You are on page 1of 105

LABORATORY

MANUAL
Phys1101, Spring 2011
University of Minnesota-Morris

Table of Contents

Pre-Lab Assignment for Experiment 1: ..................................................................................................... 5


Laboratory Experiment #1: Measurement ................................................................................................. 7
Pre-Lab Assignment for Experiment 2: Basic rules of Uncertainty propagation ..................................... 15
Laboratory Experiment #2: Motion with Constant Acceleration .............................................................. 17
Pre-Lab Assignment for Experiment 3: ................................................................................................... 21
Laboratory Experiment #3: Motion in two dimensions............................................................................. 23
Pre-Lab Assignment for Experiment 4: ................................................................................................... 27
Laboratory Experiment #4: The Addition of Vectors ................................................................................ 29
Pre-Lab Assignment for Experiment 5: ................................................................................................... 33
Laboratory Experiment #5: Newton's Second Law ................................................................................. 35
Pre-Lab Assignment for Experiment 6: Dealing with Random Errors ..................................................... 39
Laboratory Experiment #6: Centripetal Force ......................................................................................... 41
Pre-Lab Assignment for Experiment 7: ................................................................................................... 45
Laboratory Experiment #7: Collisions in One Dimension ........................................................................ 47
Pre-Lab Assignment for Experiment 8: Logarithm Rule .......................................................................... 51
Laboratory Experiment #8: Ballistic Pendulum ........................................................................................ 53
Pre-Lab Assignment for Experiment 9: ................................................................................................... 57
Laboratory Experiment #9: Static Equilibrium ......................................................................................... 59
Pre-Lab Assignment for Experiment 10: ................................................................................................. 63
Laboratory Experiment #10: Hooke's Law and Simple Harmonic Motion ............................................... 65
Pre-Lab Assignment for Experiment 11: Estimation ................................................................................ 69
Laboratory Experiment # 11: Standing Waves in a String ...................................................................... 71
Pre-Lab Assignment for Experiment 12: ................................................................................................. 77
Laboratory Experiment # 12: Calorimetry - Specific Heat ....................................................................... 79
Appendix A: Forms of Lab Reports ......................................................................................................... 81
Appendix B: Sample Lab Report ............................................................................................................. 82
Appendix C: Significant Figures .............................................................................................................. 86
Appendix D: Uncertainty ......................................................................................................................... 90
Appendix E: Propagation of Uncertainties .............................................................................................. 95
Appendix F: Finding the Slope and Slope of Linear Graphs ............................................................... 103
Appendix G: Excel Tips for Physics Labs .............................................................................................. 104

Name: _________________________________________________

Laboratory Experiment 1: Measurement


Score: _______of 15

Pre-Lab Assignment for Experiment 1:


This assignment needs to be completed before you come to lab. Turn in at the beginning of lab.
A. Study the use of significant figures. You can find information in Appendix C of the Lab manual or in
section 1.5, page 8, of the text book.
B. For the following numbers, determine the number of significant figures, and write them in scientific
notation.(4)
Physical quantity
Number of SFs
Scientific notation
520 m
420. miles
2000 years
C. Compute the net value and write it after the equal sign using the proper number of significant
figures.(4)

6.44

7.0

57.6

(3.74 - 2.83)

59.16

D. Study the role of uncertainty in Appendix D or section 1.5, pp 8ff textbook. Acquire and
understanding of the following: (a) Precision error, (b) accuracy error, (c) random error, (d) absolute
error, (e) relative error.
E. You use your wristwatch (showing time as in 3:25:12 PM) to measure the time it takes to walk from
your door to the door of the physics lab, and find: 12 minutes 20.33 seconds. For this time
measurement, analyze the uncertainty:
Type of Uncertainty
Why does the measurement have this type of
Size of this error
error?
(number and unit, if
applicable)
Precision error
(limited by scale; equal to of
smallest scale unit)

Accuracy error
(limited by measuring method
and calibration; small)

Random error
(non-reproducible variations

Total absolute uncertainty


Total relative uncertainty

Laboratory Experiment 1: Measurement

Laboratory Experiment 1: Measurement

Laboratory Experiment #1: Measurement


Objective:
To learn proper measuring techniques. To see and understand the uncertainty inherent in all physical
measurements. To become familiar with the use of significant figures and simple uncertainty calculations.
Introduction:
Meter Stick:
All physical measurements are of limited precision. The instrument, the technique, and the quantity being
measured all contribute to the uncertainty of a measurement. The meter stick, for example, has a least count
of 1 mm. Then the uncertainty due to scale limitations is 1/2 mm. For an experimenter with a sharp eye,
further discernment to the nearest 1/2 mm might be possible, reducing the uncertainty to 1/4 mm. Precision
beyond this is, however, not very reliable with the common meter stick because of the coarseness of the
black lines that mark off the millimeters. Another example is the Vernier caliper--an instrument designed to
measure lengths more accurately than the meter stick. Its use is explained below. The least count of a
standard Vernier caliper is 1/10 mm which leads to an uncertainty of 1/20 mm. But a more precise Vernier
caliper, such as is used in this lab, can have a least count of 1/50 mm. An even more precise instrument is
the micrometer caliper (also explained below). It has a least count of 1/100 mm. An experienced user might
be able to discern tenths of the least count.
The proper use of the meter stick might seem to need little explanation--one simply holds it against the object
being measured and computes the difference between the readings at the extremes of the object. Yet, even
during such a simple operation, a couple of errors are commonly made. Both involve the manner in which the
meter stick is held.
The parallax (line-of-sight) error can occur when the actual measuring scale is not up against the object being
measured. When such is the case, the reading will depend upon the line-of-sight of the experimenter. (See
Figure 1) The resulting reading will be of questionable value. This error is considered a blunder and should be
avoided.

