30.11.2014 Views

MECHANICS of FLUIDS LABORATORY - Mechanical Engineering

MECHANICS of FLUIDS LABORATORY - Mechanical Engineering

MECHANICS of FLUIDS LABORATORY - Mechanical Engineering

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

A Manual for the<br />

<strong>MECHANICS</strong> <strong>of</strong> <strong>FLUIDS</strong> <strong>LABORATORY</strong><br />

William S. Janna<br />

Department <strong>of</strong> <strong>Mechanical</strong> <strong>Engineering</strong><br />

Memphis State University


©1997 William S. Janna<br />

All Rights Reserved.<br />

No part <strong>of</strong> this manual may be reproduced, stored in a retrieval<br />

system, or transcribed in any form or by any means—electronic, magnetic,<br />

mechanical, photocopying, recording, or otherwise—<br />

without the prior written consent <strong>of</strong> William S. Janna<br />

2


TABLE OF CONTENTS<br />

Item<br />

Page<br />

Report Writing.................................................................................................................4<br />

Cleanliness and Safety ....................................................................................................6<br />

Experiment 1 Density and Surface Tension.....................................................7<br />

Experiment 2 Viscosity.........................................................................................9<br />

Experiment 3 Center <strong>of</strong> Pressure on a Submerged Plane Surface.............10<br />

Experiment 4 Measurement <strong>of</strong> Differential Pressure..................................12<br />

Experiment 5 Impact <strong>of</strong> a Jet <strong>of</strong> Water ............................................................14<br />

Experiment 6 Critical Reynolds Number in Pipe Flow...............................16<br />

Experiment 7 Fluid Meters................................................................................18<br />

Experiment 8 Pipe Flow .....................................................................................22<br />

Experiment 9 Pressure Distribution About a Circular Cylinder................24<br />

Experiment 10 Drag Force Determination .......................................................27<br />

Experiment 11 Analysis <strong>of</strong> an Airfoil................................................................28<br />

Experiment 12 Open Channel Flow—Sluice Gate .........................................30<br />

Experiment 13 Open Channel Flow Over a Weir ..........................................32<br />

Experiment 14 Open Channel Flow—Hydraulic Jump ................................34<br />

Experiment 15 Open Channel Flow Over a Hump........................................36<br />

Experiment 16 Measurement <strong>of</strong> Velocity and Calibration <strong>of</strong><br />

a Meter for Compressible Flow.............................39<br />

Experiment 17 Measurement <strong>of</strong> Fan Horsepower .........................................44<br />

Experiment 18 Measurement <strong>of</strong> Pump Performance....................................46<br />

Appendix .........................................................................................................................50<br />

3


REPORT WRITING<br />

All reports in the Fluid Mechanics<br />

Laboratory require a formal laboratory report<br />

unless specified otherwise. The report should be<br />

written in such a way that anyone can duplicate<br />

the performed experiment and find the same<br />

results as the originator. The reports should be<br />

simple and clearly written. Reports are due one<br />

week after the experiment was performed, unless<br />

specified otherwise.<br />

The report should communicate several ideas<br />

to the reader. First the report should be neatly<br />

done. The experimenter is in effect trying to<br />

convince the reader that the experiment was<br />

performed in a straightforward manner with<br />

great care and with full attention to detail. A<br />

poorly written report might instead lead the<br />

reader to think that just as little care went into<br />

performing the experiment. Second, the report<br />

should be well organized. The reader should be<br />

able to easily follow each step discussed in the<br />

text. Third, the report should contain accurate<br />

results. This will require checking and rechecking<br />

the calculations until accuracy can be guaranteed.<br />

Fourth, the report should be free <strong>of</strong> spelling and<br />

grammatical errors. The following format, shown<br />

in Figure R.1, is to be used for formal Laboratory<br />

Reports:<br />

Title Page–The title page should show the title<br />

and number <strong>of</strong> the experiment, the date the<br />

experiment was performed, experimenter's<br />

name and experimenter's partners' names.<br />

Table <strong>of</strong> Contents –Each page <strong>of</strong> the report must<br />

be numbered for this section.<br />

Object –The object is a clear concise statement<br />

explaining the purpose <strong>of</strong> the experiment.<br />

This is one <strong>of</strong> the most important parts <strong>of</strong> the<br />

laboratory report because everything<br />

included in the report must somehow relate to<br />

the stated object. The object can be as short as<br />

one sentence and it is usually written in the<br />

past tense.<br />

Theory –The theory section should contain a<br />

complete analytical development <strong>of</strong> all<br />

important equations pertinent to the<br />

experiment, and how these equations are used<br />

in the reduction <strong>of</strong> data. The theory section<br />

should be written textbook-style.<br />

Procedure – The procedure section should contain<br />

a schematic drawing <strong>of</strong> the experimental<br />

setup including all equipment used in a parts<br />

list with manufacturer serial numbers, if any.<br />

Show the function <strong>of</strong> each part when<br />

necessary for clarity. Outline exactly step-<br />

Bibliography<br />

Calibration Curves<br />

Original Data Sheet<br />

(Sample Calculation)<br />

Appendix<br />

Title Page<br />

Discussion & Conclusion<br />

(Interpretation)<br />

Results (Tables<br />

and Graphs)<br />

Procedure (Drawings<br />

and Instructions)<br />

Theory<br />

(Textbook Style)<br />

Object<br />

(Past Tense)<br />

Table <strong>of</strong> Contents<br />

Each page numbered<br />

Experiment Number<br />

Experiment Title<br />

Your Name<br />

Due Date<br />

Partners’ Names<br />

FIGURE R.1. Format for formal reports.<br />

by-step how the experiment was performed in<br />

case someone desires to duplicate it. If it<br />

cannot be duplicated, the experiment shows<br />

nothing.<br />

Results – The results section should contain a<br />

formal analysis <strong>of</strong> the data with tables,<br />

graphs, etc. Any presentation <strong>of</strong> data which<br />

serves the purpose <strong>of</strong> clearly showing the<br />

outcome <strong>of</strong> the experiment is sufficient.<br />

Discussion and Conclusion – This section should<br />

give an interpretation <strong>of</strong> the results<br />

explaining how the object <strong>of</strong> the experiment<br />

was accomplished. If any analytical<br />

expression is to be verified, calculate % error †<br />

and account for the sources. Discuss this<br />

experiment with respect to its faults as well<br />

† % error–An analysis expressing how favorably the<br />

empirical data approximate theoretical information.<br />

There are many ways to find % error, but one method is<br />

introduced here for consistency. Take the difference<br />

between the empirical and theoretical results and divide<br />

by the theoretical result. Multiplying by 100% gives the<br />

% error. You may compose your own error analysis as<br />

long as your method is clearly defined.<br />

4


as its strong points. Suggest extensions <strong>of</strong> the<br />

experiment and improvements. Also<br />

recommend any changes necessary to better<br />

accomplish the object.<br />

Each experiment write-up contains a<br />

number <strong>of</strong> questions. These are to be answered<br />

or discussed in the Discussion and Conclusions<br />

section.<br />

Appendix<br />

(1) Original data sheet.<br />

(2) Show how data were used by a sample<br />

calculation.<br />

(3) Calibration curves <strong>of</strong> instrument which<br />

were used in the performance <strong>of</strong> the<br />

experiment. Include manufacturer <strong>of</strong> the<br />

instrument, model and serial numbers.<br />

Calibration curves will usually be supplied<br />

by the instructor.<br />

(4) Bibliography listing all references used.<br />

Short Form Report Format<br />

Often the experiment requires not a formal<br />

report but an informal report. An informal report<br />

includes the Title Page, Object, Procedure,<br />

Results, and Conclusions. Other portions may be<br />

added at the discretion <strong>of</strong> the instructor or the<br />

writer. Another alternative report form consists<br />

<strong>of</strong> a Title Page, an Introduction (made up <strong>of</strong><br />

shortened versions <strong>of</strong> Object, Theory, and<br />

Procedure) Results, and Conclusion and<br />

Discussion. This form might be used when a<br />

detailed theory section would be too long.<br />

Graphs<br />

In many instances, it is necessary to compose a<br />

plot in order to graphically present the results.<br />

Graphs must be drawn neatly following a specific<br />

format. Figure R.2 shows an acceptable graph<br />

prepared using a computer. There are many<br />

computer programs that have graphing<br />

capabilities. Nevertheless an acceptably drawn<br />

graph has several features <strong>of</strong> note. These features<br />

are summarized next to Figure R.2.<br />

Features <strong>of</strong> note<br />

• Border is drawn about the entire graph.<br />

• Axis labels defined with symbols and<br />

units.<br />

• Grid drawn using major axis divisions.<br />

• Each line is identified using a legend.<br />

• Data points are identified with a<br />

symbol: “ ´” on the Q ac line to denote<br />

data points obtained by experiment.<br />

• The line representing the theoretical<br />

results has no data points represented.<br />

• Nothing is drawn freehand.<br />

• Title is descriptive, rather than<br />

something like Q vs ∆h.<br />

flow rate Q in m 3 /s<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

Q th<br />

Q ac<br />

0 0.2 0.4 0.6 0.8 1<br />

head loss ∆ h in m<br />

FIGURE R.2. Theoretical and actual volume flow rate<br />

through a venturi meter as a function <strong>of</strong> head loss.<br />

5


CLEANLINESS AND SAFETY<br />

Cleanliness<br />

There are “housekeeping” rules that the user<br />

<strong>of</strong> the laboratory should be aware <strong>of</strong> and abide<br />

by. Equipment in the lab is delicate and each<br />

piece is used extensively for 2 or 3 weeks per<br />

semester. During the remaining time, each<br />

apparatus just sits there, literally collecting dust.<br />

University housekeeping staff are not required to<br />

clean and maintain the equipment. Instead, there<br />

are college technicians who will work on the<br />

equipment when it needs repair, and when they<br />

are notified that a piece <strong>of</strong> equipment needs<br />

attention. It is important, however, that the<br />

equipment stay clean, so that dust will not<br />

accumulate too badly.<br />

The Fluid Mechanics Laboratory contains<br />

equipment that uses water or air as the working<br />

fluid. In some cases, performing an experiment<br />

will inevitably allow water to get on the<br />

equipment and/or the floor. If no one cleaned up<br />

their working area after performing an<br />

experiment, the lab would not be a comfortable or<br />

safe place to work in. No student appreciates<br />

walking up to and working with a piece <strong>of</strong><br />

equipment that another student or group <strong>of</strong><br />

students has left in a mess.<br />

Consequently, students are required to clean<br />

up their area at the conclusion <strong>of</strong> the performance<br />

<strong>of</strong> an experiment. Cleanup will include removal<br />

<strong>of</strong> spilled water (or any liquid), and wiping the<br />

table top on which the equipment is mounted (if<br />

appropriate). The lab should always be as clean<br />

or cleaner than it was when you entered. Cleaning<br />

the lab is your responsibility as a user <strong>of</strong> the<br />

equipment. This is an act <strong>of</strong> courtesy that students<br />

who follow you will appreciate, and that you<br />

will appreciate when you work with the<br />

equipment.<br />

Safety<br />

The layout <strong>of</strong> the equipment and storage<br />

cabinets in the Fluid Mechanics Lab involves<br />

resolving a variety <strong>of</strong> conflicting problems. These<br />

include traffic flow, emergency facilities,<br />

environmental safeguards, exit door locations,<br />

etc. The goal is to implement safety requirements<br />

without impeding egress, but still allowing<br />

adequate work space and necessary informal<br />

communication opportunities.<br />

Distance between adjacent pieces <strong>of</strong><br />

equipment is determined by locations <strong>of</strong> floor<br />

drains, and by the need to allow enough space<br />

around the apparatus <strong>of</strong> interest. Immediate<br />

access to the Safety Cabinet is also considered.<br />

Emergency facilities such as showers, eye wash<br />

fountains, spill kits, fire blankets and the like<br />

are not found in the lab. We do not work with<br />

hazardous materials and such safety facilities<br />

are not necessary. However, waste materials are<br />

generated and they should be disposed <strong>of</strong><br />

properly.<br />

Every effort has been made to create a<br />

positive, clean, safety conscious atmosphere.<br />

Students are encouraged to handle equipment<br />

safely and to be aware <strong>of</strong>, and avoid being<br />

victims <strong>of</strong>, hazardous situations.<br />

6


EXPERIMENT 1<br />

FLUID PROPERTIES: DENSITY AND SURFACE TENSION<br />

There are several properties simple<br />

Newtonian fluids have. They are basic<br />

properties which cannot be calculated for every<br />

fluid, and therefore they must be measured.<br />

These properties are important in making<br />

calculations regarding fluid systems. Measuring<br />

fluid properties, density and viscosity, is the<br />

object <strong>of</strong> this experiment.<br />

W 1<br />

W 2<br />

Part I: Density Measurement.<br />

Equipment<br />

Graduated cylinder or beaker<br />

Liquid whose properties are to be<br />

measured<br />

Hydrometer cylinder<br />

Scale<br />

The density <strong>of</strong> the test fluid is to be found by<br />

weighing a known volume <strong>of</strong> the liquid using the<br />

graduated cylinder or beaker and the scale. The<br />

beaker is weighed empty. The beaker is then<br />

filled to a certain volume according to the<br />

graduations on it and weighed again. The<br />

difference in weight divided by the volume gives<br />

the weight per unit volume <strong>of</strong> the liquid. By<br />

appropriate conversion, the liquid density is<br />

calculated. The mass per unit volume, or the<br />

density, is thus measured in a direct way.<br />

A second method <strong>of</strong> finding density involves<br />

measuring buoyant force exerted on a submerged<br />

object. The difference between the weight <strong>of</strong> an<br />

object in air and the weight <strong>of</strong> the object in liquid<br />

is known as the buoyant force (see Figure 1.1).<br />

FIGURE 1.1. Measuring the buoyant force on an<br />

object with a hanging weight.<br />

Referring to Figure 1.1, the buoyant force B is<br />

found as<br />

B = W 1<br />

- W 2<br />

The buoyant force is equal to the difference<br />

between the weight <strong>of</strong> the object in air and the<br />

weight <strong>of</strong> the object while submerged. Dividing<br />

this difference by the volume displaced gives the<br />

weight per unit volume from which density can be<br />

calculated.<br />

Questions<br />

1. Are the results <strong>of</strong> all the density<br />

measurements in agreement?<br />

2. How does the buoyant force vary with<br />

depth <strong>of</strong> the submerged object? Why?<br />

Part II: Surface Tension Measurement<br />

Equipment<br />

Surface tension meter<br />

Beaker<br />

Test fluid<br />

Surface tension is defined as the energy<br />

required to pull molecules <strong>of</strong> liquid from beneath<br />

the surface to the surface to form a new area. It is<br />

therefore an energy per unit area (F⋅L/L 2 = F/L).<br />

A surface tension meter is used to measure this<br />

energy per unit area and give its value directly. A<br />

schematic <strong>of</strong> the surface tension meter is given in<br />

Figure 1.2.<br />

The platinum-iridium ring is attached to a<br />

balance rod (lever arm) which in turn is attached<br />

to a stainless steel torsion wire. One end <strong>of</strong> this<br />

wire is fixed and the other is rotated. As the wire<br />

is placed under torsion, the rod lifts the ring<br />

slowly out <strong>of</strong> the liquid. The proper technique is<br />

to lower the test fluid container as the ring is<br />

lifted so that the ring remains horizontal. The<br />

force required to break the ring free from the<br />

liquid surface is related to the surface tension <strong>of</strong><br />

the liquid. As the ring breaks free, the gage at<br />

the front <strong>of</strong> the meter reads directly in the units<br />

indicated (dynes/cm) for the given ring. This<br />

reading is called the apparent surface tension and<br />

must be corrected for the ring used in order to<br />

obtain the actual surface tension for the liquid.<br />

The correction factor F can be calculated with the<br />

following equation<br />

7


FIGURE 1.2. A schematic <strong>of</strong> the<br />

surface tension meter.<br />

balance rod<br />

platinum<br />

iridium ring<br />

test liquid<br />

clamp<br />

torsion wire<br />

F = 0.725 + √⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺<br />

0.000 403 3(σ a<br />

/ρ) + 0.045 34 - 1.679(r/R)<br />

where F is the correction factor, σ a is the<br />

apparent surface tension read from the dial<br />

(dyne/cm), ρ is the density <strong>of</strong> the liquid (g/cm 3 ),<br />

and (r/R) for the ring is found on the ring<br />

container. The actual surface tension for the<br />

liquid is given by<br />

σ = Fσ a<br />

8


EXPERIMENT 2<br />

FLUID PROPERTIES: VISCOSITY<br />

One <strong>of</strong> the properties <strong>of</strong> homogeneous liquids<br />

