is one of the two most important dimensions in the flume. The other is the head measurement (Bos et al., 1991). For the head measurement, a translocated stilling well is recommended, Fig. 21.8. This method transfers the upstream depth reading to an easily measured location above the sill at the flume outlet. The translocated stilling well conveniently references the upstream head to the sill floor without the necessity of accurately leveling the flume. This makes it truly portable (Bos et al., 1991). The upstream gage reading location should be used only if the flume can be conveniently, or permanently, leveled. If the weir is located in a section of pipe, the normal depth, yn, equals the downstream depth y2. The maximum flow depth for stable, nonpriming flow is about 0.9 of the pipe diameter. Thus, y2 would need to be less than this by an amount equal to the needed head drop in the flume. Thus, y2 H p1 h1 0.9D (21.23)

Equation 21.23 gives the limits on design to provide for modular flow y1 y2 h and to keep the pipe from flowing full y1 0.9D at maximum flow. Here again we use h H. If flow in the pipe is caused by downstream backwater effects (in excess of normal depth), these criteria should be checked at low flows. With a weir at the end of a pipe and sufficient overfall, the weir can be lower to keep approach velocities high for better sediment transport. Preferably, the normal depth, yn, should be larger than or equal to y1. Hence at maximum flow: y1 p1 h1 0.9D (21.24)

All weir and pipe combinations can be considered to be Froude models of each other. This means that rating tables for sizes other than those shown in the table can be readily developed without the computer model, as long as all dimensions remain proportional.

21.8 FIELD SIMPLIFIED, EXPEDIENT MEASUREMENT TECHNIQUES


Obtaining useful design information on an existing flow system in a retrofit situation is frequently a challenge. In open channel flows, reliance on historical flow levels and the uniform flow equations can sometimes provide the needed information (Chow, 1959).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.45

FIGURE 21.9 Suggested method to obtain accurate flow profile information for computing the energy slope of a flowing canal for accurately applying Mannings formula.

Too frequently, true uniform flow does not occur even in canals that appear to have long straight reaches. Careful field surveys of the canal slope and the water surface slope over an extended distance may be required in order to calculate the needed energy slope for use in the required equations. Even then the friction value is usually nebulous. Good current metering records are valuable for both rivers and large canals. Accurate current metering in small canals is often not suitable. However, small canals can frequently be temporarily measured with portable flumes and weirs and the information used for permanent designs.

21.8.1 Channel Roughness Measurement and Water Surface Profile Measurements. An accurate and quick field technique to obtain a channel flow rate using Mannings formula, or determine a channel flow area, or obtain data to calculate energy-line slope for computing the roughness value for Mannings n is illustrated in Fig. 21.9. The technique uses a standard surveying level and two static pressure tubes inserted into the canal water at approximately equal distances upstream and downstream from the surveying instrument site. The static pressure tubes are described in Replogle (1997), and Bos et al. (1991). The distances are typically 50100 m. Using equal distances in both directions is good surveying technique in case the surveying instrument is not in good repair. Stilling wells in the form of small cups placed slightly into the water surface provide a damped water surface for the survey data. The cups need to be tall enough to survive splash from velocity impact and surface waves. Mannings formula (Eq. 21.16) can be used with an estimate of Mannings n value and the channel energy line slope, Se (Chow, 1959). As mentioned earlier, Mannings n value typically is 0.0120.014 for new concrete canals in excellent condition, increasing to 0.018 for damaged concrete surfaces, and in excess of 0.030 when heavy grass-like algae growths are present. The method has been successfully applied to concrete-lined canals to evaluate channel roughness. For this process, an independent measurement of the flow rate was obtained with portable broad-crested weirs as described in another section, and Mannings equation was solved for n. The channel flow areas and the flow velocities at the two points illustrated in Fig. 21.9, could be accurately calculated using the survey data for the water surface. These velocities then provided the velocity head that was added to the respective flow depths to establish the slope of the energy grade line, Se.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.46

Chapter Twenty-One

21.8.2 Portable Flow Measuring Flumes Portable measuring flumes can provide reliable site-specific data that can usually be translated into precise design information for permanent structures. Portable flow measuring flumes are described in Bos et al., (1991) and in Replogle and Wahlin (1998). The latter reference describes flumes with adjustable throats that are especially convenient to install in small unlined canals. They allow accurate measurements with canal ponding of less than 23 cm. The permanent structure design will often be one of the long-throated flumes described in earlier sections. This design process has been expedited by several design-oriented computer programs, which aid in the selection and installation of these devices. These have also been described in earlier sections.