Reads 7.4

Reads 7.6

Figure 1.1: The parallax plunder


By holding the measuring scale directly against the object, a good reading can be taken irrespective of the
line of sight of the experimenter. (See Figure 2)
A "zero-error" is also common because the end of the meter stick is often assumed to correspond to a
reading of "zero". For real meter sticks, such is usually not the case. The discrepancy might be due to the
manufacturing process or to the wearing-down of the meter stick with use. In Figure 3, a technique that avoids
the zero-error is illustrated. (Notice that a couple of units are worn off the left end of the stick.) The end of the
meter stick is not used. Rather some convenient reference is chosen such as the 10 unit mark. The length is
then the difference between the extremes.
7

Laboratory Experiment 1: Measurement

10

20

30

Reading = 31 units
Figure 1.2: Better positioning of a meter stick

10

20

30

Measured length = 29 - 10 units = 19 units


Figure 1.3: Avoiding the zero error.

Vernier Caliper:
A typical Vernier caliper is pictured in Figure 4. It consists primarily of two parts: the main frame and the
movable frame. Often the frame will include both English and metric scales and the number of lines on the
Vernier can be more than the number shown here.
The outside caliper jaws are used to measure the outside dimensions of an object while the inside caliper
jaws are used to measure inside dimensions (such as the inner diameter of a tube). The depth gauge can
measure the depth of holes.

inside caliper jaws

main frame

movable frame

depth gauge

outside caliper jaws


Figure 1.4: A Vernier caliper
To read the Vernier caliper, one first looks to see where, on the main scale, the zero line of the movable scale
(the "index line") lies. In general it will not line up exactly with any line on the main scale but will exceed some
line by some fraction of the least count. (See Figure 5.) The purpose of the movable or "Vernier" scale is to
8

Laboratory Experiment 1: Measurement


indicate that fraction to the nearest tenth. One counts the lines of the Vernier scale from the left (starting with
zero) until one comes to a line that is best aligned with a main-scale line. The number of that Vernier line
indicates the number of tenths by which the reading exceeds the number just to the left of the index line.
Example

30

40
Main Scale

Vernier Scale
5
0

10
Figure 1.5: A basic Vernier scale.

The index line indicates 26 mm + some fraction. Since the 7th line of the Vernier scale is
aligned with a main-scale line, that fraction is .7 mm and the reading is 26.7 mm. Since the
alignment of the seventh line was clear, we assign an uncertainty of one-half the least count,
namely .05 mm. The measurement value is then 26.7 .05 mm.
Example

60

70
Main Scale

Vernier Scale
5
0

10
Figure 1.6

The index line indicates 60. + a fraction. Since the 4th line of the Vernier scale is aligned, the
fraction is .4 mm. Thus the reading is 60.4 mm. If the alignment was clear, we assign an
uncertainty of one-half the least count, namely .05 mm. Then the recorded value should be
60.4 .05 mm.

Laboratory Experiment 1: Measurement


Example

90

100
Main Scale

Vernier Scale
5
0

10
Figure 1.7
The proper reading here is 91.9 .05 mm.

The Vernier scale works because its lines are only 9/10 as far apart as the main scale lines. Figure 8 shows
the relationship between the two scales when the zeroes are aligned.

0 1/10

3/10

5/10

7/10

9/10

Main Scale

Vernier Scale

5
0

2/10

4/10

10

6/10

8/10

10

Figure 1.8: The fundamental principle of the Vernier.


Observe that when the Vernier's zero line is aligned, all the other Vernier lines are not aligned. (The "ten" line
is not counted.) Now, if the Vernier scale was moved just 1/10 of a unit to the right, then line number "one" of
the Vernier scale would be aligned. If the scale was instead moved 6/10, then line "six" would be aligned. By
studying Figure 8, we can understand how the Vernier scale indicates fractional readings.
A systematic zero-error can also occur while using the Vernier caliper. When the calipers are closed all the
way, the reading ought to be zero. If it is not, it is simply necessary to adjust all readings accordingly. For
example, if a Vernier caliper reads .3 mm when closed, then all readings will be .3 mm too large. Just subtract
.3 mm from each reading.
The stainless steel (Scherr-Tumico) Vernier calipers that will be available in lab have a Vernier scale with fifty
lines rather than the standard ten lines discussed above. In this case, by determining which of the lines on the
Vernier is best aligned with a line on the frame scale, one can measure with a least count of .02 mm. Further
estimating is quite difficult. In fact, it will sometimes be difficult to decide which single line is best aligned. The
following example illustrates this difficulty though artificially.
Example:
Notice that the Vernier is now the upper scale and the units are cm.