is their resistance to motion. A measure <strong>of</strong> this<br />

resistance is known as viscosity. It can be<br />

measured in different, standardized methods or<br />

tests. In this experiment, viscosity will be<br />

measured with a falling sphere viscometer.<br />

The Falling Sphere Viscometer<br />

When an object falls through a fluid medium,<br />

the object reaches a constant final speed or<br />

terminal velocity. If this terminal velocity is<br />

sufficiently low, then the various forces acting on<br />

the object can be described with exact expressions.<br />

The forces acting on a sphere, for example, that is<br />

falling at terminal velocity through a liquid are:<br />

Weight - Buoyancy - Drag = 0<br />

ρ s g 4 3 πR3 - ρg 4 3 πR3 - 6πµVR = 0<br />

where ρ s and ρ are density <strong>of</strong> the sphere and<br />

liquid respectively, V is the sphere’s terminal<br />

velocity, R is the radius <strong>of</strong> the sphere and µ is<br />

the viscosity <strong>of</strong> the liquid. In solving the<br />

preceding equation, the viscosity <strong>of</strong> the liquid can<br />

be determined. The above expression for drag is<br />

valid only if the following equation is valid:<br />

average the results. With the terminal velocity<br />

<strong>of</strong> this and <strong>of</strong> other spheres measured and known,<br />

the absolute and kinematic viscosity <strong>of</strong> the liquid<br />

can be calculated. The temperature <strong>of</strong> the test<br />

liquid should also be recorded. Use at least three<br />

different spheres. (Note that if the density <strong>of</strong><br />

the liquid is unknown, it can be obtained from any<br />

group who has completed or is taking data on<br />

Experiment 1.)<br />

Questions<br />

1. Should the terminal velocity <strong>of</strong> two<br />

different size spheres be the same?<br />

2. Does a larger sphere have a higher<br />

terminal velocity?<br />

3. Should the viscosity found for two different<br />

size spheres be the same? Why or why not?<br />

4. If different size spheres give different<br />

results for the viscosity, what are the error<br />

sources? Calculate the % error and account<br />

for all known error sources.<br />

5. What are the shortcomings <strong>of</strong> this method?<br />

6. Why should temperature be recorded.<br />

7. Can this method be used for gases?<br />

8. Can this method be used for opaque liquids?<br />

9. Can this method be used for something like<br />

peanut butter, or grease or flour dough?<br />

Why or why not?<br />

ρVD<br />

µ < 1<br />

where D is the sphere diameter. Once the<br />

viscosity <strong>of</strong> the liquid is found, the above ratio<br />

should be calculated to be certain that the<br />

mathematical model gives an accurate<br />

description <strong>of</strong> a sphere falling through the<br />

liquid.<br />

Equipment<br />

Hydrometer cylinder<br />

Scale<br />

Stopwatch<br />

Several small spheres with weight and<br />

diameter to be measured<br />

Test liquid<br />

FIGURE 2.1. Terminal velocity measurement (V =<br />

d/time).<br />

V<br />

d<br />

Drop a sphere into the cylinder liquid and<br />

record the time it takes for the sphere to fall a<br />

certain measured distance. The distance divided<br />

by the measured time gives the terminal velocity<br />

<strong>of</strong> the sphere. Repeat the measurement and<br />

9


EXPERIMENT 3<br />

CENTER OF PRESSURE ON A SUBMERGED<br />

PLANE SURFACE<br />

Submerged surfaces are found in many<br />

engineering applications. Dams, weirs and water<br />

gates are familiar examples <strong>of</strong> submerged<br />

surfaces used to control the flow <strong>of</strong> water. From<br />

the design viewpoint, it is important to have a<br />

working knowledge <strong>of</strong> the forces that act on<br />

submerged surfaces.<br />

A plane surface located beneath the surface<br />

<strong>of</strong> a liquid is subjected to a pressure due to the<br />

height <strong>of</strong> liquid above it, as shown in Figure 3.1.<br />

Increasing pressure varies linearly with<br />

increasing depth resulting in a pressure<br />

distribution that acts on the submerged surface.<br />

The analysis <strong>of</strong> this situation involves<br />

determining a force which is equivalent to the<br />

pressure, and finding the location <strong>of</strong> this force.<br />

FIGURE 3.1. Pressure distribution on a submerged<br />

plane surface and the equivalent force.<br />

For this case, it can be shown that the<br />

equivalent force is:<br />

F = ρgy c<br />

A (3.1)<br />

in which ρ is the liquid density, y c is the distance<br />

from the free surface <strong>of</strong> the liquid to the centroid<br />

<strong>of</strong> the plane, and A is the area <strong>of</strong> the plane in<br />

contact with liquid. Further, the location <strong>of</strong> this<br />

force y F below the free surface is<br />

y F = I xx<br />

y c<br />

A + y c<br />

(3.2)<br />

in which I xx is the second area moment <strong>of</strong> the<br />

plane about its centroid. The experimental<br />

F<br />

y F<br />

verification <strong>of</strong> these equations for force and<br />

distance is the subject <strong>of</strong> this experiment.<br />

Center <strong>of</strong> Pressure Measurement<br />

Equipment<br />

Center <strong>of</strong> Pressure Apparatus<br />

Weights<br />

Figure 3.2 gives a schematic <strong>of</strong> the apparatus<br />

used in this experiment. The torus and balance<br />

arm are placed on top <strong>of</strong> the tank. Note that the<br />

pivot point for the balance arm is the point <strong>of</strong><br />

contact between the rod and the top <strong>of</strong> the tank.<br />

The zeroing weight is adjusted to level the<br />

balance arm. Water is then added to a<br />

predetermined depth. Weights are placed on the<br />

weight hanger to re-level the balance arm. The<br />

amount <strong>of</strong> needed weight and depth <strong>of</strong> water are<br />

then recorded. The procedure is then repeated for<br />

four other depths. (Remember to record the<br />

distance from the pivot point to the free surface<br />

for each case.)<br />

From the depth measurement, the equivalent<br />

force and its location are calculated using<br />

Equations 3.1 and 3.2. Summing moments about the<br />

pivot allows for a comparison between the<br />

theoretical and actual force exerted. Referring to<br />

Figure 3.2, we have<br />

WL<br />

F =<br />

(y + y F )<br />

(3.3)<br />

where y is the distance from the pivot point to<br />

the free surface, y F is the distance from the free<br />

surface to the line <strong>of</strong> action <strong>of</strong> the force F, and L is<br />

the distance from the pivot point to the line <strong>of</strong><br />

action <strong>of</strong> the weight W. Note that both curved<br />

surfaces <strong>of</strong> the torus are circular with centers at<br />

the pivot point. For the report, compare the force<br />

obtained with Equation 3.1 to that obtained with<br />

Equation 3.3. When using Equation 3.3, it will be<br />

necessary to use Equation 3.2 for y F .<br />

Questions<br />

1. In summing moments, why isn't the buoyant<br />

force taken into account?<br />

2. Why isn’t the weight <strong>of</strong> the torus and the<br />

balance arm taken into account?<br />

10


L<br />

level<br />

R i<br />

y<br />

zeroing weight<br />

torus<br />

pivot point<br />

(point <strong>of</strong> contact)<br />

weight<br />

hanger<br />

R o<br />

y F<br />

h<br />

F<br />

w<br />

FIGURE 3.2. A schematic <strong>of</strong> the center <strong>of</strong> pressure apparatus.<br />

11


EXPERIMENT 4<br />

MEASUREMENT OF DIFFERENTIAL PRESSURE<br />

Pressure can be measured in several ways.<br />

Bourdon tube gages, manometers, and transducers<br />

are a few <strong>of</strong> the devices available. Each <strong>of</strong> these<br />

instruments actually measures a difference in<br />

pressure; that is, measures a difference between<br />

the desired reading and some reference pressure,<br />

usually atmospheric. The measurement <strong>of</strong><br />

differential pressure with manometers is the<br />

subject <strong>of</strong> this experiment.<br />

Manometry<br />

A manometer is a device used to measure a<br />

pressure difference and display the reading in<br />

terms <strong>of</strong> height <strong>of</strong> a column <strong>of</strong> liquid. The height<br />

is related to the pressure difference by the<br />

hydrostatic equation.<br />

Figure 4.1 shows a U-tube manometer<br />

connected to two pressure vessels. The manometer<br />

reading is ∆h and the manometer fluid has<br />

density ρ m . One pressure vessel contains a fluid <strong>of</strong><br />

density ρ 1 while the other vessel contains a fluid<br />

<strong>of</strong> density ρ 2 . The pressure difference can be found<br />

by applying the hydrostatic equation to each<br />

limb <strong>of</strong> the manometer. For the left leg,<br />

p 1<br />

p 2<br />

1<br />

p A<br />

z 1<br />

FIGURE 4.1. A U-tube manometer connected to<br />

two pressure vessels.<br />

p 1<br />

+ ρ 1<br />

gz 1<br />

= p A<br />

z 2<br />

Likewise for the right leg,<br />

p 2<br />

+ ρ 2<br />

gz 2<br />

+ ρ m<br />

g∆h = p A<br />

Equating these expressions and solving for the<br />

pressure difference gives<br />

h<br />

p A<br />

m<br />

2<br />

p 1<br />

- p 2<br />

= ρ 2<br />

gz 2<br />

+ ρ 1<br />

gz 1<br />

+ ρ m<br />

g∆h<br />

If the fluids above the manometer liquid are both<br />

gases, then ρ 1 and ρ 2 are small compared to ρ µ .<br />

The above equation then becomes<br />

p 1<br />

- p 2<br />

= ρ m<br />

g∆h<br />

Figure 4.2 is a schematic <strong>of</strong> the apparatus<br />

used in this experiment. It consists <strong>of</strong> three U-tube<br />

manometers, a well-type manometer, a U-<br />

tube/inclined manometer and a differential<br />

pressure gage. There are two tanks (actually, two<br />

capped pieces <strong>of</strong> pipe) to which each manometer<br />

and the gage are connected. The tanks have bleed<br />

valves attached and the tanks are connected<br />

with plastic tubing to a squeeze bulb. The bulb<br />

lines also contain valves. With both bleed valves<br />

closed and with both bulb line valves open, the<br />

bulb is squeezed to pump air from the low pressure<br />

tank to the high pressure tank. The bulb is<br />

squeezed until any <strong>of</strong> the manometers reaches its<br />

maximum reading. Now both valves are closed<br />

and the liquid levels are allowed to settle in<br />

each manometer. The ∆h readings are all<br />

recorded. Next, one or both bleed valves are<br />

opened slightly to release some air into or out <strong>of</strong> a<br />

tank. The liquid levels are again allowed to<br />

settle and the ∆h readings are recorded. The<br />

procedure is to be repeated until 5 different sets <strong>of</strong><br />

readings are obtained. For each set <strong>of</strong> readings,<br />

convert all readings into psi or Pa units, calculate<br />

the average value and the standard deviation.<br />

Before beginning, be sure to zero each manometer<br />

and the gage.<br />

Questions<br />

1. Manometers 1, 2 and 3 are U-tube types and<br />

each contains a different liquid. Manometer<br />

4 is a well-type manometer. Is there an<br />

advantage to using this one over a U-tube<br />

type?<br />

2. Manometer 5 is a combined U/tube/inclined<br />

manometer. What is the advantage <strong>of</strong> this<br />

type?<br />

3. Note that some <strong>of</strong> the manometers use a<br />

liquid which has a specific gravity<br />

different from 1.00, yet the reading is in<br />

inches <strong>of</strong> water. Explain how this is<br />

possible.<br />

4. What advantages or disadvantages does<br />

the gage have over the manometers?<br />

12


5. Is a low value <strong>of</strong> the standard deviation<br />

expected? Why?<br />

6. What does a low standard deviation<br />

imply?<br />

7. In your opinion, which device gives the<br />

most accurate reading. What led you to this<br />

conclusion?<br />

High pressure tank<br />

Low pressure tank<br />

Bleed valves<br />

Gage<br />

U-tube manometers<br />

Well-type<br />

manometer<br />

U-tube/inclined<br />

manometer<br />

FIGURE 4.2. A schematic <strong>of</strong> the apparatus used in this experiment.<br />

13


EXPERIMENT 5<br />

IMPACT OF A JET OF WATER<br />

A jet <strong>of</strong> fluid striking a stationary object<br />

exerts a force on that object. This force can be<br />

measured when the object is connected to a spring<br />

balance or scale. The force can then be related to<br />

the velocity <strong>of</strong> the jet <strong>of</strong> fluid and in turn to the<br />

rate <strong>of</strong> flow. The force developed by a jet stream<br />

<strong>of</strong> water is the subject <strong>of</strong> this experiment.<br />

Impact <strong>of</strong> a Jet <strong>of</strong> Liquid<br />

Equipment<br />

Jet Impact Apparatus<br />

Object plates<br />

Figure 5.1 is a schematic <strong>of</strong> the device used in<br />

this experiment. The device consists <strong>of</strong> a tank<br />

within a tank. The interior tank is supported on a<br />

pivot and has a lever arm attached to it. As<br />

water enters this inner tank, the lever arm will<br />

reach a balance point. At this time, a stopwatch<br />

is started and a weight is placed on the weight<br />

hanger (e.g., 10 lbf). When enough water has<br />

entered the tank (10 lbf), the lever arm will<br />

again balance. The stopwatch is stopped. The<br />

elapsed time divided into the weight <strong>of</strong> water<br />

collected gives the weight or mass flow rate <strong>of</strong><br />

water through the system (lbf/sec, for example).<br />

The outer tank acts as a support for the table<br />

top as well as a sump tank. Water is pumped from<br />

the outer tank to the apparatus resting on the<br />

table top. As shown in Figure 5.1, the impact<br />

apparatus contains a nozzle that produces a high<br />

velocity jet <strong>of</strong> water. The jet is aimed at an object<br />

(such as a flat plate or hemisphere). The force<br />

exerted on the plate causes the balance arm to<br />

which the plate is attached to deflect. A weight<br />

is moved on the arm until the arm balances. A<br />

summation <strong>of</strong> moments about the pivot point <strong>of</strong><br />