21.8.3

Surface Floats

The velocity of surface floats is sometimes used to estimate flows in lined channels. Unless survey data exist to define a natural channel, the surface velocity times an unknown area of flow produces an unreliable discharge rate. Lined prismatic channels, after probing to determine sedimentation conditions, are a fair candidate for this method of flow estimation. Rather than using a single float, multiple floats consisting usually of buoyant ditchbank trash, popcorn, or even numerous ice cubes, provide a surface float method that is reliable to 10 percent. Ice cubes are particularly useful in windy conditions. The recommended procedure is to apply the floating material across the channel about two or more top widths upstream from a designated point A. Measure the time for the initial front of the floating material to pass from point A to a point B established 3050 m downstream. Because of the many particles, the time for the front edge of the floating material to reach point B should provide the velocity of the fastest surface flow element. This flow element is a function of the channel shape and the channel roughness. Also, the roughness and the length between stations A and B contribute to the longitudinal dispersion of the floating material. While this length dispersion pattern has not been established well enough to firmly define Mannings n, it does appear to be related. For field flow estimates, when the channel is rough or large, the average velocity is about 90 percent of the maximum surface velocity. When the channel is small or smooth, the value approaches 70 percent. While this trend is evident, precise values have not been established. The current information does support an average value of 80 percent. These values differ from the rod floats as used and reported in USBR (1997) where the average is represented by range from 66 percent to 80 percent with an average of 73 percent. However, those floats sampled at least the top quarter of the flow depth as opposed to only the surface velocity suggested here.

21.8.4 Checking a Flow Profile Sometimes there is a need to inexpensively check how the velocity profile is behaving near a measuring device, and to check if measures taken to condition it have been effective. One way to obtain quick and easy results is the rising bubble method. Trickle irrigation tubing is weighted so that it will stay in a straight line across the bottom of the channel of interest. Pressurized air or other gas is released at a rather fast rate from

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.47

FIGURE 21.10 Air bubbles used to check velocity patterns.

the many small holes, Fig. 21.10. The predominant larger bubbles rise uniformly enough to define an undisturbed water surface area between the line of injection and the predominant emergence. The smaller bubbles rise more slowly and emerge in the downstream bubble trail. One immediate observation is the symmetry of the emergence line. A ragged, nonsymmetrical line in a prismatic channel indicates velocity profile distortions. Using rising bubbles as a flow measurement method. This same system can be used to measure discharge rate. The discharge is calculated quite simply by the product of the surface area defined by the emerging bubbles and the release line, multiplied by the average bubble rise velocity of 0.218 m/s (Herschy, 1985). A limitation of this rising-bubble method is the difficulty of measuring the surface area accurately, but this method will give good discharge estimates in poorly defined earth channels and automatically adjusts for both velocity profile and channel shape.

21.8.5 Low-Pressure Pipe Venturi The Venturi tube, although old compared to the modern sonic meters, is still a viable method of measuring fluid flows. It has good anticlogging characteristics and is sometimes used on pressure flows of sewage. In low-pressure pipelines associated with agricultural irrigation (less than 30 m.of pressure head) Venturi tubes were fashioned from standard Polyvinyl chloride (PVC) pipe fittings for 30.5 cm (12-in) main lines and a throat diameter of 20.3 cm (8 in) (Replogle and Wahlin, 1994). Figure 21.11 shows the basic construction. Standard pipe fittings were used to reduce to 10-in pipe then to the 8in pipe. Fig. 21.12 shows a method of making suitable manometer taps.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.48

Chapter Twenty-One

FIGURE 21.11 Configuration installed in the field.