10

Laboratory Experiment 1: Measurement


Vernier scale
0

main scale

10

Figure 1.9: An example with a fifty-line Vernier scale. The digital


nature of this computer rendition causes the erroneous appearance
of five lines that all seem to be perfectly aligned.
The zero line of the Vernier indicates that the value of the measurement is 2.1 cm plus some additional
fraction. The additional fraction is determined by looking at which line of the Vernier is best aligned with a line
of the main scale. Since five lines all look nicely aligned (58, 60, 62, 64, and 66), we will take the reading that
lies in the center of the five, namely 62. Then the reading is 2.162 cm. A reasonable estimate of uncertainly
might be .002 since it was not very clear which line was best aligned. Written value: 2.162 .002 cm.
Micrometer Caliper:
A typical micrometer caliper is shown in Figure 10.

Figure 1.10: The micrometer caliper.


The movable jaw advances or recedes when the thimble is rotated due to screw threads inside the sleeve.
(Important: when closing the jaws, do not turn the thimble directly but rather turn the slip knob. This will
prevent damage to the instrument and guarantee a consistent exertion of pressure on the jaws.) Because one
revolution of the screw advances the jaw exactly .5 mm, the fifty equally spaced markings around the
circumference of the thimble each correspond to 1/50th of .5 mm or .01 mm. Furthermore, as the thimble
turns it reveals more of the sleeve scale. Between these two scales, very accurate readings are possible. An
experimenter with a sharp eye can even make estimates of readings that lie between lines on the thimble
scale.

11

Laboratory Experiment 1: Measurement


Example

sleeve
0

thimble
35
30
25

Figure 1.11
From the location of the thimble edge on the sleeve scale, one reads 7.5 mm + a fraction. From the
location of the index line (the horizontal line on the sleeve) on the thimble scale one reads 30
hundredths. Thus: reading = 7.80 mm. Actually, because the index line was so well lined up with
the 30, one might dare to say the reading is 7.800 .0005 mm.
Example

thimble

sleeve
15

20

25
20

Figure 1.12
From the sleeve scale, the reading is 23 mm + fraction. From the thimble scale, the fraction is .23
plus some little bit. Estimating that little bit to be 8/10 of the distance between "23" and "24" on the
thimble scale, one gets a thimble reading of .238. Then the total reading is 23.238 mm. Since the
last digit, the 8, was an estimate, its uncertainty might 2. Then the reading should be recorded as
23.238 .002 mm.
Example

thimble

sleeve
5

10

35
30

Figure 1.13
From the sleeve one reads 13.5 mm plus a fraction. From the thimble: fraction = .322 .002. Then
the reading is 13.822 .002 mm.
The possibility of a zero-error also occurs with the micrometer caliper. To check the calibration, simply close
the jaws completely, using the slip knob, and take the reading. If the reading is non-zero, all readings with that
particular instrument will have to be adjusted accordingly.
Equipment:
Meter stick, Scherr-Tumico Vernier caliper, micrometer caliper, 150g (m=.01g) digital mass balance,
sphere, small cylinder, block (iron).

12

Laboratory Experiment 1: Measurement


Procedure:
Measure the dimensions of the sphere using the meter stick, the Vernier caliper and the micrometer caliper.
When measuring, watch for uncertainty inherent in the quantity being measured as well as for uncertainty of
the measuring devices. (Irregularities in the dimensions of the object being measured may exceed the
uncertainty associated with the measuring device. Check for this by measuring the dimension at different
locations and noting variations of readings.) Make sure the measurements made with the three devices agree
with each other (within your assigned uncertainties).
Also measure the dimensions of the cylinder and the block using the three measuring devices, keeping in
mind the tips above.
Measure the mass of each object. Compute the volume and density for each, complete with uncertainty.
(Express the volume and density uncertainties as percentages.)
Uncertainty of volumes:

V
V

W
W

L
L

H
H

V
V

D
D

V
V

D
D

H
H

d
d

m
m

V
V

Comments:
Density, d, is defined as mass, M, divided by volume, V.

m
V

Express your densities in units of g/cm 3.


For your information, the textbook values of a few densities are:
aluminum--2.70 .01 g/cm3
steel--7.9 .05 g/cm3
brass--8.4 .05 g/cm3
copper--8.93 .01 g/cm3
gold--19.28 .05 g/cm3.
Question:
See Spreadsheet.
1. Make three statements about the composition of the three objects, using the micrometer caliper
measurement. State whether your result agrees with the standard densities within uncertainty, and give
possible sources for the discrepancy.
2.

Give reasons for your choice for the value of uncertainty for the diameter of the cylinder as measured
with a Vernier caliper.