the arm allows for calculating the force exerted<br />

by the jet.<br />

Water is fed through the nozzle by means <strong>of</strong><br />

a centrifugal pump. The nozzle emits the water in<br />

a jet stream whose diameter is constant. After the<br />

water strikes the object, the water is channeled to<br />

the weighing tank inside to obtain the weight or<br />

mass flow rate.<br />

The variables involved in this experiment<br />

are listed and their measurements are described<br />

below:<br />

1. Mass rate <strong>of</strong> flow–measured with the<br />

weighing tank inside the sump tank. The<br />

volume flow rate is obtained by dividing<br />

mass flow rate by density: Q = m/ρ.<br />

2. Velocity <strong>of</strong> jet–obtained by dividing volume<br />

flow rate by jet area: V = Q/A. The jet is<br />

cylindrical in shape with a diameter <strong>of</strong> 0.375<br />

in.<br />

3. Resultant force—found experimentally by<br />

summation <strong>of</strong> moments about the pivot point<br />

<strong>of</strong> the balance arm. The theoretical resultant<br />

force is found by use <strong>of</strong> an equation derived by<br />

applying the momentum equation to a control<br />

volume about the plate.<br />

Impact Force Analysis<br />

The total force exerted by the jet equals the<br />

rate <strong>of</strong> momentum loss experienced by the jet after<br />

it impacts the object. For a flat plate, the force<br />

equation is:<br />

F = ρQ2<br />

A<br />

For a hemisphere,<br />

F = 2ρQ2<br />

A<br />

(flat plate)<br />

(hemisphere)<br />

For a cone whose included half angle is α,<br />

F = ρQ2<br />

(1 + cos α) (cone)<br />

A<br />

For your report, derive the appropriate<br />

equation for each object you use. Compose a graph<br />

with volume flow rate on the horizontal axis,<br />

and on the vertical axis, plot the actual and<br />

theoretical force. Use care in choosing the<br />

increments for each axis.<br />

14


pivot<br />

balancing<br />

weight<br />

lever arm with<br />

flat plate attached<br />

water<br />

jet<br />

flat plate<br />

nozzle<br />

drain<br />

flow control<br />

valve<br />

weigh tank<br />

tank pivot<br />

plug<br />

weight hanger<br />

sump tank<br />

motor<br />

pump<br />

FIGURE 5.1. A schematic <strong>of</strong> the jet impact apparatus.<br />

15


EXPERIMENT 6<br />

CRITICAL REYNOLDS NUMBER IN PIPE FLOW<br />

The Reynolds number is a dimensionless ratio<br />

<strong>of</strong> inertia forces to viscous forces and is used in<br />

identifying certain characteristics <strong>of</strong> fluid flow.<br />

The Reynolds number is extremely important in<br />

modeling pipe flow. It can be used to determine<br />

the type <strong>of</strong> flow occurring: laminar or turbulent.<br />

Under laminar conditions the velocity<br />

distribution <strong>of</strong> the fluid within the pipe is<br />

essentially parabolic and can be derived from the<br />

equation <strong>of</strong> motion. When turbulent flow exists,<br />

the velocity pr<strong>of</strong>ile is “flatter” than in the<br />

laminar case because the mixing effect which is<br />

characteristic <strong>of</strong> turbulent flow helps to more<br />

evenly distribute the kinetic energy <strong>of</strong> the fluid<br />

over most <strong>of</strong> the cross section.<br />

In most engineering texts, a Reynolds number<br />

<strong>of</strong> 2 100 is usually accepted as the value at<br />

transition; that is, the value <strong>of</strong> the Reynolds<br />

number between laminar and turbulent flow<br />

regimes. This is done for the sake <strong>of</strong> convenience.<br />

In this experiment, however, we will see that<br />

transition exists over a range <strong>of</strong> Reynolds numbers<br />

and not at an individual point.<br />

The Reynolds number that exists anywhere in<br />

the transition region is called the critical<br />

Reynolds number. Finding the critical Reynolds<br />

number for the transition range that exists in pipe<br />

flow is the subject <strong>of</strong> this experiment.<br />

Critical Reynolds Number Measurement<br />

Equipment<br />

Critical Reynolds Number Determination<br />

Apparatus<br />

Figure 6.1 is a schematic <strong>of</strong> the apparatus<br />

used in this experiment. The constant head tank<br />

provides a controllable, constant flow through<br />

the transparent tube. The flow valve in the tube<br />

itself is an on/<strong>of</strong>f valve, not used to control the<br />

flow rate. Instead, the flow rate through the tube<br />

is varied with the rotameter valve at A. The<br />

head tank is filled with water and the overflow<br />

tube maintains a constant head <strong>of</strong> water. The<br />

liquid is then allowed to flow through one <strong>of</strong> the<br />

transparent tubes at a very low flow rate. The<br />

valve at B controls the flow <strong>of</strong> dye; it is opened<br />

and dye is then injected into the pipe with the<br />

water. The dye injector tube is not to be placed in<br />

the pipe entrance as it could affect the results.<br />

Establish laminar flow by starting with a very<br />

low flow rate <strong>of</strong> water and <strong>of</strong> dye. The injected<br />

dye will flow downstream in a threadlike<br />

pattern for very low flow rates. Once steady state<br />

is achieved, the rotameter valve is opened<br />

slightly to increase the water flow rate. The<br />

valve at B is opened further if necessary to allow<br />

more dye to enter the tube. This procedure <strong>of</strong><br />

increasing flow rate <strong>of</strong> water and <strong>of</strong> dye (if<br />

necessary) is repeated throughout the<br />

experiment.<br />

Establish laminar flow in one <strong>of</strong> the tubes.<br />

Then slowly increase the flow rate and observe<br />

what happens to the dye. Its pattern may<br />

change, yet the flow might still appear to be<br />

laminar. This is the beginning <strong>of</strong> transition.<br />

Continue increasing the flow rate and again<br />

observe the behavior <strong>of</strong> the dye. Eventually, the<br />

dye will mix with the water in a way that will<br />

be recognized as turbulent flow. This point is the<br />

end <strong>of</strong> transition. Transition thus will exist over a<br />

range <strong>of</strong> flow rates. Record the flow rates at key<br />

points in the experiment. Also record the<br />

temperature <strong>of</strong> the water.<br />

The object <strong>of</strong> this procedure is to determine<br />

the range <strong>of</strong> Reynolds numbers over which<br />

transition occurs. Given the tube size, the<br />

Reynolds number can be calculated with:<br />

Re = VD<br />

ν<br />

where V (= Q/A) is the average velocity <strong>of</strong><br />

liquid in the pipe, D is the hydraulic diameter <strong>of</strong><br />

the pipe, and ν is the kinematic viscosity <strong>of</strong> the<br />

liquid.<br />

The hydraulic diameter is calculated from<br />

its definition:<br />

D =<br />

4 x Area<br />

Wetted Perimeter<br />

For a circular pipe flowing full, the hydraulic<br />

diameter equals the inside diameter <strong>of</strong> the pipe.<br />

For a square section, the hydraulic diameter will<br />

equal the length <strong>of</strong> one side (show that this is<br />

the case). The experiment is to be performed for<br />

both round tubes and the square tube. With good<br />

technique and great care, it is possible for the<br />

transition Reynolds number to encompass the<br />

traditionally accepted value <strong>of</strong> 2 100.<br />

16


Questions<br />

1. Can a similar procedure be followed for<br />

gases?<br />

2. Is the Reynolds number obtained at<br />

transition dependent on tube size or shape?<br />

3. Can this method work for opaque liquids?<br />

dye reservoir<br />

drilled partitions<br />

B<br />

transparent tube<br />

on/<strong>of</strong>f valve<br />

rotameter<br />

inlet to<br />

tank<br />

overflow<br />

to drain<br />

FIGURE 6.1. The critical Reynolds number determination apparatus.<br />

A<br />

to drain<br />

17


EXPERIMENT 7<br />

FLUID METERS IN INCOMPRESSIBLE FLOW<br />

There are many different meters used in pipe<br />

flow: the turbine type meter, the rotameter, the<br />

orifice meter, the venturi meter, the elbow meter<br />

and the nozzle meter are only a few. Each meter<br />

works by its ability to alter a certain physical<br />

characteristic <strong>of</strong> the flowing fluid and then<br />

allows this alteration to be measured. The<br />

measured alteration is then related to the flow<br />

rate. A procedure <strong>of</strong> analyzing meters to<br />

determine their useful features is the subject <strong>of</strong><br />

this experiment.<br />

The Venturi Meter<br />

The venturi meter is constructed as shown in<br />

Figure 7.1. It contains a constriction known as the<br />

throat. When fluid flows through the<br />

constriction, it must experience an increase in<br />

velocity over the upstream value. The velocity<br />

increase is accompanied by a decrease in static<br />

pressure at the throat. The difference between<br />

upstream and throat static pressures is then<br />

measured and related to the flow rate. The<br />

greater the flow rate, the greater the pressure<br />

drop ∆p. So the pressure difference ∆h (= ∆p/ρg)<br />

can be found as a function <strong>of</strong> the flow rate.<br />

1<br />

h<br />

2<br />

FIGURE 7.1. A schematic <strong>of</strong> the Venturi meter.<br />

Using the hydrostatic equation applied to<br />

the air-over-liquid manometer <strong>of</strong> Figure 7.1, the<br />

pressure drop and the head loss are related by<br />

(after simplification):<br />

p 1<br />

- p 2<br />

ρg<br />

= ∆h<br />

By combining the continuity equation,<br />

Q = A 1<br />

V 1<br />

= A 2<br />

V 2<br />

with the Bernoulli equation,<br />

p 1<br />

ρ + V 1 2<br />

2 = p 2<br />

ρ + V 2 2<br />

2<br />

and substituting from the hydrostatic equation, it<br />

can be shown after simplification that the<br />

volume flow rate through the venturi meter is<br />

given by<br />

Q = A th 2<br />

√⎺⎺⎺⎺<br />

2g∆h<br />

1 - (D 24<br />

/D 14<br />

)<br />

(7.1)<br />

The preceding equation represents the theoretical<br />

volume flow rate through the venturi meter.<br />

Notice that is was derived from the Bernoulli<br />

equation which does not take frictional effects<br />

into account.<br />

In the venturi meter, there exists small<br />

pressure losses due to viscous (or frictional)<br />

effects. Thus for any pressure difference, the<br />

actual flow rate will be somewhat less than the<br />

theoretical value obtained with Equation 7.1<br />

above. For any ∆h, it is possible to define a<br />

coefficient <strong>of</strong> discharge C v as<br />

C v<br />

= Q ac<br />

Q th<br />

For each and every measured actual flow rate<br />

through the venturi meter, it is possible to<br />

calculate a theoretical volume flow rate, a<br />

Reynolds number, and a discharge coefficient.<br />

The Reynolds number is given by<br />

Re = V 2 D 2<br />

(7.2)<br />

ν<br />

where V 2<br />

is the velocity at the throat <strong>of</strong> the<br />

meter (= Q ac<br />

/A 2<br />

).<br />

The Orifice Meter and<br />

Nozzle-Type Meter<br />

The orifice and nozzle-type meters consist <strong>of</strong><br />

a throttling device (an orifice plate or bushing,<br />

respectively) placed into the flow. (See Figures<br />

7.2 and 7.3). The throttling device creates a<br />

measurable pressure difference from its upstream<br />

to its downstream side. The measured pressure<br />

difference is then related to the flow rate. Like<br />

the venturi meter, the pressure difference varies<br />

with flow rate. Applying Bernoulli’s equation to<br />

points 1 and 2 <strong>of</strong> either meter (Figure 7.2 or Figure<br />

7.3) yields the same theoretical equation as that<br />

for the venturi meter, namely, Equation 7.1. For<br />

any pressure difference, there will be two<br />

associated flow rates for these meters: the<br />

theoretical flow rate (Equation 7.1), and the<br />

18


actual flow rate (measured in the laboratory).<br />

The ratio <strong>of</strong> actual to theoretical flow rate leads<br />

to the definition <strong>of</strong> a discharge coefficient: C o<br />

for<br />

the orifice meter and C n<br />

for the nozzle.<br />

rotor supported<br />

on bearings<br />

(not shown)<br />

to receiver<br />

h<br />

1 2<br />

flow<br />

straighteners<br />

turbine rotor<br />

rotational speed<br />

proportional to<br />

flow rate<br />

FIGURE 7.4. A schematic <strong>of</strong> a turbine-type flow<br />

meter.<br />

FIGURE 7.2. Cross sectional view <strong>of</strong> the orifice<br />

meter.<br />

h<br />

1 2<br />

FIGURE 7.3. Cross sectional view <strong>of</strong> the nozzletype<br />

meter, and a typical nozzle.<br />

For each and every measured actual flow<br />

rate through the orifice or nozzle-type meters, it<br />

is possible to calculate a theoretical volume flow<br />

rate, a Reynolds number and a discharge<br />

coefficient. The Reynolds number is given by<br />

Equation 7.2.<br />

The Turbine-Type Meter<br />

The turbine-type flow meter consists <strong>of</strong> a<br />

section <strong>of</strong> pipe into which a small “turbine” has<br />

been placed. As the fluid travels through the<br />

pipe, the turbine spins at an angular velocity<br />

that is proportional to the flow rate. After a<br />

certain number <strong>of</strong> revolutions, a magnetic pickup<br />

sends an electrical pulse to a preamplifier which<br />

in turn sends the pulse to a digital totalizer. The<br />

totalizer totals the pulses and translates them<br />

into a digital readout which gives the total<br />

volume <strong>of</strong> liquid that travels through the pipe<br />

and/or the instantaneous volume flow rate.<br />

Figure 7.4 is a schematic <strong>of</strong> the turbine type flow<br />

meter.<br />

The Rotameter (Variable Area Meter)<br />

The variable area meter consists <strong>of</strong> a tapered<br />

metering tube and a float which is free to move<br />

inside. The tube is mounted vertically with the<br />

inlet at the bottom. Fluid entering the bottom<br />

raises the float until the forces <strong>of</strong> buoyancy, drag<br />

and gravity are balanced. As the float rises the<br />

annular flow area around the float increases.<br />

Flow rate is indicated by the float position read<br />

against the graduated scale which is etched on<br />

the metering tube. The reading is made usually at<br />

the widest part <strong>of</strong> the float. Figure 7.5 is a sketch<br />

<strong>of</strong> a rotameter.<br />

freely<br />

suspended<br />

float<br />

outlet<br />

tapered, graduated<br />

transparent tube<br />

inlet<br />

FIGURE 7.5. A schematic <strong>of</strong> the rotameter and its<br />

operation.<br />

Rotameters are usually manufactured with<br />

one <strong>of</strong> three types <strong>of</strong> graduated scales:<br />

1. % <strong>of</strong> maximum flow–a factor to convert scale<br />

reading to flow rate is given or determined for<br />

the meter. A variety <strong>of</strong> fluids can be used<br />

with the meter and the only variable<br />

19


encountered in using it is the scale factor. The<br />

scale factor will vary from fluid to fluid.<br />

2. Diameter-ratio type–the ratio <strong>of</strong> cross<br />

sectional diameter <strong>of</strong> the tube to the<br />

diameter <strong>of</strong> the float is etched at various<br />

locations on the tube itself. Such a scale<br />

requires a calibration curve to use the meter.<br />

3. Direct reading–the scale reading shows the<br />

actual flow rate for a specific fluid in the<br />

units indicated on the meter itself. If this<br />

type <strong>of</strong> meter is used for another kind <strong>of</strong> fluid,<br />

then a scale factor must be applied to the<br />

readings.<br />

Experimental Procedure<br />

Equipment<br />

Fluid Meters Apparatus<br />

Stopwatch<br />

The fluid meters apparatus is shown<br />

schematically in Figure 7.6. It consists <strong>of</strong> a<br />

centrifugal pump, which draws water from a<br />

sump tank, and delivers the water to the circuit<br />

containing the flow meters. For nine valve<br />

positions (the valve downstream <strong>of</strong> the pump),<br />

record the pressure differences in each<br />

manometer. For each valve position, measure the<br />

actual flow rate by diverting the flow to the<br />

volumetric measuring tank and recording the time<br />

required to fill the tank to a predetermined<br />

volume. Use the readings on the side <strong>of</strong> the tank<br />

itself. For the rotameter, record the position <strong>of</strong><br />

the float and/or the reading <strong>of</strong> flow rate given<br />

directly on the meter. For the turbine meter,<br />

record the flow reading on the output device.<br />

Note that the venturi meter has two<br />

manometers attached to it. The “inner”<br />

manometer is used to calibrate the meter; that is,<br />

to obtain ∆h readings used in Equation 7.1. The<br />

“outer” manometer is placed such that it reads<br />

the overall pressure drop in the line due to the<br />

presence <strong>of</strong> the meter and its attachment fittings.<br />

We refer to this pressure loss as ∆H (distinctly<br />

different from ∆h). This loss is also a function <strong>of</strong><br />

flow rate. The manometers on the turbine-type<br />

and variable area meters also give the incurred<br />

loss for each respective meter. Thus readings <strong>of</strong><br />

∆H vs Q ac are obtainable. In order to use these<br />

parameters to give dimensionless ratios, pressure<br />

coefficient and Reynolds number are used. The<br />

Reynolds number is given in Equation 7.2. The<br />

pressure coefficient is defined as<br />

C p<br />

= g∆H<br />

V 2 /2<br />

(7.3)<br />

All velocities are based on actual flow rate and<br />

pipe diameter.<br />

The amount <strong>of</strong> work associated with the<br />

laboratory report is great; therefore an informal<br />

group report is required rather than individual<br />

reports. The write-up should consist <strong>of</strong> an<br />

Introduction (to include a procedure and a<br />

derivation <strong>of</strong> Equation 7.1), a Discussion and<br />

Conclusions section, and the following graphs:<br />

1. On the same set <strong>of</strong> axes, plot Q ac vs ∆h and<br />

Q th vs ∆h with flow rate on the vertical<br />

axis for the venturi meter.<br />

2. On the same set <strong>of</strong> axes, plot Q ac vs ∆h and<br />

Q th vs ∆h with flow rate on the vertical<br />

axis for the orifice meter.<br />

3. Plot Q ac<br />

vs Q th<br />

for the turbine type meter.<br />

4. Plot Q ac<br />

vs Q th<br />

for the rotameter.<br />

5. Plot C v vs Re on a log-log grid for the<br />

venturi meter.<br />

6. Plot C o vs Re on a log-log grid for the orifice<br />

meter.<br />

7. Plot ∆H vs Q ac for all meters on the same set<br />

<strong>of</strong> axes with flow rate on the vertical axis.<br />

8. Plot C p vs Re for all meters on the same set<br />

<strong>of</strong> axes (log-log grid) with C p vertical axis.<br />

Questions<br />

1. Referring to Figure 7.2, recall that<br />

Bernoulli's equation was applied to points 1<br />

and 2 where the pressure difference<br />

measurement is made. The theoretical<br />

equation, however, refers to the throat area<br />

for point 2 (the orifice hole diameter)<br />

which is not where the pressure<br />

measurement was made. Explain this<br />

discrepancy and how it is accounted for in<br />

the equation formulation.<br />

2. Which meter in your opinion is the best one<br />

to use?<br />

3. Which meter incurs the smallest pressure<br />

loss? Is this necessarily the one that should<br />

always be used?<br />

4. Which is the most accurate meter?<br />

5. What is the difference between precision<br />

and accuracy?<br />

20


manometer<br />

orifice meter<br />

venturi meter<br />

volumetric<br />

measuring<br />

tank<br />

rotameter<br />

return<br />

sump tank<br />

turbine-type meter<br />

motor pump valve<br />

FIGURE 7.6. A schematic <strong>of</strong> the Fluid Meters Apparatus. (Orifice and Venturi meters: upstream<br />