FIGURE 21.12 Pressure tap showing glued-on boss and pipe fitting used in field units.

The basic expression for discharge, Q, is derived from the classical Bernoulli Equation and can be written in a form that is applicable to round pipes or other conduit shapes as Q Cd Ap At
2 p 2 t

2g

(hp

ht)

A A where Cd discharge coefficient, usually 0.90 Cd 0.99, Ap area of approach piping,,At area of contracted throat section, G gravitational constant (9.81 m/s2 = 32.16

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.49

ft/s2), velocity distribution coefficient (estimated as 1.02), and hp ht head between upstream pipe tapping and throat tapping.

Differential

These Venturi meters fashioned from plastic pipe fittings of the kind usually used by the irrigation industry have an expected accuracy of 2 percent, not including the errors of the readout method. It is recommended that a throat length of three times the throat diameter be used for these plastic Venturi meter constructions. Shorter throat lengths appear to cause difficulties in pressure detection due to flow separation. Longer throat lengths produce excessive headloss. The rate of contraction of the reducer fittings caused no significant change in Cd, and thus the meter calibration. However, commercial fittings vary in their contraction angle and the fittings with a 25 contraction angle (measured from pipe centerline) exhibited a greater total head loss through the meter than those with a less severe 15 contraction rate (18 percent and 25 percent of head reading respectively). These constructions conform to expected Venturi meter behavior with a discharge coefficient about 2 percent to 5 percent less than standard Herscheltype Venturi tubes. The discharge coefficient, Cd, can be estimated by the following empirical equation: Cd where m1 0.964; m2 pipe diameter). 0.0466; m3 m1 m2 e(
Rn/m3)

254,000; and Rn

Reynolds number (based on

The most important construction factor is the fabrication of the pressure taps and the immediate connections. They should be drilled with appropriate backing blocks to reduce burrs and with a guide to assure that they are constructed perpendicular to the pipe wall. A slight rounding of the piezometer tap holes will help reduce burrs. It is recommended that the pressure taps be installed on the sides of the meter to prevent air bubbles from entering the pressure lines. It is not necessary that the pressure taps be on the same horizontal line and the meter can be mounted at any angle. Slow-setting PVC cement is recommended to allow workers sufficient time to uniformly assemble and adjust large pipe parts. For an accepted error of 2 percent, one properly constructed pressure tap at each of the two required piping locations is adequate.

25.9 ACKNOWLEDGMENTS
We hereby acknowledge M. G. Bos who encouraged and contributed to a significant cooperative effort with Replogle and Clemmens to produce the first comprehensive treatment of long-throated flumes. That publications by Bos et al., (1991) formed the basis for much of the discussion presented on the newer computable flumes, including the computer program developed to compute them (Clemmens et al., 1993). Much of the remainder of the text for this chapter was condensed from the recent U.S. Bureau of Reclamation publication (USBR, 1997) to which the current authors contributed significantly, but to which we acknowledge the editorial efforts of Mr. Russ Dodge, and other Reclamation employees, who undertook much responsibility to bring that publication to fruition.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