Submission:
Completed spreadsheet by e-mail

13

Experiment 2: Motion at constant acceleration

14

Experiment 2: Motion at constant acceleration


Name: ________________________________________________

score:______of 10

Pre-Lab Assignment for Experiment 2: Basic rules of Uncertainty propagation


A. Study the rules for uncertainty propagation in Appendix E of the Lab manual. Note, in particular,

A B C 2D then Y
A B C 2 D
A B
Y
A
B
C
D
b. The multiplication/division rule: Y
then
C D
Y
A
B
C
D
Y
A
A
A
A
n
c. The power rule: Y A
A A
A then
n
Y
A
A
A
A
a. The addition/subtraction rule:

B. In Lab Experiment #2 the acceleration due to gravity g is calculated using


2DL
t 2h
In which D, L and h are length measurements, and t is a time measurement. One measurement gave the
following values:
L 50.0 0.1cm , h 2.7 0.1cm , D 90.6 0.1cm , and t 1.864 0.047s .
g

a) Calculate the value for g from this measurement.

b) Find the % error for each (L, h, D, and t).

L/ L

D/ D

h/ h

t /t

c) Using the Multiplication/Division and Powers Rule, write an equation for the relative uncertainty of g,
containing four terms (one for each of the measured variables).

d) What is the result for g g from this measurement?

15

Experiment 2: Motion at constant acceleration

16

Experiment 2: Motion at constant acceleration

Laboratory Experiment #2: Motion with Constant Acceleration


Objective:
To determine, experimentally, the local acceleration due to gravity. To observe and analyze the motion
of an object having constant acceleration. Familiarize with the tools of graphing by hand.
Part A: objects in free fall
Part B: object on frictionless inclined track
Introduction:
All objects in the vicinity of the earth's surface are drawn toward the center of the earth. In the absence
of any restraints, such objects will fall with velocities that increase uniformly with time. The rate of
change of these velocities (their acceleration) is the same for all objects, namely about 9.8 m/s 2.
Free Fall: When an object is falling with this constant acceleration, its velocity will be uniformly
increasing in the downward (negative y) direction. Unfortunately this all occurs so rapidly or over such
short distances that one cannot visually discern the rate of change of the velocity. One might notice
that the velocity starts out small and increases but without making some measurements one cannot
say how much it was at a given time. For a quantitative analysis, some permanent record of the motion
is needed such as might be provided by a strobe-illuminated time-exposure photograph or by a sparktimer output tape. Such records would show the location of the falling object at equally-spaced
intervals of time.
From this kind of data the average velocity during each time interval can be computed:
vavg = y / t
t. (Since the motion is one-dimensional,
vector notation can be omitted.)
It turns out that one can also determine the instantaneous velocity at certain times from the above type
of data provided the acceleration is constant. In fact, the average velocity calculated above for a given
time interval is equal to the instantaneous velocity at a time halfway through that time interval. To
prove this, consider an interval t to t + t. Then
vavg = [v(t) + v(t + t)] / 2.
But
v(t + t) = v(t) + a t
so that
vavg = [v(t) + v(t) + a t]/2
= v(t) + a t / 2
= v(t + t / 2)
Thus the average velocity during the interval t to t + t is equal to the instantaneous velocity at the midtime t + t /2.
Once the instantaneous velocity is known at a number of different times, a graph can be made
showing v(t). The data points should form a straight line whose slope equals the acceleration of
gravity, g. Thus g can be experimentally determined.
17

Experiment 2: Motion at constant acceleration


Inclined Track: A second method for measuring g is to time the fall of an object through a measured
2
distance, y. From y = gt /2, one could then obtain the acceleration. One difficulty with this method is
that falls over human-scale distances occur in short time intervals that are challenging to measure
accurately. To address this difficulty, it is possible to make the "fall" last longer. This can be
accomplished by letting an object slide down an incline. Then, in the absence of friction, the
component of g parallel to the incline will cause acceleration.

g sin

Figure 2.1: An inclined plane.


time of "fall" down a certain length, D, of frictionless incline, the acceleration can be obtained from
D = a t2/2. Substituting for the acceleration, we obtain D = (g h / L) t2/2. Solving for g yields
g

2DL
t 2h

(1)

Equipment:
One free-fall apparatus with spark timer and D.C. power supply, transparent plastic ruler, meter stick,
air track with air supply, glider with flag, electronic timer with two photogates (t=1%), and vernier
caliper; graph paper.
Procedure:
Part A:
The free-fall apparatus is designed to make a permanent record of a falling object's motion. The object
falls between two vertical wires. A spark generator sends high voltage jolts through the falling object
(via the two wires) at equally spaced time intervals. The resulting sparks make spots on a waxed strip
of paper so that the location of the object is recorded. An analysis of those spots will yield g.
Step 1:

Obtain data by setting the spark timer to 30 sparks per second (t = 1/30 second .1%) and release
the object. Remove the paper tape for analysis. (An alternative is to set the spark timer to 60 sparks
per second and use every other spot.)

Step 2:

Consider ten consecutive, clear spots on the tape. Number them from top to bottom. Measure the
distances between the spots, y, and record these in the suggested tabular form. Measure as precisely
as possible using the transparent rulers! Be sure to record uncertainties.
18

Experiment 2: Motion at constant acceleration


Step 3:

Compute the average velocity during each time interval vavg = y / t.