diameter is 1.025 inches; throat diameter is 0.625 inches.)<br />

21


EXPERIMENT 8<br />

PIPE FLOW<br />

Experiments in pipe flow where the presence<br />

<strong>of</strong> frictional forces must be taken into account are<br />

useful aids in studying the behavior <strong>of</strong> traveling<br />

fluids. Fluids are usually transported through<br />

pipes from location to location by pumps. The<br />

frictional losses within the pipes cause pressure<br />

drops. These pressure drops must be known to<br />

determine pump requirements. Thus a study <strong>of</strong><br />

pressure losses due to friction has a useful<br />

application. The study <strong>of</strong> pressure losses in pipe<br />

flow is the subject <strong>of</strong> this experiment.<br />

Pipe Flow<br />

Equipment<br />

Pipe Flow Test Rig<br />

Figure 8.1 is a schematic <strong>of</strong> the pipe flow test<br />

rig. The rig contains a sump tank which is used as<br />

a water reservoir from which a centrifugal pump<br />

discharges water to the pipe circuit. The circuit<br />

itself consists <strong>of</strong> four different diameter lines and<br />

a return line all made <strong>of</strong> drawn copper tubing. The<br />

circuit contains valves for directing and<br />

regulating the flow to make up various series and<br />

parallel piping combinations. The circuit has<br />

provision for measuring pressure loss through the<br />

use <strong>of</strong> static pressure taps (manometer board not<br />

shown in schematic). Finally, because the circuit<br />

also contains a rotameter, the measured pressure<br />

losses can be obtained as a function <strong>of</strong> flow rate.<br />

As functions <strong>of</strong> the flow rate, measure the<br />

pressure losses in inches <strong>of</strong> water for (as specified<br />

by the instructor):<br />

1. 1 in. copper tube 5. 1 in. 90 T-joint<br />

2. 3 /4-in. copper tube 6. 1 in. 90 elbow (ell)<br />

3. 1 /2-in copper tube 7. 1 in. gate valve<br />

4. 3 /8 in copper tube 8. 3 /4-in gate valve<br />

• The instructor will specify which <strong>of</strong> the<br />

pressure loss measurements are to be taken.<br />

• Open and close the appropriate valves on the<br />

apparatus to obtain the desired flow path.<br />

• Use the valve closest to the pump on its<br />

downstream side to vary the volume flow<br />

rate.<br />

• With the pump on, record the assigned<br />

pressure drops and the actual volume flow<br />

rate from the rotameter.<br />

• Using the valve closest to the pump, change<br />

the volume flow rate and again record the<br />

pressure drops and the new flow rate value.<br />

• Repeat this procedure until 9 different<br />

volume flow rates and corresponding pressure<br />

drop data have been recorded.<br />

With pressure loss data in terms <strong>of</strong> ∆h, the<br />

friction factor can be calculated with<br />

2g∆h<br />

f =<br />

V 2 (L/D)<br />

It is customary to graph the friction factor as a<br />

function <strong>of</strong> the Reynolds number:<br />

Re = VD<br />

ν<br />

The f vs Re graph, called a Moody Diagram is<br />

traditionally drawn on a log-log grid. The graph<br />

also contains a third variable known as the<br />

roughness coefficient ε/D. For this experiment<br />

the roughness factor ε is that for drawn tubing.<br />

Where fittings are concerned, the loss<br />

incurred by the fluid is expressed in terms <strong>of</strong> a loss<br />

coefficient K. The loss coefficient for any fitting<br />

can be calculated with<br />

K =<br />

∆ h<br />

V 2 /2g<br />

where ∆h is the pressure (or head) loss across the<br />

fitting. Values <strong>of</strong> K as a function <strong>of</strong> Q ac are to be<br />

obtained in this experiment.<br />

For the report, calculate friction factor f and<br />

graph it as a function <strong>of</strong> Reynolds number Re for<br />

items 1 through 4 above as appropriate. Compare<br />

to a Moody diagram. Also calculate the loss<br />

coefficient for items 5 through 8 above as<br />

appropriate, and determine if the loss coefficient<br />

K varies with flow rate or Reynolds number.<br />

Compare your K values to published ones.<br />

Note that gate valves can have a number <strong>of</strong><br />

open positions. For purposes <strong>of</strong> comparison it is<br />

<strong>of</strong>ten convenient to use full, half or one-quarter<br />

open.<br />

22


otameter<br />

tank<br />

valve<br />

motor<br />

static pressure tap<br />

pump<br />

FIGURE 8.1. Schematic <strong>of</strong> the pipe friction apparatus.<br />

23


EXPERIMENT 9<br />

PRESSURE DISTRIBUTION ABOUT A CIRCULAR CYLINDER<br />

In many engineering applications, it may be<br />

necessary to examine the phenomena occurring<br />

when an object is inserted into a flow <strong>of</strong> fluid. The<br />

wings <strong>of</strong> an airplane in flight, for example, may<br />

be analyzed by considering the wings stationary<br />

with air moving past them. Certain forces are<br />

exerted on the wing by the flowing fluid that<br />

tend to lift the wing (called the lift force) and to<br />

push the wing in the direction <strong>of</strong> the flow (drag<br />

force). Objects other than wings that are<br />

symmetrical with respect to the fluid approach<br />

direction, such as a circular cylinder, will<br />

experience no lift, only drag.<br />

Drag and lift forces are caused by the<br />

pressure differences exerted on the stationary<br />

object by the flowing fluid. Skin friction between<br />

the fluid and the object contributes to the drag<br />

force but in many cases can be neglected. The<br />

measurement <strong>of</strong> the pressure distribution existing<br />

around a stationary cylinder in an air stream to<br />

find the drag force is the object <strong>of</strong> this<br />

experiment.<br />

Consider a circular cylinder immersed in a<br />

uniform flow. The streamlines about the cylinder<br />

are shown in Figure 9.1. The fluid exerts pressure<br />

on the front half <strong>of</strong> the cylinder in an amount<br />

that is greater than that exerted on the rear<br />

half. The difference in pressure multiplied by the<br />

projected frontal area <strong>of</strong> the cylinder gives the<br />

drag force due to pressure (also known as form<br />

drag). Because this drag is due primarily to a<br />

pressure difference, measurement <strong>of</strong> the pressure<br />

distribution about the cylinder allows for finding<br />

the drag force experimentally. A typical pressure<br />

distribution is given in Figure 9.2. Shown in<br />

Figure 9.2a is the cylinder with lines and<br />

arrowheads. The length <strong>of</strong> the line at any point<br />

on the cylinder surface is proportional to the<br />

pressure at that point. The direction <strong>of</strong> the<br />

arrowhead indicates that the pressure at the<br />

respective point is greater than the free stream<br />

pressure (pointing toward the center <strong>of</strong> the<br />

cylinder) or less than the free stream pressure<br />

(pointing away). Note the existence <strong>of</strong> a<br />

separation point and a separation region (or<br />

wake). The pressure in the back flow region is<br />

nearly the same as the pressure at the point <strong>of</strong><br />

separation. The general result is a net drag force<br />

equal to the sum <strong>of</strong> the forces due to pressure<br />

acting on the front half (+) and on the rear half<br />

(-) <strong>of</strong> the cylinder. To find the drag force, it is<br />

necessary to sum the components <strong>of</strong> pressure at<br />

each point in the flow direction. Figure 9.2b is a<br />

graph <strong>of</strong> the same data as that in Figure 9.2a<br />

except that 9.2b is on a linear grid.<br />

Freestream<br />

Velocity V<br />

Stagnation<br />

Streamline<br />

Wake<br />

FIGURE 9.1. Streamlines <strong>of</strong> flow about a circular<br />

cylinder.<br />

separation<br />

point<br />

p<br />

0 30 60 90 120 150 180<br />

separation<br />

point<br />

(a) Polar Coordinate Graph<br />

(b) Linear Graph<br />

FIGURE 9.2. Pressure distribution around a circular cylinder placed in a uniform flow.<br />

24


Pressure Measurement<br />

Equipment<br />

A Wind Tunnel<br />

A Right Circular Cylinder with Pressure<br />

Taps<br />

Figure 9.3 is a schematic <strong>of</strong> a wind tunnel. It<br />

consists <strong>of</strong> a nozzle, a test section, a diffuser and a<br />

fan. Flow enters the nozzle and passes through<br />

flow straighteners and screens. The flow is<br />

directed through a test section whose walls are<br />

made <strong>of</strong> a transparent material, usually<br />

Plexiglas or glass. An object is placed in the test<br />

section for observation. Downstream <strong>of</strong> the test<br />

section is the diffuser followed by the fan. In the<br />

tunnel that is used in this experiment, the test<br />

section is rectangular and the fan housing is<br />

circular. Thus one function <strong>of</strong> the diffuser is to<br />

gradually lead the flow from a rectangular<br />

section to a circular one.<br />

Figure 9.4 is a schematic <strong>of</strong> the side view <strong>of</strong><br />

the circular cylinder. The cylinder is placed in<br />

the test section <strong>of</strong> the wind tunnel which is<br />

operated at a preselected velocity. The pressure<br />

tap labeled as #1 is placed at 0° directly facing<br />

the approach flow. The pressure taps are<br />

attached to a manometer board. Only the first 18<br />

taps are connected because the expected pr<strong>of</strong>ile is<br />

symmetric about the 0° line. The manometers will<br />

provide readings <strong>of</strong> pressure at 10° intervals<br />

about half the cylinder. For two different<br />

approach velocities, measure and record the<br />

pressure distribution about the circular cylinder.<br />

Plot the pressure distribution on polar coordinate<br />

graph paper for both cases. Also graph pressure<br />

difference (pressure at the point <strong>of</strong> interest minus<br />

the free stream pressure) as a function <strong>of</strong> angle θ<br />

on linear graph paper. Next, graph ∆p cosθ vs θ<br />

(horizontal axis) on linear paper and determine<br />

the area under the curve by any convenient<br />

method (counting squares or a numerical<br />

technique).<br />

The drag force can be calculated by<br />

integrating the flow-direction-component <strong>of</strong> each<br />

pressure over the area <strong>of</strong> the cylinder:<br />

π<br />

D f<br />

= 2RL<br />

0<br />

∫ ∆p cosθdθ<br />

The above expression states that the drag force is<br />

twice the cylinder radius (2R) times the cylinder<br />

length (L) times the area under the curve <strong>of</strong> ∆p<br />

cosθ vs θ.<br />

Drag data are usually expressed as drag<br />

coefficient C D<br />

vs Reynolds number Re. The drag<br />

coefficient is defined as<br />

D f<br />

C D<br />

=<br />

ρV 2 A/2<br />

The Reynolds number is<br />

Re = ρVD<br />

µ<br />

inlet flow<br />

straighteners<br />

nozzle<br />

test section<br />

diffuser<br />

fan<br />

FIGURE 9.3. A schematic <strong>of</strong> the wind tunnel used in this experiment.<br />