21.50

Chapter Twenty-One

REFERENCES
Ackers, P., W. R. White, J. A. Perkins, and A. J. M. Harrison, Weirs and Flumes for Flow Measurement, John Wiley & Sons, New York, 1978. ASME, Fluid Meters, Their Theory and Application, 5th ed., Report of ASME Research Committee on Fluid Meters, American Society of Mechanical Engineers, New York. 1959. ASME., Fluid Meters, Their Theory and Application, 6th ed., H. S. Bean, ed., Report of ASME Research Committee on Fluid Meters, American Society of Mechanical Engineers, New York. 1971. ASTM, Annual Book of ASTM Standards, Water Environment Technology, Section 11, Vol. 11.01 Water (1), American Society for Testing and Materials, 1988. Bos, M. G. ed.,. Discharge Measurement Structures. 3rd rev. ed., Publication 20,. International Institute for Land Reclamation and Improvement/ILRI, Wageningen, The Netherlands. 1989. Bos, M. G., J. A. Replogle, and A. J. Clemmens, Flow Measuring Flumes for Open Channel Systems, American Society of Agricultural Engineers, St. Joseph, MI, (republication of 1984 edition by John Wiley & Sons), 1991. Brater, E. F., and H. W. King, Handbook of Hydraulics, 6th ed., McGraw-Hill, New York, 1982. Chow, V. Te, Open Channel Hydraulics, McGraw-Hill, New York, 1959. Clemmens, A. J., M. G. Bos, and Replogle, J. A. RBC Broad-Crested Weirs for Circular Sewers and Pipes, Journal of Hydrology (Netherlands), 68:349-368, 1984; Stout, G. E., and G. H. Davis, eds., The Ven Te Chow Memorial Volume. 1984 Clemmens, A. J., M. G., Bos, and Replogle, J. A., FLUME: Design and Calibration of LongThroated Measuring Flumes (with computer disk, Version 3.0), Publication No. 54, International Institute for Land Reclamation and Improvement/ILRI, Wageningen, The Netherlands, 1993. Daugherty, R. L., and A. C. Ingersoll, Fluid Mechanics, 5th ed., McGraw-Hill, New York, 1954. Herschy, R. W., Streamflow Measurement, Elsevier Applied Science Publishers, London and New York, 1985. ISO Standard 5167, Measurement of Fluid Flow by Means of Plates, Nozzles and Venturi Tubes Inserted in Circular Cross-Section Conduits Running Full, ISO 5167-1980(E), Geneva, Switzerland, 1991. ISO, Measurement of Liquid Flows in Open Channels, Handbook No. 15, International Organization for Standardization. Geneva, Switzerland, 1983. Miller, R. W., Flow Measurement Engineering Handbook, 3rd ed., McGraw-Hill, New York, 1996. Reginald, W., Streamflow Measurement, Elsevier Applied Science Publishers, London and New York, 1985. Replogle, J. A., A. J., Clemmens, and Bos, M. G. Measuring Irrigation Water,. in T. A. Howell, K. H. Solomon, and G. Hoffman eds. Management of Farm Irrigation Systems, American Society of Agricultural Engineers, St. Joseph, MI, 1990. Replogle, J. A., and B. T. Wahlin, Venturi Meter Construction for Plastic Irrigation Pipelines, Applied Engineering in Agriculture,10(1):2126, 1994. Replogle, J. and B. Wahlin, Portable and Permanent Flumes with Adjustable Throats, Irrigation and Drainage Systems, Vol. 12, No. 1, Kluwer Academic Publishers, Netherlands, pgs. 23-34., 1998 Replogle, J. A., Practical Technologies for Irrigation Flow Control and Measurement, Journal of Irrigation and Drainage Systems, 11(3):241259, 1997. Spitzer, D. W., Industrial Flow Measurement. Resources for Measurement and Control Series, Instrument Society of America, 1990.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

HYDRAULIC DESIGN OF FLOW MEASURING STRUCTURES

Hydraulic Design of Flow Measuring Structures 21.51 USBR, Water Measurement Manual, U.S. Department of Interior, Bureau of Reclamation, 3rd ed., Superintendent of Documents, U.S. Government Printing Office, Washington, DC, 1997. USGS, National Handbook of Recommended Methods of Water-Data Acquisition, U.S. Geological Survey, Office of Water Data Coordination, Government Printing Office, Washington, DC, 1980. Wahl, T. L., and A. J. Clemmens, Improved Software for Design of Long-Throated Flumes. in Proceedings of the 14th Technical Conference on Irrigation, Drainage and Flood Control, U.S. Committee on Irrigation and Drainage. Phoenix, AZ, 1998.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com) Copyright 2004 The McGraw-Hill Companies. All rights reserved. Any use is subject to the Terms of Use as given at the website.

You might also like