Step 4: Assuming t=0 corresponds to spot #1, compute the times, tM that correspond to the midpoint of
each interval. (tM is the time halfway between the times corresponding to the endpoints of an interval.
See the table below that illustrates this.)
Spot
t (s)
Interv
tM (s)
y
al
(cm)
1
00
1-2
0.5 (t t)
2
(t t)
2 -3
1.5 (t t)
3
2 (t t)
3-4
2.5 (t t)
4
3 (t t)
4-5
3.5 (t t)
5
4 (t t)
5-6
4.5 (t t)
6
5 (t t)
6-7
5.5 (t t)
7
6 (t t)
7-8
6.5 (t t)
8
7 (t t)
8-9
9
9 - 10
10

Step 5:

Manually, graph vavg as a function of tM. Use a ruler and a sharp pencil (no pen). Include error bars,
using the uncertainties in t and y for each point. Draw the "best" straight line through the points to get
g. Choose two points (NOT any of your data points), far apart, on that line. Read their positions

v1

v1 , t1

t1 and v2

v2 , t2

t2 on the axes of your graph and label the points. Estimate

the uncertainties for v and t in each point by holding your ruler in the positions of the steepest and
flattest line you may fit to your data points (within uncertainty) and observe by how much your values
of

v and t would deviate. The slope

slope a

v2 v1
is the acceleration component. Record it on the
t2 t1

graph. Find the he uncertainty of the acceleration (using add/subtraction rule and multiplication/division
rule) by

slope
slope

a
a

v1
v2
v2 v1

t1
t2
, see
t2 t1

discussion in appendix F.

Part B:
The air track is a device on which a glider can move with almost no friction. If the air track is tilted
slightly from the horizontal, then a glider released from rest at the top of the incline will accelerate
down with acceleration a = g sin.
Step 1:

Carefully level the air track and adjust the air supply so that the glider can move freely. With too little
airflow, the glider may drag on the air track. With too much airflow, the glider may wobble and/or
experience friction from individual air jets. Never slide a glider along the track with the air supply off -you may scratch the track and/or the glider! If the air track is slightly bowed, level it as best you can,
then document where it has peaks and valleys. Try to perform this experiment to minimize the affect of
any bumps.
19

Experiment 2: Motion at constant acceleration


Step 2:

Raise one end of the air track by placing an object of known height, h, under one of the feet of the air
track. Since the feet are not at the ends of the air track, notice that in your analysis L is not the total
length of the air track!

Step 3:

Place the two photogates above the air track so that the glider can pass completely through both of
them. Set the photogate timer resolution to 1 ms. (t inherent is 1%.) Allow enough space below the
second photogate to catch the glider with your hands before it smashes into the end stop, if necessary.
Use the light emitting diodes (LEDs) on the sides of the photogates and the scale on the air track itself
to determine how far the glider travels between tripping the two photogates. You will have to decide
which of the three timer modes to use. Remember not to move either photogate once you have made
this measurement.

Step 4:

Practice releasing the glider smoothly from rest as close as possible to the upper photogate. Use the
LED on the photogate and the scale on the air track to help you do this reproducibly. To characterize
the variability in this measurement, time the glider for 6 good runs.
Remember, Equation 1 assumes that the glider starts from rest at t = 0. With this timing setup, this
cannot be precisely true, but it will be a good approximation if you do the experiment carefully.

Step 5:

Compute a value of g g from the average of the measured times.

Step 6:

Repeat the experiment for a different inclination angle. Then do it again for a third time. Remeasure L
and D each time.
Comment
The accepted value for the acceleration of gravity in Morris, MN is 980.57 0.05 cm/s2.
Question:
In Part 1, if the graph of vavg as a function of tM is extrapolated back to vavg = 0, what is the
corresponding value of t? What is the physical meaning of this time?
Submission:

1.
2.

One hand-drawn graph per student


Electronic lab spreadsheet

20

Experiment 3: Motion in two dimensions


Name: ________________________________________________

Score: ___________of 10

Pre-Lab Assignment for Experiment 3:


For a ball thrown from the right across the room, the following motion diagram was observed:
y

x
A. Sketch the x-t graph and y-t graph for this motion.

B. Sketch the vx-t and the vy-t graphs. Place axes as needed.

vx

vy

C. Sketch the ax-t and the ay-t graphs for this motion. Place axes as needed.

ax

ay

D. Using the Addition/Subtraction rule, and the Multiplication/Division rule, find the equation for the relative
uncertainty for