25


where V is the free stream velocity (upstream <strong>of</strong><br />

the cylinder), A is the projected frontal area <strong>of</strong><br />

the cylinder (2RL), D is the cylinder diameter, ρ<br />

is the air density and µ is the air viscosity.<br />

Compare the results to those found in texts.<br />

60<br />

90<br />

120<br />

0<br />

30<br />

static pressure<br />

taps attach to<br />

manometers<br />

150<br />

180<br />

FIGURE 9.4. Schematic <strong>of</strong> the experimental<br />

apparatus used in this experiment.<br />

26


EXPERIMENT 10<br />

DRAG FORCE DETERMINATION<br />

An object placed in a uniform flow is acted<br />

upon by various forces. The resultant <strong>of</strong> these<br />

forces can be resolved into two force components,<br />

parallel and perpendicular to the main flow<br />

direction. The component acting parallel to the<br />

flow is known as the drag force. It is a function <strong>of</strong><br />

a skin friction effect and an adverse pressure<br />

gradient. The component perpendicular to the<br />

flow direction is the lift force and is caused by a<br />

pressure distribution which results in a lower<br />

pressure acting over the top surface <strong>of</strong> the object<br />

than at the bottom. If the object is symmetric<br />

with respect to the flow direction, then the lift<br />

force will be zero and only a drag force will exist.<br />

Measurement <strong>of</strong> the drag force acting on an object<br />

immersed in the uniform flow <strong>of</strong> a fluid is the<br />

subject <strong>of</strong> this experiment.<br />

Equipment<br />

Subsonic Wind Tunnel<br />

Objects<br />

A description <strong>of</strong> a subsonic wind tunnel is<br />

given in Experiment 9 and is shown schematically<br />

in Figure 9.3. The fan at the end <strong>of</strong> the tunnel<br />

draws in air at the inlet. An object is mounted on a<br />

stand that is pre calibrated to read lift and drag<br />

forces exerted by the fluid on the object. A<br />

schematic <strong>of</strong> the test section is shown in Figure<br />

10.1. The velocity <strong>of</strong> the flow at the test section is<br />

also pre calibrated. The air velocity past the<br />

object can be controlled by changing the angle <strong>of</strong><br />

the inlet vanes located within the fan housing.<br />

Thus air velocity, lift force and drag force are<br />

read directly from the tunnel instrumentation.<br />

There are a number <strong>of</strong> objects that are<br />

available for use in the wind tunnel. These<br />

include a disk, a smooth surfaced sphere, a rough<br />

surface sphere, a hemisphere facing upstream,<br />

and a hemisphere facing downstream. For<br />

whichever is assigned, measure drag on the object<br />

as a function <strong>of</strong> velocity.<br />

Data on drag vs velocity are usually graphed<br />

in dimensionless terms. The drag force D f is<br />

customarily expressed in terms <strong>of</strong> the drag<br />

coefficient C D (a ratio <strong>of</strong> drag force to kinetic<br />

energy):<br />

in which ρ is the fluid density, V is the free<br />

stream velocity, and A is the projected frontal<br />

area <strong>of</strong> the object. Traditionally, the drag<br />

coefficient is graphed as a function <strong>of</strong> the<br />

Reynolds number, which is defined as<br />

Re = VD<br />

ν<br />

where D is a characteristic length <strong>of</strong> the object<br />

and ν is the kinematic viscosity <strong>of</strong> the fluid. For<br />

each object assigned, graph drag coefficient vs<br />

Reynolds number and compare your results to<br />

those published in texts. Use log-log paper if<br />

appropriate.<br />

Questions<br />

1. How does the mounting piece affect the<br />

readings?<br />

2. How do you plan to correct for its effect, if<br />

necessary?<br />

uniform flow<br />

lift force<br />

measurement<br />

object<br />

mounting stand<br />

drag force<br />

measurement<br />

FIGURE 10.1. Schematic <strong>of</strong> an object mounted in<br />

the test section <strong>of</strong> the wind tunnel.<br />

D f<br />

C D<br />

=<br />

ρV 2 A/2<br />

27


EXPERIMENT 11<br />

ANALYSIS OF AN AIRFOIL<br />

A wing placed in the uniform flow <strong>of</strong> an<br />

airstream will experience lift and drag forces.<br />

Each <strong>of</strong> these forces is due to a pressure<br />

difference. The lift force is due to the pressure<br />

difference that exists between the lower and<br />

upper surfaces. This phenomena is illustrated in<br />

Figure 11.1. As indicated the airfoil is immersed<br />

in a uniform flow. If pressure could be measured at<br />

selected locations on the surface <strong>of</strong> the wing and<br />

the results graphed, the pr<strong>of</strong>ile in Figure 11.1<br />

would result. Each pressure measurement is<br />

represented by a line with an arrowhead. The<br />

length <strong>of</strong> each line is proportional to the<br />

magnitude <strong>of</strong> the pressure at the point. The<br />

direction <strong>of</strong> the arrow (toward the horizontal<br />

axis or away from it) represents whether the<br />

pressure at the point is less than or greater than<br />

the free stream pressure measured far upstream <strong>of</strong><br />

the wing.<br />

Experiment I<br />

Mount the wing with pressure taps in the<br />

tunnel and attach the tube ends to manometers.<br />

Select a wind speed and record the pressure<br />

distribution for a selected angle <strong>of</strong> attack (as<br />

assigned by the instructor). Plot pressure vs chord<br />

length as in Figure 11.1, showing the vertical<br />

component <strong>of</strong> each pressure acting on the upper<br />

surface and on the lower surface. Determine<br />

where separation occurs for each case.<br />

Mount the second wing on the lift and drag<br />

balance (Figure 11.2). For the same wind speed<br />

and angle <strong>of</strong> attack, measure lift and drag exerted<br />

on the wing.<br />

lift<br />

c<br />

drag<br />

c<br />

uniform flow<br />

mounting stand<br />

stagnation<br />

point<br />

C p<br />

pressure<br />

coefficient<br />

negative pressure<br />

gradient on upper<br />

surface<br />

lift force<br />

measurement<br />

drag force<br />

measurement<br />

stagnation<br />

point<br />

chord, c<br />

positive pressure<br />

on lower surface<br />

FIGURE 11.2. Schematic <strong>of</strong> lift and drag<br />

measurement in a test section.<br />

FIGURE 11.1. Streamlines <strong>of</strong> flow about a wing<br />

and the resultant pressure distribution.<br />

Lift and Drag Measurements for a Wing<br />

Equipment<br />

Wind Tunnel (See Figure 9.3)<br />

Wing with Pressure Taps<br />

Wing for Attachment to Lift & Drag<br />

Instruments (See Figure 11.2)<br />

The wing with pressure taps provided<br />

pressure at selected points on the surface <strong>of</strong> the<br />

wing. Use the data obtained and sum the<br />

horizontal component <strong>of</strong> each pressure to obtain<br />

the drag force. Compare to the results obtained<br />

with the other wing. Use the data obtained and<br />

sum the vertical component <strong>of</strong> each pressure to<br />

obtain the lift force. Compare the results<br />

obtained with the other wing. Calculate %<br />

errors.<br />

28


Experiment II<br />

For a number <strong>of</strong> wings, lift and drag data<br />

vary only slightly with Reynolds number and<br />

therefore if lift and drag coefficients are graphed<br />

as a function <strong>of</strong> Reynolds number, the results are<br />

not that meaningful. A more significant<br />

representation <strong>of</strong> the results is given in what is<br />

known as a polar diagram for the wing. A polar<br />

diagram is a graph on a linear grid <strong>of</strong> lift<br />

coefficient (vertical axis) as a function <strong>of</strong> drag<br />

coefficient. Each data point on the graph<br />

corresponds to a different angle <strong>of</strong> attack, all<br />

measured at one velocity (Reynolds number).<br />

Referring to Figure 11.2 (which is the<br />

experimental setup here), the angle <strong>of</strong> attack α is<br />

measured from a line parallel to the chord c to a<br />

line that is parallel to the free stream velocity.<br />

If so instructed, obtain lift force, drag force and<br />

angle <strong>of</strong> attack data using a pre selected velocity.<br />

Allow the angle <strong>of</strong> attack to vary from a negative<br />

angle to the stall point and beyond. Obtain data<br />

at no less than 9 angles <strong>of</strong> attack. Use the data to<br />

produce a polar diagram.<br />

Analysis<br />

Lift and drag data are usually expressed in<br />

dimensionless terms using lift coefficient and drag<br />

coefficient. The lift coefficient is defined as<br />

L f<br />

C L<br />

=<br />

ρV 2 A/2<br />

where L f is the lift force, ρ is the fluid density, V<br />

is the free stream velocity far upstream <strong>of</strong> the<br />

wing, and A is the area <strong>of</strong> the wing when seen<br />

from a top view perpendicular to the chord<br />

length c. The drag coefficient is defined as<br />

D f<br />

C D<br />

=<br />

ρV 2 A/2<br />

in which D f<br />

is the drag force.<br />

29


EXPERIMENT 12<br />

OPEN CHANNEL FLOW—SLUICE GATE<br />

Liquid motion in a duct where a surface <strong>of</strong> the<br />

fluid is exposed to the atmosphere is called open<br />

channel flow. In the laboratory, open channel<br />

flow experiments can be used to simulate flow in a<br />

river, in a spillway, in a drainage canal or in a<br />

sewer. Such modeled flows can include flow over<br />

bumps or through dams, flow through a venturi<br />

flume or under a partially raised gate (a sluice<br />

gate). The last example, flow under a sluice gate,<br />

is the subject <strong>of</strong> this experiment.<br />

Flow Through a Sluice Gate<br />

Equipment<br />

Open Channel Flow Apparatus<br />

Sluice Gate Model<br />

Figure 12.1 shows a schematic <strong>of</strong> the side<br />

view <strong>of</strong> the sluice gate. Flow upstream <strong>of</strong> the gate<br />

has a depth h o while downstream the depth is h.<br />

The objective <strong>of</strong> the analysis is to formulate an<br />

equation to relate the volume flow rate through<br />

(or under) the gate to the upstream and<br />

downstream depths.<br />

sluice gate<br />

p atm<br />

h o<br />

hand crank<br />

direction <strong>of</strong><br />

movement<br />

h<br />

p atm<br />

FIGURE 12.1. Schematic <strong>of</strong> flow under a sluice<br />

gate.<br />

The flow rate through the gate is maintained at<br />

nearly a constant value. For various raised<br />

positions <strong>of</strong> the sluice gate, different liquid<br />

heights h o and h will result. Applying the<br />

Bernoulli equation to flow about the gate gives<br />

p 0<br />

ρg + V 0 2<br />

2g + h 0 = p ρg + V2<br />

2g + h<br />

Pressures at the free surface are both equal to<br />

atmospheric pressure, so they cancel. Rearranging<br />

gives<br />

h 0 = V2<br />

2g - V 0 2<br />

2g +h<br />

In terms <strong>of</strong> flow rate, the velocities are written as<br />

V 0 = Q A = Q<br />

bh 0<br />

V = Q bh<br />

where b is the channel width at the gate.<br />

Substituting into the Bernoulli Equation and<br />

simplifying gives<br />

h 0 = Q2<br />

2gb 2 ⎝ ⎛ 1<br />

h ⎠ ⎞ 2 - 1<br />

h<br />

2 + h<br />

0<br />

Dividing by h 0 ,<br />

Q<br />

1 =<br />

2<br />

2gb 2 h 0 ⎝ ⎛ 1<br />

h ⎠ ⎞<br />

2 - 1<br />

h<br />

2 + h 0 h 0<br />

Rearranging further,<br />

Q 2<br />

⎛1 - h ⎞ =<br />

⎝ h 0 ⎠ 2gb 2 h 2 h 0 ⎝ ⎛ 1 - h 2<br />

h ⎠ ⎞ 2 0<br />

Multiplying both sides by h 2 /h 0 2 , and continuing<br />

to simplify, we finally obtain<br />

h 2 /h<br />

2 0 Q<br />

=<br />

2<br />

1 + h/h 0 2gb 2 h<br />

3 0<br />

here Q is the theoretical volume flow rate. The<br />

right hand side <strong>of</strong> this equation is recognized as<br />

1/2 <strong>of</strong> the upstream Froude number. So by<br />

measuring the depth <strong>of</strong> liquid before and after<br />

the sluice gate, the theoretical flow rate can be<br />

calculated with the above equation. The<br />

theoretical flow rate can then be compared to the<br />

actual flow rate obtained by measurements using<br />

the orifice meters.<br />

For 9 different raised positions <strong>of</strong> the sluice<br />

gate, measure the upstream and downstream<br />

depths and calculate the actual flow rate. In<br />

addition, calculate the upstream Froude number<br />

for each case and determine its value for<br />

maximum flow conditions. Graph h/h 0 (vertical<br />

30


axis) versus (Q 2 /b 2 h 0 3 g). Determine h/h 0<br />

corresponding to maximum flow. Note that h/h 0<br />

varies from 0 to 1.<br />

Figure 12.2 is a sketch <strong>of</strong> the open channel<br />

flow apparatus. It consists <strong>of</strong> a sump tank with a<br />

pump/motor combination on each side. Each pump<br />

draws in water from the sump tank and<br />

discharges it through the discharge line to<br />

calibrated orifice meters and then to the head<br />

tank. Each orifice meter is connected to its own<br />

manometer. Use <strong>of</strong> the calibration curve<br />

(provided by the instructor) allows for finding<br />

the actual flow rate into the channel. The head<br />

tank and flow channel have sides made <strong>of</strong><br />

Plexiglas. Water flows downstream in the<br />

channel past the object <strong>of</strong> interest (in this case a<br />

sluice gate) and then is routed back to the sump<br />

tank.<br />

Questions<br />

1. For the required report, derive the sluice<br />

gate equation in detail.<br />

2. What if it was assumed that V 0


EXPERIMENT 13<br />

OPEN CHANNEL FLOW OVER A WEIR<br />

Flow meters used in pipes introduce an<br />

obstruction into the flow which results in a<br />

measurable pressure drop that in turn is related to<br />

the volume flow rate. In an open channel, flow<br />

rate can be measured similarly by introducing an<br />

obstruction into the flow. A simple obstruction,<br />

called a weir, consists <strong>of</strong> a vertical plate<br />

extending the entire width <strong>of</strong> the channel. The<br />

plate may have an opening, usually rectangular,<br />

trapezoidal, or triangular. Other configurations<br />

exist and all are about equally effective. The use<br />

<strong>of</strong> a weir to measure flow rate in an open channel<br />

is the subject <strong>of</strong> this experiment.<br />

Flow Over a Weir<br />

Equipment<br />

Open Channel Flow Apparatus (See<br />

Figure 12.1)<br />

Several Weirs<br />

The open channel flow apparatus allows for<br />

the insertion <strong>of</strong> a weir and measurement <strong>of</strong> liquid<br />

depths. The channel is fed by two centrifugal<br />

pumps. Each pump has a discharge line which<br />

contains an orifice meter attached to a<br />

manometer. The pressure drop reading from the<br />

manometers and a calibration curve provide the<br />

means for determining the actual flow rate into<br />

the channel.<br />

Figure 13.1 is a sketch <strong>of</strong> the side and<br />

upstream view <strong>of</strong> a 90 degree (included angle) V-<br />

notch weir. Analysis <strong>of</strong> this weir is presented<br />

here for illustrative purposes. Note that<br />

upstream depth measurements are made from the<br />

lowest point <strong>of</strong> the weir over which liquid flows.<br />

This is the case for the analysis <strong>of</strong> all<br />

conventional weirs. A coordinate system is<br />

imposed whose origin is at the intersection <strong>of</strong> the<br />

free surface and a vertical line extending upward<br />

from the vertex <strong>of</strong> the V-notch. We select an<br />

element that is dy thick and extends the entire<br />

width <strong>of</strong> the flow cross section. The velocity <strong>of</strong><br />

the liquid through this element is found by<br />

applying Bernoulli's equation:<br />

p a<br />

ρ + V o 2<br />

2 + gh = p a<br />

ρ + V2<br />

+ g(h - y)<br />

2<br />

Note that in pipe flow, pressure remained in the<br />

equation when analyzing any <strong>of</strong> the differential<br />

pressure meters (orifice or venturi meters). In open<br />

channel flows, the pressure terms represents<br />

atmospheric pressure and cancel from the<br />

Bernoulli equation. The liquid height is<br />

therefore the only measurement required here.<br />

From the above equation, assuming V o<br />

negligible:<br />

V = √⎺⎺⎺2gy (13.1)<br />

Equation 13.1 is the starting point in the analysis<br />

<strong>of</strong> all weirs. The incremental flow rate <strong>of</strong> liquid<br />

through layer dy is:<br />

dQ = 2Vxdy = √⎺⎺⎺2gy(2x)dy<br />

From the geometry <strong>of</strong> the V-notch and with<br />

respect to the coordinate axes, we have y = h - x.<br />

V o<br />

p a<br />

h<br />

V<br />

p a<br />

y axis<br />

y<br />

dy<br />

x<br />

FIGURE 13.1. Side and upstream views <strong>of</strong> a 90° V-notch weir.<br />

x axis<br />

32


Therefore,<br />

Q = ∫<br />

0<br />

Integration gives<br />

h<br />

(2√⎺⎺2g)y 1/2 (h - y)dy<br />

Q th = 8 15 √⎺⎺ 2g h 5/2 =Ch 5/2 (13.2)<br />

where C is a constant. The above equation<br />

represents the ideal or theoretical flow rate <strong>of</strong><br />

liquid over the V-notch weir. The actual<br />

discharge rate is somewhat less due to frictional<br />

and other dissipative effects. As with pipe<br />

meters, we introduce a discharge coefficient<br />

defined as:<br />

C' = Q ac<br />

Q th<br />

The equation that relates the actual volume flow<br />

rate to the upstream height then is<br />

Q ac<br />

= C'Ch 5/2<br />

It is convenient to combine the effects <strong>of</strong> the<br />

constant C and the coefficient C’ into a single<br />

coefficient C vn for the V-notch weir. Thus we<br />

reformulate the previous two equations to obtain:<br />

C vn ≈ Q ac<br />

Q th<br />

(13.3)<br />

Q ac<br />

= C vn h 5/2 (13.4)<br />

Each type <strong>of</strong> weir will have its own coefficient.<br />

Calibrate each <strong>of</strong> the weirs assigned by the<br />

instructor for 7 different upstream height<br />

measurements. Use the flow rate chart provided<br />

with the open channel flow apparatus to obtain<br />

the actual flow rate. Derive an appropriate<br />

equation for each weir used (similar to Equation<br />

13.4) above. Determine the coefficient applicable<br />

for each weir tested. List the assumptions made<br />

in each derivation. Discuss the validity <strong>of</strong> each<br />

assumption, pointing out where they break down.<br />

Graph upstream height vs actual and theoretical<br />

volume flow rates. Plot the coefficient <strong>of</strong><br />

discharge (as defined in Equation 13.3) as a<br />

function <strong>of</strong> the upstream Froude number.<br />

FIGURE 13.2. Other types <strong>of</strong> weirs–semicircular, contracted and suppressed.<br />