xn

xn 1
, in which xn 1 , xn 1 and t are measured quantities.
2 t

21

Experiment 3: Motion in two dimensions

22

Experiment 3: Motion in two dimensions

Laboratory Experiment #3: Motion in two dimensions


Objective:
To study position, velocity and acceleration for situations in which motion occurs in two dimensions
simultaneously.
Introduction:
When the motion of objects is confined to a plane, we refer to it as two-dimensional (2-D) motion. Itis instructive
to examine the horizontal and the vertical components of such motion independently. When this is done, many
familiar examples of 2-D motion are revealed to consist of either constant velocity or constant acceleration
motion horizontally, coupled with either constant velocity or constant acceleration motion vertically.
Motion in two dimensions may be studied by analysis of video clips in which the motion occurs in a plane
perpendicular to the line of sight of the camera. A characteristic feature on the moving object is marked in each
frame of the video so that the X-position and Y-position of that feature may be tracked as time goes by. By
knowing the time separation of the frames in the video and the scale of the motion across the screen, one may
deduce position, velocity, and acceleration versus time (and more!). The time information and distance scaling
have already been documented in the video clips that we will examine.
Choose two of the following examples:
A. A toy wind-up car rolling across the floor
B. A ball thrown between two students.
C. The bob of a pendulum is moving back and forth.
The movies will be analyzed frame by frame, positions in x and y will be measured.
Equipment:
Large graph paper, toy car, ball, tape, pendulum, a digital camera, a tripod, meterstick, connector cable to
computer, jump drive to transfer the movies to all work stations

Procedure:
Choose two of the options. The procedure is exactly the same for each of the two types of motion. Follow these
steps:
1.
Cover the reference background with large-scale graph paper. That would be the floor for the car, or the
wall for projectile motion and glider. Make a decision as to x, y directions and mark every tenth box in each
direction. Record the size of 20 units in x and y, as well as their uncertainties.
2.
Bring the camera into a position that allows to view a large enough area to follow your object in its motion.
Use the movie setting at 30 frames per second. Set the maximum frame size possible for better esolution.
3.
Film the motion of your object. Make sure that the object moves through a large portion of the view. For
the projectile, try to throw the ball close and parallel to the marked wall, so that the influence of perspective in
the movie will be minimized, when you determine the position of the ball later-on.
4.

Transfer the movies to your workstation.

23

Experiment 3: Motion in two dimensions


5.
Open them in a program called ImageBrowser. The icon is located on the left-hand panel on your screen.
Double-click on the selected movie and open it in full screen. The play button () will run the movie, which can
be stopped with (). On the right, you find the buttons to move by single frames (| and | ). You can leave
the full-frame view with a double-click on the right side of the screen.
6.
Choose a point on your object, which is visible throughout the movie. Step through the movie frame by
frame (as long as your object is visible), and record the x- and y- position in the data table provided. Also,
assign an uncertainty to your x and y values. Since the movie is using the screen, you have to record on paper
first.
7.

Transfer your data to the Lab3 spreadsheet.

Evaluation:
8.
Plot the trajectory of the object in a x-y graph in a separate sheet in your spreadsheet. Label the sheet,
the axes and the graph. Include error bars in both directions. Help with graphing in Excel can be found in
Appendix F.
9.
Plot each, the x-t graph and the y-t graph in the spreadsheet. Include error bars in both directions, and
label the graphs properly. Ask for assistance if you are not sure how to do this.
10.

For each, the x- and y-direction, find the velocity versus time. This can be best achieved in the following

way: at time tn, the instantaneous velocity of the object can be approximated as

vn

xn

xn 1
. Use Excel
2 t

features wisely to do this if not sure ask for help. Also, determine the uncertainty for each velocity value. Note
that you will not have a velocity value for your first and your last data point.
11. Analyze the acceleration for the y component in the projectile motion In a hand-drawn graph, plot the y
velocity versus time. Use a ruler and a sharp pencil (no pen) for everything. Label axes and graph properly.
Include error bars based on your data uncertainty on each data point..
12.

Draw a best-fit line. Choose two points (NOT any of your data points), far apart, on that line. Read their

positions v1

v1 , t1

t1 and v2

v2 , t2

t2 on the axes of your graph and label the points. Estimate

the uncertainties for v and t in each point by holding your ruler in the positions of the steepest and flattest line
you may fit to your data points (within uncertainty) and observe by how much your values of v and t would
deviate.

slope a

v2 v1
is the y component of the acceleration. Record it on the graph.
t2 t1

13.

The slope

14.

Find the he uncertainty of the acceleration (using add/subtraction rule and multiplication/division rule) by

slope
slope

a
a

v1
v2
v2 v1

t1
t2
, see discussion in appendix F.
t2 t1

15. Now, that you have the acceleration components, go back into the spread sheet, and draw a vector with
the direction of the acceleration into the x-y graph for the data set you chose into three different data points.
Submission:
1. Electronic spreadsheet for each group.
2. One hand-drawn graph for each student.

24

Experiment 3: Motion in two dimensions

Motion A:___________________________________________
Grid dimension (20 units

Uncertainty (20 units)

X direction (cm)
Y direction (cm)
Frame rate (fps)

30

Frame-by-frame positions (beginning when motion is visible):


Frame
X position in grid units
x
1

Y position in grid units

2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25

25

Experiment 3: Motion in two dimensions


Motion B:_____________________________________________________
Grid dimension (20 units

Uncertainty (20 units)

X direction (cm)
Y direction (cm)
Frame rate (fps)

30

Frame-by-frame positions (beginning when motion is visible):


Frame
X position in grid units
x
1

Y position in grid units

2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25

26

Experiment 4: Addition of vectors


Name: ________________________________________________

Score: ___________of 10

Pre-Lab Assignment for Experiment 4: basic uncertainty approximation


A. Study Appendix E again. You will need the basic uncertainty approximation equation for the absolute
error

f
a
a

f
t

f
y

y
f
treats all variables except
a

if f is a function of the measured variables a, t and y. A partial derivative

the a (in this case) as constants, but is otherwise executed as any normal derivative. For the work in B
through E, use the space below and the backside of the paper.
B. For the equation Fx

F cos

Fx

show that

cos

F
o

F sin

where is expressed

F cos

where is expressed

in radians. Find Fx Fx when F = 2.92 .05 N and = 10 .5 .


C. For the equation

Fy

F sin , show that

Fx

sin

F
o

Fx

Fy

Fy

Fy

Fx

in radians. Find Fy Fy when F = 2.92 .05 N and = 10 .5 .


D. For the equation

Fx2

Fy2 , show that

Fx

. Find F F when Fx = 1.02

.02 N and Fy = 1.95 .03 N.


E. For the equation

arctan

Fy
Fx

, show that

Fx

Fy
F

(in radians). Find when

Fx = 1.02 .02 N and Fy = 1.95 .03 N. Remember that will come out in radians. Be sure to report
1
both and in the same units.

The following partial derivatives may be useful:

arctan

a
b

b
a

and

arctan

a
b

a
a b2
2

27

Experiment 4: Addition of vectors

28

Experiment 4: Addition of vectors

Laboratory Experiment #4: The Addition of Vectors


Objective:
To observe and investigate the manner in which vector quantities add.
Introduction:
In order to clearly describe many physical quantities, both a magnitude and a direction must be specified. Such
quantities are called vector quantities. Some examples are position, velocity, acceleration, and force.
The addition (or subtraction) of vector quantities is more complicated than the addition of ordinary numbers
(scalars) because of the effect of the directions. For example, a walk of 4 miles to the East followed by a walk of
3 miles to the North results in a net displacement of only 5 miles:

5 miles
3 miles

4 miles
Figure 4.1
Two different methods exist for computing the sum of vectors--the graphical method and the analytical
(component) method. In the graphical method, the vectors to be added are drawn, with some scale factor, on
graph paper. The first vector is drawn outward form the origin at the angle of the vector with a length
corresponding to its magnitude. The second vector begins where the first ended and continues in its direction
for a distance corresponding to its magnitude. Any further vectors are added consecutively in like fashion. (The
orientation of the vectors is very important.) The sum or "resultant" vector is the vector from the origin to the end
of the last vector added. Figure 4.2 shows the resultant, R, of three vectors A, B, C.

C
R

B
Origin

A
Figure 4.2

The "equilibrant" vector, E, is that vector which, when added to a sum of other vectors, yields a total of zero.
For example:

E + (A + B + C) = E + R = 0
Then
E = - R.
So the equilibrant is simply the opposite of the resultant.

In the analytical method, the vectors to be added are first resolved into components along some set of axes
(e.g. x and y). All the x-components are then added to yield the x-component of the resultant and the ycomponents are added to give the y-component of the resultant.
29

Experiment 4: Addition of vectors


Equipment:
Force table, each with four mounted pulleys and weight hangers, 600g or 1200g digital balances (m=.1g), and
assorted masses.
Procedure:
The force table is a small, round, horizontal table around whose circumference all 360 are indicated. A small
ring at the center of the table is tied to several weight hangers on which various masses can be placed. The
force of gravity on those masses therefore causes forces on the central ring via the strings. (The vertical force
of gravity is turned into a horizontal force by the (frictionless?) pulleys.) See Fig. 4.3.

Figure 4.3: The force table.


2

The force of gravity is F = mg where m is the mass expressed in kilograms (kg) and g = 9.8057 .0005 m/s .
Then the unit of force is the Newton (N). For example, a mass of 100. grams will experience a gravitational
2
force (.100 kg)(9.8057 m/s ) = .981 N.
When all of the forces on the central ring are balanced (i.e. their sum is zero), then the ring will not be pulled
preferentially in any particular direction. Then any one of the forces acting on the ring is the equilibrant of all the
others.
Trial #1
Hang 300.g at 10 and 240.g at 120. (Be sure to include the mass of the weight hanger in the 300.g total.
Some of the weight hangers are 50g each and some are 8g each.) Find the experimental equilibrant on the
force table by hanging masses on a third string. Vary the amount of mass as well as the angle until balance is
achieved. To characterize the variability of this measurement, determine the mass range and angle range over
which balance is maintained. Dont rely on the values stamped into the masses--put them on the digital balance
to get the actual mass value and use that value in the experimental, analytical, and graphical methods. The
balancing vector is the equilibrant. Record the experimental equilibrant in your results table.
Also determine the equilibrant graphically by making a scale drawing on graph paper. Put only one drawing per
page and make the drawing large so as to fill the page. Indicate the scale on each drawing (for example: 1N = 5
cm.)
Also determine the equilibrant analytically. Use the experimental uncertainties of the specified masses and
angles to calculate the uncertainty of the analytical resultant.
Trial #2
Hang 100.g at 10, 200.g at 300, and 250.g at 200. Find the equilibrant in the three usual ways.
30

Experiment 4: Addition of vectors


Properly determined uncertainties will be expected on the experimental and analytical results but you are
excused from any error analysis of the graphical additions. It's not that it's excessively difficult, but it is quite
tedious and time-consuming.
Comments
Please do not overtighten the pulley mounts!
For best results, keep the strings aligned with the radial direction:

Right

Wrong

Figure 4.4: Proper string alignment on the force table ring.