33


EXPERIMENT 14<br />

OPEN CHANNEL FLOW—HYDRAULIC JUMP<br />

When spillways or other similar open<br />

channels are opened by the lifting <strong>of</strong> a gate,<br />

liquid passing below the gate has a high velocity<br />

and an associated high kinetic energy. Due to the<br />

erosive properties <strong>of</strong> a high velocity fluid, it<br />

may be desirable to convert the high kinetic<br />

energy (e.g. high velocity) to a high potential<br />

energy (e.g., a deeper stream). The problem then<br />

becomes one <strong>of</strong> rapidly varying the liquid depth<br />

over a short channel length. Rapidly varied flow<br />

<strong>of</strong> this type produces what is known as a<br />

hydraulic jump.<br />

Consider a horizontal, rectangular open<br />

channel <strong>of</strong> width b, in which a hydraulic jump<br />

has developed. Figure 14.1 shows a side view <strong>of</strong> a<br />

hydraulic jump. Figure 14.1 also shows the depth<br />

<strong>of</strong> liquid upstream <strong>of</strong> the jump to be h 1<br />

, and a<br />

downstream depth <strong>of</strong> h 2<br />

. Pressure distributions<br />

upstream and downstream <strong>of</strong> the jump are drawn<br />

in as well. Because the jump occurs over a very<br />

short distance, frictional effects can be neglected.<br />

A force balance would therefore include only<br />

pressure forces. Applying the momentum equation<br />

in the flow direction gives:<br />

p 1<br />

A 1<br />

- p 2<br />

A 2<br />

= ρQ(V 2<br />

- V 1<br />

)<br />

Pressure in the above equation represents the<br />

pressure that exists at the centroid <strong>of</strong> the cross<br />

section. Thus p = ρg(h/2). With a rectangular<br />

cross section <strong>of</strong> width b (A = bh), the above<br />

equation becomes<br />

h 1<br />

g<br />

2 (h 1 b) - h 2 g<br />

2 (h 2 b) = Q(V 2 - V 1 )<br />

From continuity, A 1<br />

V 1<br />

= A 2<br />

V 2<br />

= Q. Combining and<br />

rearranging,<br />

h<br />

2<br />

1<br />

- h<br />

2<br />

2<br />

= Q 2<br />

2 gb 2 ⎝ ⎛ 1<br />

- 1 h ⎠ ⎞<br />

2<br />

h 1<br />

Simplifying,<br />

h 2<br />

2<br />

+ h 2<br />

h 1<br />

- 2<br />

Q 2<br />

gb 2 h 1<br />

= 0<br />

Solving for the downstream height yields one<br />

physically (nonnegative) possible solution:<br />

h 2<br />

= - h 1<br />

2 √⎺⎺⎺⎺<br />

+ 2Q 2<br />

4<br />

gb 2 h 1<br />

+ h 1 2<br />

from which the downstream height can be found.<br />

By applying Bernoulli’s Equation along the free<br />

surface, the energy lost irreversibly can be<br />

calculated as<br />

Lost Energy = E =<br />

g(h 2<br />

- h 1<br />

) 3<br />

4h 2<br />

h 1<br />

and the rate <strong>of</strong> energy loss is<br />

dW<br />

dt = ρQE<br />

The above equations are adequate to properly<br />

describe a hydraulic jump.<br />

Hydraulic Jump Measurements<br />

Equipment<br />

Open Channel Flow Apparatus (Figure 12.1)<br />

The channel can be used in either a<br />

horizontal or a sloping configuration. The device<br />

contains two pumps which discharge water<br />

through calibrated orifice meters connected to<br />

manometers. The device also contains on the<br />

channel bottom two forward facing brass tubes.<br />

Each tube is connected to a vertical Plexiglas<br />

tube. The height <strong>of</strong> the water in either <strong>of</strong> these<br />

tubes represents the energy level at the<br />

respective tube location. The difference in height<br />

is the actual lost energy (E) for the jump <strong>of</strong><br />

interest.<br />

FIGURE 14.1. Schematic <strong>of</strong> a<br />

hydraulic jump in an open<br />

channel.<br />

h 1<br />

V 2<br />

V 1<br />

p 1<br />

p 2<br />

h 2<br />

34


Develop a hydraulic jump in the channel;<br />

record upstream and downstream heights,<br />

manometer readings (from which the actual<br />

volume flow rate is obtained) and the lost energy<br />

E. By varying the flow rate, upstream height,<br />

downstream height and/or the channel slope,<br />

record measurements on different jumps. Derive<br />

the applicable equations in detail and substitute<br />

appropriate values to verify the predicted<br />

downstream height and lost energy. In other<br />

words, the downstream height <strong>of</strong> each jump is to<br />

be measured and compared to the downstream<br />

height calculated with Equation 14.1. The same<br />

is to be done for the rate <strong>of</strong> energy loss (Equation<br />

14.2).<br />

Analysis<br />

Data on a hydraulic jump is usually specified<br />

in two ways both <strong>of</strong> which will be required for<br />

the report. Select any <strong>of</strong> the jumps you have<br />

measurements for and construct a momentum<br />

diagram . A momentum diagram is a graph <strong>of</strong><br />

liquid depth on the vertical axis vs momentum on<br />

the horizontal axis. The momentum <strong>of</strong> the flow is<br />

given by:<br />

M = 2Q2<br />

gb 2 h + h2<br />

4<br />

Another significant graph <strong>of</strong> hydraulic jump<br />

data is <strong>of</strong> depth ratio h 2 /h 1 (vertical axis) as a<br />

function <strong>of</strong> the upstream Froude number, Fr 1 (=<br />

Q 2 /gb 2 h 1<br />

3<br />

). Construct such a graph for any <strong>of</strong> the<br />

jumps for which you have taken measurements.<br />

35


EXPERIMENT 15<br />

OPEN CHANNEL FLOW OVER A HUMP<br />

Flow over a hump in an open channel is a<br />

problem that can be successfully modeled in order<br />

to make predictions about the behavior <strong>of</strong> the<br />

fluid. This experiment involves making<br />

appropriate measurements for such a system, and<br />

relating flow rate to critical depth. The flow<br />

rate, critical depth, and specific energy are<br />

determined theoretically and experimentally.<br />

Theory<br />

Flow in a channel is modeled in terms <strong>of</strong> a<br />

parameter called the specific energy head (or just<br />

specific energy) <strong>of</strong> the flow, E. The specific<br />

energy head is defined as<br />

Q 2<br />

E = h +<br />

2gh 2 b 2 (15.1)<br />

where h is the depth <strong>of</strong> the flow, Q is the volume<br />

flow rate, g is gravity, and b is the channel<br />

width. The dimension <strong>of</strong> the specific energy head<br />

is L (ft or m).<br />

Figure 15.1 is a sketch <strong>of</strong> flow over a hump,<br />

with flow from left to right. Shown is the channel<br />

bed and the hump. Upstream <strong>of</strong> the hump<br />

(subscript 1 notation), the flow is subcritical;<br />

downstream (subscript 2) the flow is supercritical.<br />

Just at the highest point <strong>of</strong> the hump,<br />

the flow is critical (subscript c). Also shown in<br />

the figure is the total energy line, which we<br />

assume is parallel to the flow channel bed; i.e.,<br />

the total energy remains a constant in the flow.<br />

Upstream <strong>of</strong> the hump, the total specific<br />

energy head <strong>of</strong> the flow is denoted as E 1 , and the<br />

depth <strong>of</strong> the liquid is h 1 , as shown graphically in<br />

Figure 15.1. At any location z on the hump before<br />

z c , the energy head is E, and the depth is h. At<br />

this same height z downstream <strong>of</strong> z c , the liquid<br />

depth is h’, but the energy head is still E. At the<br />

highest point <strong>of</strong> the hump z c , the energy head is<br />

E c and the liquid depth is h c . The total specific<br />

energy head and the liquid depth anywhere are<br />

related according to Equation 15.1.<br />

total energy line<br />

flow<br />

direction<br />

h<br />

E<br />

h c<br />

h'<br />

h 2<br />

E 1<br />

h 1<br />

z<br />

E c<br />

z c<br />

hump<br />

channel bed<br />

FIGURE 15.1. Flow over a<br />

hump in an open<br />

channel.<br />

We can illustrate the relationship between<br />

these parameters graphically by drawing a<br />

specific energy head diagram, as illustrated in<br />

Figure 15.2. This graph has flow depth on the<br />

vertical axis and specific energy head on the<br />

horizontal axis. The condition <strong>of</strong> the flow is<br />

represented by the solid line with arrows<br />

showing how the flow changes from subcritical to<br />

supercritical. At the location on the hump where<br />

the height is z, the energy head is E. We draw a<br />

vertical line at this value <strong>of</strong> the specific energy<br />

head; it will intersect the line at h (upstream)<br />

and h’ (downstream).<br />

depth h<br />

h 1<br />

h<br />

z<br />

z c<br />

h c<br />

h'<br />

h 2<br />

E c E E 1 , E 2<br />

specific energy head E<br />

FIGURE 15.2. Specific energy diagram.<br />

subcritical<br />

supercritical<br />

36


At any upstream (<strong>of</strong> the hump) location, say<br />

h 1 , we see that the corresponding specific energy<br />

head is E 1 . The vertical line that locates E 1 also<br />

locates the energy E 2 which is downstream <strong>of</strong> the<br />

hump. A vertical line drawn at E 1 intersects the<br />

line at h 1 and h 2 , which are the upstream and<br />

downstream liquid heights, respectively. Note<br />

that the minimum specific energy head is at the<br />

highest point <strong>of</strong> the hump z c , and the energy<br />

head there is E c .<br />

As water flows over the hump, the initial<br />

specific energy head E 1 is reduced to a value E by<br />

an amount equal to the height <strong>of</strong> the hump. So at<br />

any location along the hump, the specific energy<br />

head is E 1 - z, where z is the elevation above the<br />

channel bed. At the point where the flow is<br />

critical, the critical depth h c is given by<br />

Q 2<br />

1/3 2E<br />

h c = ⎛ ⎞ = c<br />

⎝ b 2 g⎠<br />

3<br />

Flow Over a Hump<br />

(15.2)<br />

Equipment<br />

Open Channel Flow Apparatus (Figure<br />

12.1)<br />

Installed hump<br />

The open channel flow apparatus is described<br />

in Experiment 12 and illustrated schematically in<br />

Figure 12.1. Adjust the channel so that it is<br />

horizontal. Make every effort to minimize<br />

leakage <strong>of</strong> water past the sides <strong>of</strong> the hump.<br />

Start both pumps and adjust the valves to give a<br />

smooth water surface pr<strong>of</strong>ile over the hump. For<br />

one set <strong>of</strong> conditions, take readings from the<br />

manometers to determine the volume flow rate<br />

over the hump.<br />

The open channel flow apparatus has a<br />

depth gage attached. It will be necessary to<br />

measure the water depth at certain specific<br />

locations on or about the hump. These locations<br />

are shown in Figure 15.3 (dimensions are in feet).<br />

There are 8 water depths to be measured. So for<br />

one flow rate, two manometer readings and 8<br />

water depths will be recorded. Gather data for<br />

the assigned number <strong>of</strong> flow rates.<br />

Results<br />

Although the data taken in this experiment<br />

seem simple, the calculations required to reduce<br />

the data appropriately can occupy much time.<br />

With the data obtained:<br />

Determine the flow rate using the manometer<br />

readings. This value will be referred to as the<br />

actual flow rate Q AC (subscript AC will refer<br />

to an actual value, while TH refers to a<br />

theoretical value).<br />

Calculate the flow rate using a rearranged<br />

form <strong>of</strong> Equation 15.2. This value will be<br />

referred to as the theoretical flow rate Q TH .<br />

Compare the two flow rates and find % error.<br />

Use Equation 15.2 to find the value <strong>of</strong> the<br />

critical depth using Q AC. Compare this value<br />

to the measured value, and find % error.<br />

Calculate the theoretical and actual values<br />

<strong>of</strong> the minimum energy E c using Equation 15.2.<br />

Compare the results.<br />

Calculate the actual specific energy head E AC<br />

at each measurement station using Equation<br />

15.1. Determine also the total energy head<br />

H AC (= E AC + z) for all readings.<br />

Compose a chart using the column and row<br />

headings shown in Table 15.1.<br />

1 2 3 4 5 6 7 8<br />

flow<br />

direction<br />

hump<br />

2<br />

0.276<br />

0.313<br />

0.313 0.313 0.313<br />

1<br />

FIGURE 15.3. Water depth<br />

measurement locations<br />

for flow over a hump.<br />

(Dimensions in feet.)<br />

37


TABLE 15.1. Data reduction table for flow over a hump.<br />

Station 1 2 3 4 5 6 7 8<br />

Depth <strong>of</strong> flow h AC in ft<br />

Specific energy head E AC in ft<br />

Height <strong>of</strong> hump above channel bed z<br />

in ft<br />

Total energy head H AC in ft<br />

Construct a graph <strong>of</strong> the flow configuration.<br />

On the horizontal axis, plot distance<br />

downstream, and plot depth on the vertical<br />

axis. On this set <strong>of</strong> axes, plot (a) the total<br />

energy line (H AC ); (b) the water surface<br />

pr<strong>of</strong>ile; and, (c) the elevation z. Show data<br />

points on the graph.<br />

Construct a specific energy head diagram<br />

similar to that <strong>of</strong> Figure 15.2. Show the<br />

theoretical results (based on Q TH ), and show<br />

the actual data points.<br />

Derive Equation 2.<br />

Questions<br />

1. What is the value <strong>of</strong> the Froude number (a)<br />

upstream <strong>of</strong> the hump, (b) at the highest<br />

point <strong>of</strong> the hump, and (c) downstream <strong>of</strong> the<br />

hump?<br />

2. Is the Froude number used in finding the<br />

critical depth in Equation 15.2?<br />

3. What equations is used to develop the<br />

expression for specific energy head (Equation<br />

15.1)?<br />

4. How is the second term in Equation 15.1 (i.e.;<br />

Q 2 /2gh 2 b 2 ) related to the Froude number?<br />

5. Is the total energy line (H AC ) a constant as we<br />

assumed with reference to Figure 15.1, or does<br />

it change?<br />

38


EXPERIMENT 16<br />

MEASUREMENT OF VELOCITY<br />

AND<br />

CALIBRATION OF A METER FOR COMPRESSIBLE FLOW<br />

The objective <strong>of</strong> this experiment is to<br />

determine a calibration curve for a meter placed<br />

in a pipe that is conveying air. The meters <strong>of</strong><br />

interest are an orifice meter and a venturi meter.<br />

These meters are calibrated in this experiment by<br />

using a pitot-static tube to measure the velocity,<br />

from which the flow rate is calculated.<br />

Pitot Static Tube<br />

When a fluid flows through a pipe, it exerts<br />

pressure that is made up <strong>of</strong> static and dynamic<br />

components. The static pressure is indicated by a<br />

measuring device moving with the flow or that<br />

causes no velocity change in the flow. Usually, to<br />

measure static pressure, a small hole<br />

perpendicular to the flow is drilled through the<br />

container wall and connected to a manometer (or<br />

pressure gage) as indicated in Figure 16.1.<br />

The dynamic pressure is due to the movement<br />

<strong>of</strong> the fluid. The dynamic pressure and the static<br />

pressure together make up the total or stagnation<br />

pressure. The stagnation pressure can be measured<br />

in the flow with a pitot tube. The pitot tube is an<br />

open ended tube facing the flow directly. Figure<br />

16.1 gives a sketch <strong>of</strong> the measurement <strong>of</strong><br />

stagnation pressure.<br />

flow<br />

stagnation pressure<br />

measurement<br />

static pressure<br />

measurement<br />

pitot tube<br />

h<br />

FIGURE 16.1. Measurement <strong>of</strong> static and<br />

stagnation pressures.<br />

The pitot-static tube combines the effects <strong>of</strong><br />

static and stagnation pressure measurement into<br />

one device. Figure 16.2 is a schematic <strong>of</strong> the pitotstatic<br />

tube. It consists <strong>of</strong> a tube within a tube<br />

which is placed in the duct facing upstream. The<br />

pressure tap that faces the flow directly gives a<br />

measurement <strong>of</strong> the stagnation pressure, while<br />

h<br />

the tap that is perpendicular to the flow gives<br />

the static pressure.<br />

When the pitot-static tube is immersed in the<br />

flow <strong>of</strong> a fluid, the pressure difference<br />

(stagnation minus static) can be read directly<br />

using a manometer and connecting the pressure<br />

taps to each leg. Applying the Bernoulli equation<br />

between the two pressure taps yields:<br />

four to eight holes<br />

equally spaced<br />

flow direction<br />

section A-A<br />

enlarged<br />

A<br />

A<br />

manometer<br />

connections<br />

FIGURE 16.2. Schematic <strong>of</strong> a pitot-static tube.<br />

p 1<br />

ρg + V 1 2<br />

2g + z 1 = p 2<br />

ρg + V 2 2<br />

2g + z 2<br />

where state “1” as the stagnation state (which<br />

will be changed to subscript “t”), and state “2” as<br />

the static state (no subscript). Elevation<br />

differences are negligible, and at the point where<br />

stagnation pressure is measured, the velocity is<br />

zero. The Bernoulli equation thus reduces to:<br />

p t<br />

ρg = p ρg + V2<br />

2g<br />

Next, we rearrange the preceding equation and<br />

solve for velocity<br />

V =<br />

√⎺⎺⎺<br />

2(p t - p)<br />

ρ<br />

A manometer connected to the pitot-static tube<br />

would provide head loss readings ∆h given by<br />

39


∆h = p t - p<br />

ρg<br />

where density is that <strong>of</strong> the flowing fluid. So<br />

velocity in terms <strong>of</strong> head loss is<br />

V = √⎺⎺⎺⎺⎺ 2g∆h<br />

Note that this equation applies only to<br />

incompressible flows. Compressibility effects are<br />

not accounted for. Furthermore, ∆h is the head<br />

loss in terms <strong>of</strong> the flowing fluid and not in terms<br />

<strong>of</strong> the reading on the manometer.<br />

For flow in a duct, manometer readings are to<br />

be taken at a number <strong>of</strong> locations within the cross<br />

section <strong>of</strong> the flow. The velocity pr<strong>of</strong>ile is then<br />

plotted using the results. Velocities at specific<br />

points are then determined from these pr<strong>of</strong>iles.<br />

The objective here is to obtain data, graph a<br />

velocity pr<strong>of</strong>ile and then determine the average<br />

velocity.<br />

Average Velocity<br />

The average velocity is related to the flow<br />

rate through a duct as<br />

V = Q A<br />

where Q is the volume flow rate and A is the<br />

cross sectional area <strong>of</strong> the duct. We can divide<br />

the flow area into five equal areas, as shown in<br />

Figure 16.3. The velocity is to be obtained at<br />

those locations labeled in the figure. The chosen<br />

positions divide the cross section into five equal<br />

concentric areas. The flow rate through each area<br />

labeled from 1 to 5 is found as<br />

Q 1 = A 1 V 1 Q 2 = A 2 V 2<br />

Q 3 = A 3 V 3 Q 4 = A 4 V 4<br />

Q 5 = A 5 V 5<br />

0.548 R<br />

0.316 R<br />

R<br />

0.837 R<br />

0.707 R 0.949 R<br />

FIGURE 16.3. Five positions within the cross<br />

section where velocity is to be determined.<br />

The total flow rate through the entire cross<br />

section is the sum <strong>of</strong> these:<br />

5<br />

Q total = ∑Q i = A 1 V 1 + A 2 V 2 + A 3 V 3 + A 4 V 4<br />

1<br />

+ A 5 V 5<br />

or Q total = A 1 (V 1 + V 2 + V 3 + V 4 + V 5 )<br />

The total area A total is 5A 1 and so<br />

V = Q total<br />

A total<br />

= (A total/5)(V 1 + V 2 + V 3 + V 4 + V 5 )<br />

A total<br />

The average velocity then becomes<br />

V = (V 1 + V 2 + V 3 + V 4 + V 5 )<br />

5<br />

The importance <strong>of</strong> the five chosen radial<br />

positions for measuring V 1 through V 5 is now<br />

evident.<br />

Velocity Measurements<br />

Equipment<br />

Axial flow fan apparatus<br />

Pitot-static tube<br />

Manometer<br />

The fan <strong>of</strong> the apparatus is used to move air<br />

through the system at a rate that is small enough<br />

to allow the air to be considered incompressible.<br />

While the fan is on, make velocity pr<strong>of</strong>ile<br />

measurements at a selected location within the<br />

duct at a cross section that is several diameters<br />

downstream <strong>of</strong> the fan. Repeat these<br />

measurements at different fan speed settings so<br />

that 9 velocity pr<strong>of</strong>iles will result. Use the<br />

velocity pr<strong>of</strong>iles to determine the average<br />

velocity and the flow rate.<br />

Questions<br />

1. Why is it appropriate to take velocity<br />

measurements at several diameters<br />

downstream <strong>of</strong> the fan?<br />

2. Suppose the duct were divided into 6 equal<br />

areas and measurements taken at select<br />

positions in the cross section. Should the<br />

average velocity using 6 equal areas be the<br />

same as the average velocity using 5 or 4<br />

equal areas?<br />

40


Incompressible Flow Through a Meter<br />

Incompressible flow through a venturi and an<br />

orifice meter was discussed in Experiment 9. For<br />

our purposes here, we merely re-state the<br />

equations for convenience. For an air over liquid<br />

manometer, the theoretical equation for both<br />

meters is<br />

Q th = A √⎺⎺⎺⎺⎺<br />

2g∆h<br />

2<br />

(1 - D 4 2 /D 4 1 )<br />

Now for any pressure drop ∆h i , there are two<br />

corresponding flow rates: Q ac and Q th . The ratio <strong>of</strong><br />

these flow rates is the venturi discharge<br />

coefficient C v , defined as<br />

C v = Q ac<br />

Q th<br />

= 0.985<br />

for turbulent flow. The orifice discharge<br />

coefficient can be expressed in terms <strong>of</strong> the Stolz<br />

equation:<br />

C o = 0.595 9 + 0.031 2β 2.1 - 0.184β 8 +<br />

+ 0.002 9β 2.5 10<br />

⎛<br />

6 0.75<br />

⎞<br />

⎝ Re β⎠<br />

where Re = ρV oD o<br />

µ<br />

L 1 = 0<br />

L 1 = 1/D 1<br />

L 1 = 1<br />

+ 0.09L 1<br />

⎝ ⎛ β 4<br />

1 - β ⎠ ⎞ 4 - L 2 (0.003 37β 3 )<br />

= 4ρQ ac<br />

πD o µ<br />

for corner taps<br />

for flange taps<br />

for 1D & 1 2 D taps<br />

β = D o<br />

D 1<br />

β 4<br />

and if L 1 ≥ 0.433 3, the coefficient <strong>of</strong> the ⎛ ⎞<br />

⎝ 1 - β 4 ⎠<br />

term becomes 0.039.<br />

L 2 = 0 for corner taps<br />

L 2 = 1/D 1 for flange taps<br />

L 2 = 0.5 - E/D 1 for 1D & 1 2 D taps<br />

E = orifice plate thickness<br />

Compressible Flow Through a Meter<br />

When a compressible fluid (vapor or gas)<br />

flows through a meter, compressibility effects<br />

must be accounted for. This is done by introduction<br />

<strong>of</strong> a compressibility factor which can be<br />

determined analytically for some meters<br />

(venturi). For an orifice meter, on the other hand,<br />

the compressibility factor must be measured.<br />

The equations and formulation developed<br />

thus far were for incompressible flow through a<br />

meter. For compressible flows, the derivation is<br />

somewhat different. When the fluid flows<br />

through a meter and encounters a change in area,<br />

the velocity changes as does the pressure. When<br />

pressure changes, the density <strong>of</strong> the fluid changes<br />

and this effect must be accounted for in order to<br />

obtain accurate results. To account for<br />

compressibility, we will rewrite the descriptive<br />

equations.<br />

Venturi Meter<br />

Consider isentropic, subsonic, steady flow <strong>of</strong><br />

an ideal gas through a venturi meter. The<br />

continuity equation is<br />

ρ 1 A 1 V 1 = ρ 2 A 2 V 2 = ·m isentropic = ·m s<br />

where section 1 is upstream <strong>of</strong> the meter, and<br />

section 2 is at the throat. Neglecting changes in<br />

potential energy (negligible compared to changes<br />

in enthalpy), the energy equation is<br />

h 1 + V 1 2<br />

2 = h 2 + V 2 2<br />

2<br />

The enthalpy change can be found by assuming<br />

that the compressible fluid is ideal:<br />

h 1 - h 2 = C p (T 1 - T 2 )<br />

Combining these equations and rearranging gives<br />

or<br />

C p T 1 +<br />

·<br />

m s<br />

2<br />

2ρ 1 2 A 1<br />

2 = C pT 2 +<br />

m<br />

· 2 s<br />

2ρ 2 2 A<br />

2 2<br />

m<br />

· 1<br />

2 s<br />

⎛<br />

⎞<br />

⎝ ρ 2 2 A<br />

2 - 1<br />

2 ρ 2 1 A<br />

2 1 ⎠<br />

= 2C p(T 1 - T 2 )<br />

= 2C p T 1<br />

⎝ ⎛ 1 - T 2<br />

T ⎠ ⎞<br />

1<br />

If we assume an isentropic compression process<br />

through the meter, then we can write<br />

p 2 T<br />

= ⎛ 2<br />

⎞<br />

p 1 ⎝ T 1 ⎠<br />

γ<br />

γ - 1<br />

where γ is the ratio <strong>of</strong> specific heats (γ = C p /C v ).<br />

Also, recall that for an ideal gas,<br />

C p = R γ<br />

γ - 1<br />

Substituting, rearranging and simplifying, we get<br />

41


m<br />

· 2 s<br />

ρ 2 2 A<br />

2 2 ⎝ ⎛ 1 - ρ 2 2 A<br />

2 2<br />

ρ ⎠ ⎞ 2 1 A<br />

2 = 2 R γ<br />

1 γ - 1 T 1<br />

⎢ ⎡ p<br />

1 - ⎛ 2<br />

⎞<br />

⎣ ⎝ ⎠<br />

γ - 1<br />

γ<br />

p 1<br />

For an ideal gas, we write ρ = p/RT. Substituting<br />

for the RT 1 term in the preceding equation yields<br />

·<br />

m s<br />

2<br />

A 2<br />

2 = 2ρ 2 2<br />

γ<br />

γ - 1 ⎝ ⎛<br />

p 1<br />

ρ 1<br />

(γ - 1)/γ<br />

⎠ ⎞ 1 - (p 2 /p 1 )<br />

1 - (ρ 2 2 A 2 2 /ρ 2 1 A 2 1 )<br />

For an isentropic process, we can also write<br />

or<br />

p 1<br />

ρ 1<br />

γ = p 2<br />

ρ 2<br />

γ<br />

p<br />

ρ 2 = ⎛ 2<br />

⎞<br />

⎝ p 1 ⎠<br />

1/γ<br />

ρ1<br />

from which we obtain<br />

ρ 2 p<br />

2 = ⎛ 2<br />

⎞<br />

⎝ ⎠<br />

p 1<br />

2/γ<br />

ρ1<br />

2<br />

Substituting into the mass flow equation, we get<br />

after considerable manipulation Equation 16.1 <strong>of</strong><br />

Table 16.1, which summarizes the results.<br />

Thus for compressible flow through a venturi<br />

meter, the measurements needed are p 1 , p 2 , T 1 ,<br />

the venturi dimensions, and the fluid properties.<br />

By introducing the venturi discharge coefficient<br />

C v , the actual flow rate through the meter is<br />

determined to be<br />

·<br />

m ac = C v<br />

·ms<br />

Combining this result with Equation 16.1 gives<br />

Equation 16.2 <strong>of</strong> Table 16.1.<br />

It would be convenient if we could re-write<br />

Equation 16.2 in such a way that the<br />

compressibility effects could be consolidated into<br />

one term. We attempt this by using the flow rate<br />

equation for the incompressible case multiplied<br />

by another coefficient called the compressibility<br />

factor Y; we therefore write<br />

m<br />

·<br />

2(p<br />

ac = C v Yρ 1 A 2<br />

√⎺⎺⎺⎺⎺⎺<br />

1 - p 2 )<br />

ρ 1 (1 - D 4 2 /D 4 1 )<br />

We now set the preceding equation equal to<br />

Equation 16.2 and solve for Y. We obtain Equation<br />

16.3 <strong>of</strong> the table.<br />

The ratio <strong>of</strong> specific heats γ will be known for<br />

a given compressible fluid, and so Equation 16.3<br />

⎦ ⎥⎤<br />

could be plotted as compressibility factor Y versus<br />

pressure ratio p 2 /p 1 for various values <strong>of</strong> D 2 /D 1 .<br />

The advantage <strong>of</strong> using this approach is that a<br />

pressure drop term appears just as with the<br />

incompressible case, which is convenient if a<br />

manometer is used to measure pressure. Moreover,<br />

the compressibility effect has been isolated into<br />

one factor Y.<br />

Orifice Meter<br />

The equations and formulation <strong>of</strong> an analysis<br />

for an orifice meter is the same as that for the<br />

venturi meter. The difference is in the evaluation<br />

<strong>of</strong> the compressibility factor. For an orifice meter<br />

the compressibility factor is much lower than<br />

that for a venturi meter. The compressibility<br />

factor for an orifice meter cannot be derived, but<br />

instead must be measured. Results <strong>of</strong> such tests<br />

have yielded the Buckingham equation, Equation<br />

16.4 <strong>of</strong> Table 16.1, which is valid for most<br />

manometer connection systems.<br />

Calibration <strong>of</strong> a Meter<br />

Figures 16.4 and 16.5 show how the apparatus<br />

is set up. An axial flow fan is attached to the<br />

shaft <strong>of</strong> a DC motor. The rotational speed <strong>of</strong> the<br />

motor, and hence the volume flow rate <strong>of</strong> air, is<br />

controllable. The fan moves air through a duct<br />

into which a pitot-static tube is attached. The<br />

pitot static tube is movable so that the velocity<br />

at any radial location can be measured. An orifice<br />

or a venturi meter can be placed in the duct<br />

system.<br />

The pitot static tube has pressure taps which<br />

are to be connected to a manometer. Likewise each<br />

meter also has pressure taps, and these will be<br />

connected to a separate manometer.<br />

A meter for calibration will be assigned by<br />

the instructor. For the experiment, make<br />

measurements <strong>of</strong> velocity using the pitot-static<br />

tube to obtain a velocity pr<strong>of</strong>ile. Draw the<br />

velocity pr<strong>of</strong>ile to scale. Obtain data from the<br />

velocity pr<strong>of</strong>ile and determine a volume flow<br />

rate.<br />

For one velocity pr<strong>of</strong>ile, measure the pressure<br />

drop associated with the meter. Graph volume<br />

flow rate as a function <strong>of</strong> head loss ∆h obtained<br />

from the meter, with ∆h on the horizontal axis.<br />

Determine the value <strong>of</strong> the compressibility factor<br />

experimentally and again using the appropriate<br />

equation (Equation 16.3 or 16.4) for each data<br />

point. A minimum <strong>of</strong> 9 data points should be<br />

obtained. Compare the results <strong>of</strong> both<br />

calculations for Y.<br />

42


TABLE 16.1. Summary <strong>of</strong> equations for compressible flow through a venturi or an orifice meter.<br />

m<br />

·<br />

s = A 2 ⎨ ⎧<br />

⎩<br />

2p 1 ρ 1 (p 2 /p 1 ) 2/γ [γ/(γ - 1)] [1 - (p 2 /p 1 ) (γ - 1)/γ 1/2<br />

]<br />

1 - (p ⎭ ⎬⎫<br />

2 /p 1 ) 2/γ (D 4 2 /D 4 1 )<br />

(16.1)<br />

m<br />

·<br />

ac = C v A 2 ⎨ ⎧<br />

⎩<br />

2p 1 ρ 1 (p 2 /p 1 ) 2/γ [γ/(γ - 1)] [1 - (p 2 /p 1 ) (γ - 1)/γ 1/2<br />

]<br />

1 - (p ⎭ ⎬⎫<br />

2 /p 1 ) 2/γ (D 4 2 /D 4 1 )<br />

(16.2)<br />

Y = √⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺⎺<br />

γ [(p 2 /p 1 ) 2/γ - (p 2 /p 1 ) (γ + 1)/γ ](1 - D 4 2 /D 4 1 )<br />

γ - 1 [1 - (D 4 2 /D 4 1 )(p 2 /p 1 ) 2/γ ](1 - p 2 /p 1 )<br />

Y = 1 - (0.41 + 0.35β 4 ) (1 - p 2/p 1 )<br />

γ<br />

(venturi meter) (16.3)<br />

(orifice meter) (16.4)<br />

rounded<br />

inlet<br />

manometer<br />

connections<br />

motor<br />

pitot-static<br />

tube<br />

axial flow outlet duct<br />

fan<br />

FIGURE 16.4. Experimental setup for calibrating a venturi meter.<br />

venturi meter<br />

manometer<br />

connections<br />

rounded<br />

inlet<br />

motor<br />

axial flow<br />

fan<br />

outlet duct<br />

pitot-static<br />

tube<br />

orifice plate<br />

FIGURE 16.5. Experimental setup for calibrating an orifice meter.<br />

43


EXPERIMENT 17<br />

MEASUREMENT OF FAN HORSEPOWER<br />

The objective <strong>of</strong> this experiment is to measure<br />

performance characteristics <strong>of</strong> an axial flow fan,<br />

and display the results graphically.<br />

Figure 17.1 shows a schematic <strong>of</strong> the<br />

apparatus used in this experiment. A DC motor<br />

rotates an axial flow fan which moves air<br />

through a duct. The sketch shows a venturi meter<br />

used in the outlet duct to measure flow rate.<br />

However, an orifice meter or a pitot-static tube<br />

can be used instead. (See Experiment 16.) The<br />

control volume from section 1 to 2 includes all the<br />

fluid inside. The inlet is labeled as section 1, and<br />

has an area (indicated by the dotted line) so huge<br />

that the velocity at 1 is negligible compared to<br />

the velocity at 2. The pressure at 1 equals<br />

atmospheric pressure. The fan thus accelerates<br />

the flow from a velocity <strong>of</strong> 0 to a velocity we<br />

identify as V 2 . The continuity equation is<br />

m· 1 = m· 2<br />

The energy equation is<br />

0 = - dW<br />

dt + m· ⎡<br />

1 h ⎣<br />

1 + V 1 2 ⎤ - m· ⎡<br />

2 h<br />

2 ⎦ ⎣<br />

2 + V 2 2 ⎤<br />

2 ⎦<br />

where dW/dt is the power input from the fan to<br />

the air, which is what we are solving for. By<br />

substituting the enthalpy terms according to the<br />

definition (h = u + pv), the preceding equation<br />

becomes<br />

dW<br />

dt = m·<br />

(u 1 - u 2 )<br />

+ m·<br />

⎧ p 1<br />

⎨<br />

⎩<br />

p 2<br />

⎡ ⎤<br />

⎣ ρ + V 1 2<br />

- ⎡ ⎤<br />

2 ⎦ ⎣ ρ + V 2 2<br />

2 ⎦<br />

Assuming ideal gas behavior, we have<br />

⎫<br />

⎬<br />

⎭<br />

u 1 - u 2 = C v (T 1 - T 2 )<br />

With a fan, however, we assume an isothermal<br />

process, so that T 1 ≈ T 2 and ρ 1 ≈ ρ 2 = ρ. With m·<br />

=<br />

ρAV (evaluated at the outlet, section 2), the<br />

equation for power becomes<br />

dW<br />

dt<br />

= A 2 V 2 ⎨ ⎧ ⎡p ⎩ ⎣<br />

1 + ρV 1 2 ⎤ - ⎡p 2 ⎦ ⎣<br />

2 + ρV 2 2 ⎤<br />

2 ⎦ ⎭ ⎬⎫<br />

Recall that in this analysis, we set up our control<br />

volume so that the inlet velocity V 1 = 0; actually<br />

V 1


placed on the torque arm to reposition the motor<br />

to its balanced position. The product <strong>of</strong> weight<br />

and torque arm length gives the torque input from<br />

motor to fan.<br />

A tachometer is used to measure the<br />

rotational speed <strong>of</strong> the motor. The product <strong>of</strong><br />

torque and rotational speed gives the power input<br />

to the fan:<br />

dW a<br />

dt<br />

= Tω (17.2)<br />

This is the power delivered to the fan from the<br />

motor.<br />

The efficiency <strong>of</strong> the fan can now be<br />

calculated using Equations 1 and 2:<br />

η = dW/dt<br />

dW a /dt<br />

(17.3)<br />

Thus for one setting <strong>of</strong> the motor controller, the<br />

following readings should be obtained:<br />

1. An appropriate reading for the flow meter.<br />

2. Weight needed to balance the motor, and its<br />

position on the torque arm.<br />

3. Rotational speed <strong>of</strong> the fan and motor.<br />

4. The static pressure at section 2.<br />

With these data, the following parameters<br />

can be calculated, again for each setting <strong>of</strong> the<br />

motor controller:<br />

1. Outlet velocity at section 2: V 2 = Q/A 2 .<br />

2. The power using Equation 17.1.<br />

3. The input power using Equation 17.2.<br />

4. The efficiency using Equation 17.3.<br />

Presentation <strong>of</strong> Results<br />

On the horizontal axis, plot volume flow<br />

rate. On the vertical axis, graph the power using<br />

Equation 1, and Equation 2, both on the same set <strong>of</strong><br />

axes. Also, again on the same set <strong>of</strong> axes, graph<br />

total pressure ∆p t as a function <strong>of</strong> flow rate. On a<br />

separate graph, plot efficiency versus flow rate<br />

(horizontal axis).<br />

45


EXPERIMENT 18<br />

MEASUREMENT OF PUMP PERFORMANCE<br />

The objective <strong>of</strong> this experiment is to perform<br />

a test <strong>of</strong> a centrifugal pump and display the<br />

results in the form <strong>of</strong> what is known as a<br />

performance map.<br />

Figure 18.1 is a schematic <strong>of</strong> the pump and<br />

piping system used in this experiment. The pump<br />

contains an impeller within its housing. The<br />

impeller is attached to the shaft <strong>of</strong> the motor<br />

and the motor is mounted so that it is free to<br />

rotate, within limits. As the motor rotates and<br />

the impeller moves liquid through the pump, the<br />

motor housing tends to rotate in the opposite<br />

direction from that <strong>of</strong> the impeller. A calibrated<br />

measurement system gives a readout <strong>of</strong> the torque<br />

exerted by the motor on the impeller.<br />

The rotational speed <strong>of</strong> the motor is obtained<br />

with a tachometer. The product <strong>of</strong> rotational<br />

speed and torque is the input power to the<br />

impeller from the motor.<br />

Gages in the inlet and outlet lines about the<br />

pump give the corresponding pressures in gage<br />

pressure units. The gages are located at known<br />

heights from a reference plane.<br />

After moving through the system, the water<br />

is discharged into an open channel containing a<br />

V-notch weir. The weir is calibrated to provide<br />

the volume flow rate through the system.<br />

The valve in the outlet line is used to control<br />

the volume flow rate. As far as the pump is<br />

concerned, the resistance <strong>of</strong>fered by the valve<br />

simulates a piping system with a controllable<br />

friction loss. Thus for any valve position, the<br />

following data can be obtained: torque, rotational<br />

speed, inlet pressure, outlet pressure, and volume<br />

flow rate. These parameters are summarized in<br />

Table 18.1.<br />

TABLE 18.1. Pump testing parameters.<br />

Raw Data<br />

Parameter Symbol Dimensions<br />

torque T F·L<br />

rotational speed ω 1/T<br />

inlet pressure p 1 F/L 2<br />

outlet pressure p 2 F/L 2<br />

volume flow rate Q L 3 /T<br />

The parameters used to characterize the<br />

pump are calculated with the raw data obtained<br />

from the test (listed above) and are as follows:<br />

input power to the pump, the total head<br />

difference as outlet minus inlet, the power<br />

imparted to the liquid, and the efficiency. These<br />

parameters are summarized in Table 18.2. These<br />

parameters must be expressed in a consistent set <strong>of</strong><br />

units.<br />

TABLE 18.2. Pump characterization parameters.<br />

Reduced Data<br />

Parameter Symbol Dimensions<br />

input power dW a /dt F·L/T<br />

total head diff ∆H L<br />

power to liquid dW/dt F·L/T<br />

efficiency η —<br />

The raw data are manipulated to obtain the<br />

reduced data which in turn are used to<br />

characterize the performance <strong>of</strong> the pump. The<br />

input power to the pump from the motor is the<br />

product <strong>of</strong> torque and rotational speed:<br />

- dW a<br />

dt<br />

= Tω (18.1)<br />

where the negative sign is added as a matter <strong>of</strong><br />

convention. The total head at section 1, where<br />

the inlet pressure is measured (see Figure 18.1), is<br />

defined as<br />

H 1<br />

= p 1<br />

ρg + V 1 2<br />

2g + z 1<br />

where ρ is the liquid density and V 1<br />

(= Q/A) is<br />

the velocity in the inlet line. Similarly, the<br />

total head at position 2 where the outlet pressure<br />

is measured is<br />

H 2<br />

= p 2<br />

ρg + V 2 2<br />

2g + z 2<br />

The total head difference is given by<br />

46


∆H = H 2<br />

- H 1<br />

= p 2<br />

ρg + V 2 2<br />

2g + z 2<br />

p<br />

- ⎛ 1<br />

⎞<br />

⎝ ρg + V 1 2<br />

2g + z 1⎠<br />

The dimension <strong>of</strong> the head H is L (ft or m). The<br />

power imparted to the liquid is calculated with<br />

the steady flow energy equation applied from<br />

section 1 to 2:<br />

- dW<br />

dt<br />

p 2<br />

= m· g<br />

⎡⎛<br />

⎞<br />

⎣⎝ ρg + V 2 2<br />

2g + z 2<br />

⎠<br />

-<br />

⎛<br />

⎝<br />

p 1<br />

ρg + V 1 2<br />

In terms <strong>of</strong> total head H, we have<br />

⎞⎤<br />

2g + z 1⎠<br />

⎦<br />

- dW<br />

dt = m· g (H 2<br />

- H 1<br />

) = m· g ∆H (18.2)<br />

The efficiency is determined with<br />

η = dW/dt<br />

dW a<br />

/dt<br />

(18.3)<br />

Experimental Method<br />

The experimental technique used in obtaining<br />

data depends on the desired method <strong>of</strong> expressing<br />

performance characteristics. For this experiment,<br />

data are taken on only one impeller-casing-motor<br />

combination. One data point is first taken at a<br />

certain valve setting and at a preselected<br />

rotational speed. The valve setting would then be<br />

changed and the speed control on the motor (not<br />

shown in Figure 18.1) is adjusted if necessary so<br />

that the rotational speed remains constant, and<br />

the next set <strong>of</strong> data are obtained. This procedure<br />

is continued until 6 data points are obtained for<br />

one rotational speed.<br />

Next, the rotational speed is changed and<br />

the procedure is repeated. Four rotational speeds<br />

should be used, and at least 6 data points per<br />

rotational speed should be obtained.<br />

v-notch weir<br />

return<br />

valve<br />

control panel<br />

and gages<br />

pressure<br />

tap<br />

1 nominal<br />

schedule 40<br />

PVC pipe<br />

•<br />

sump tank<br />

inlet<br />

z 2<br />

pressure<br />

tap<br />

•<br />

valve<br />

motor<br />

motor<br />

shaft<br />

pump<br />

z 1<br />

1-1/2 nominal<br />

schedule 40<br />

PVC pipe<br />

FIGURE 18.1. Centrifugal pump testing setup.<br />

47


Performance Map<br />

A performance map is to be drawn to<br />

summarize the performance <strong>of</strong> the pump over its<br />

operating range. The performance map is a graph<br />

if the total head ∆H versus flow rate Q<br />

(horizontal axis). Four lines, corresponding to the<br />

four pre-selected rotational speeds, would be<br />

drawn. Each line has 6 data points, and the<br />

efficiency at each point is calculated. Lines <strong>of</strong><br />

equal efficiency are then drawn, and the resulting<br />

graph is known as a performance map. Figure 18.2<br />

is an example <strong>of</strong> a performance map.<br />

Total head in ft<br />

40<br />

30<br />

20<br />

10<br />

3600 rpm<br />

2700<br />

1760<br />

900<br />

65%<br />

Efficiency in %<br />

75%<br />

80%<br />

85%<br />

75%<br />

0<br />

0 200 400 600 800<br />

Volume flow rate in gallons per minute<br />

65%<br />

FIGURE 18.2. Example <strong>of</strong> a performance map <strong>of</strong><br />

one impeller-casing-motor combination<br />

obtained at four different rotational speeds.<br />

Dimensionless Graphs<br />

To illustrate the importance <strong>of</strong><br />

dimensionless parameters, it is prudent to use the<br />

data obtained in this experiment and produce a<br />

dimensionless graph.<br />

A dimensional analysis can be performed for<br />

pumps to determine which dimensionless groups<br />

are important. With regard to the flow <strong>of</strong> an<br />

incompressible fluid through a pump, we wish to<br />

relate three variables introduced thus far to the<br />

flow parameters. The three variables <strong>of</strong> interest<br />

here are the efficiency η, the energy transfer rate<br />

g∆H, and the power dW/dt. These three<br />

parameters are assumed to be functions <strong>of</strong> fluid<br />

properties density ρ and viscosity µ, volume flow<br />

rate through the machine Q, rotational speed ω,<br />

and a characteristic dimension (usually impeller<br />

diameter) D. We therefore write three functional<br />

dependencies:<br />

η = f 1<br />

(ρ, µ, Q, ω, D )<br />

dW<br />

dt<br />

= f 3<br />

(ρ, µ, Q, ω, D)<br />

Performing a dimensional analysis gives the<br />

following results:<br />

where<br />

ρωD<br />

η = f 1<br />

⎛<br />

2<br />

⎞<br />

⎝ µ , Q<br />

ωD 3 ⎠<br />

g∆H<br />

ω 2 D 2 = f 2<br />

⎝ ⎛ ρωD 2<br />

µ , Q<br />

ωD ⎠ ⎞ 3<br />

dW/dt<br />

ρω 3 D 5 = f 3<br />

⎝ ⎛ ρωD 2<br />

µ , Q<br />

ωD ⎠ ⎞ 3<br />

g∆H<br />

ω 2 D2 = energy transfer coefficient<br />

Q<br />

ωD 3<br />

ρωD 2<br />

µ<br />

dW/dt<br />

ρω 3 D 5<br />

= volumetric flow coefficient<br />

= rotational Reynolds number<br />

= power coefficient<br />

Experiments conducted with pumps show that the<br />

rotational Reynolds number (ρωD 2 /µ) has a<br />

smaller effect on the dependent variables than<br />

does the flow coefficient. So for incompressible<br />

flow through pumps, the preceding equations<br />

reduce to<br />

Q<br />

η ≈ f 1<br />

⎛ ⎞<br />

⎝ ωD 3 (18.4)<br />

⎠<br />

g∆H<br />

ω 2 D 2 ≈ f 2<br />

⎝ ⎛ Q<br />

ωD ⎠ ⎞ 3 (18.5)<br />

dW/dt<br />

ρω 3 D 5<br />

Q<br />

≈ f 3<br />

⎛ ⎞<br />

⎝ ωD 3 (18.6)<br />

⎠<br />

For this experiment, construct a graph <strong>of</strong><br />

efficiency, energy transfer coefficient, and power<br />

coefficient all as functions <strong>of</strong> the volumetric flow<br />

coefficient. Three different graphs can be drawn,<br />

or all lines can be placed on the same set <strong>of</strong> axes.<br />

g∆H = f 2<br />

(ρ, µ, Q, ω, D)<br />

48


Specific Speed<br />

A dimensionless group known as specific<br />

speed can also be derived. Specific speed is found<br />

by combining head coefficient and flow<br />

coefficient in order to eliminate characteristic<br />

length D:<br />

ω ss<br />

= ⎝<br />

⎛<br />

or ω ss<br />

=<br />

Q<br />

⎞<br />

⎠<br />

ωD 3 1/2<br />

ωQ 1/2<br />

(g∆H) 3/4<br />

ω<br />

⎛<br />

2 D 2<br />

⎞<br />

⎝ g∆H ⎠<br />

3/4<br />

[dimensionless]<br />

Exponents other than 1/2 and 3/4 could be used (to<br />

eliminate D), but 1/2 and 3/4 are customarily<br />

selected for modeling pumps. Another definition<br />

for specific speed is given by<br />

ω s<br />

= ωQ1/2<br />

∆H 3/4 ⎣ ⎡ rpm = rpm(gpm)1/2<br />

⎦ ⎤<br />

ft 3/4<br />

in which the rotational speed ω is expressed in<br />

rpm, volume flow rate Q is in gpm, total head ∆H<br />

is in ft <strong>of</strong> liquid, and specific speed ω s<br />

is<br />

arbitrarily assigned the unit <strong>of</strong> rpm. The equation<br />

for specific speed ω ss<br />

is dimensionless whereas<br />

ω s<br />

is not.<br />

The specific speed <strong>of</strong> a pump can be<br />

calculated at any operating point, but<br />

customarily specific speed for a pump is<br />

determined only at its maximum efficiency. For<br />

the pump <strong>of</strong> this experiment, calculate its<br />

specific speed using both equations.<br />

49


Appendix<br />

Calibration Curves<br />

Orifice plates—open channel flow apparatus ....................................51<br />

V-notch weir—turbomachinery experiments...................................... 52<br />

50


1<br />

large orifice<br />

volume flow rate in ft 3 /s<br />

0.1<br />

small orifice<br />

0.01<br />

0.01 0.1 1 10<br />

manometer deflection in ft <strong>of</strong> water<br />

FIGURE A.1. Calibration curve for the open channel flow device.<br />

51


120<br />

100<br />

80<br />

height reading in mm<br />

60<br />

40<br />

20<br />

0<br />

0 50 100 150 200 250 300 350<br />

volume flow rate in liters/min<br />

FIGURE A.2. Calibration curve for the V-notch weir, turbomachinery experiments.<br />

52

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!