Some new Excel functions you will need are:
ABS(...) = absolute value of argument in parentheses
ATAN(...)=arctangent of argument in parentheses
PI()=3.14159. . . (dont put anything inside the parentheses)
Submission:
1. completed spreadsheet by e-mail
2. Two hand-drawn graphical solutions from each student

31

Experiment 5: Newtons Second Law

32

Experiment 5: Newtons Second Law


Name: ____________________________________

Score: __________of 10

Pre-Lab Assignment for Experiment 5:


A. Refer to appendix E again: The logarithm rule allows to calculate the relative error of a function as

f
f

follows:

ln f
x

ln f
y

ln f
z

z . You will need this rule below.

B. For an arrangement as used in Experiment 5, the theoretical acceleration of the glider is found as

m
M

g . Note that the mass m appears twice in this equation, prohibiting the use of the

multiplication/division rule. Use the logarithm rule to show that the corresponding error equation is

A
A

g
g

M m
m M m

M
.
M m

C. If m = 0.065 .0005 kg, M = 0.1245 .0005 kg, and g = 980.57 .0005 cm/s find acceleration and
uncertainty A A/A (as a percent).

D. Write the equation for the relative uncertainty of acceleration


quantities.

33

2D
, if both, D and t are measured
t2

Experiment 5: Newtons Second Law

34

Experiment 5: Newtons Second Law

Laboratory Experiment #5: Newton's Second Law


Objective:
To explore how the one-dimensional acceleration of an object depends upon the net force applied to the object
and its mass.
Introduction:
The net force on a glider moving on a leveled air track is essentially zero. The airflow provides the normal force
to counter the weight of the glider, so there is no net vertical force. The glider may move along the track with
very little friction because it is not in direct contact with the track, so there is no net horizontal force either.
According to Newton's 1st Law, a glider at rest on a level air track should remain at rest, and a glider moving
along a level air track should move with a constant velocity (until it hits an end stop). In fact, these expectations
may be used to help level an air track, since a glider will accelerate down any slope. If an air track is not
perfectly straight, the slight acceleration of the glider at different points along the track can help identify even
miniscule peaks and valleys. For the subsequent discussion, it is assumed that the air track is perfectly level.
A net force may be applied to a glider on an air track by running a string from one end of the glider, over a
pulley at one end of the air track, to a block which drops vertically as shown in Figure 5.1. The magnitude of the
acceleration of the glider will match that of the dropping block, as long as the string connecting them does not
stretch. Assuming the mass of the string may be neglected and the pulley is nearly ideal (i.e. frictionless and
massless), this one-dimensional acceleration should be a constant.

a
string

pulley

glider

air track
a
m

block
Figure 5.1

The acceleration of the glider and the dropping block may be determined using two photogate timers as follows.
Place the photogates as far apart as possible so that the same feature of the glider, such as the leading edge of
the flag, trips first one gate and then the other. Set the timer mode to record the time between these two
events. For one-dimensional constant-acceleration motion, the experimental acceleration (call it a) is related to
the elapsed time t and the distance traveled d the glider travels by
2

D = (1/2) a t +v0
If the glider starts from rest v0 = 0 so this simplifies to
D = (1/2) a t

(1)

This may be rearranged to deduce the acceleration a from measurements of D and t. Remember that Equation
1 only applies when the glider is timed starting from rest!
This may be accomplished experimentally by placing the glider so that it does not quite block the first photogate
(the indicator light on the side of the photogate will be helpful here) then smoothly releasing the glider so that it

35

Experiment 5: Newtons Second Law


does not wobble as it picks up speed. In this way, the glider will be essentially at rest when the timing begins so
Equation 1 may be used.
The vertical forces on the glider remain balanced, so the net force on the glider is simply the horizontal tension
applied by the string. Notice that the tension in the string is not the same as the weight of the dropping block! If
that were the case, the net force on the dropping block would be zero and it should move at a constant velocity,
which contradicts the observation that the dropping block in fact accelerates downward.
Predict the acceleration of the glider (call this theoretical value A), for comparison with the experimental
acceleration a, as follows. Draw a free-body diagram of the dropping block then apply Newton's 2nd Law
F = ma
to the block to determine how the weight of the block w = mg (directed downward) and the tension in the string
T (directed upward) are related to the acceleration of the block A (directed downward). Considering the
direction of A, which must be larger, T or w? Now rearrange your expression to deduce T in terms of m and A.
Next, draw a free-body diagram of the glider and apply Newton's 2nd Law separately in the horizontal and the
vertical directions. Be sure to use a different variable for the glider mass (perhaps M) so as not to confuse it
with the mass of the block m. Even though the glider accelerates sideways and the block accelerates
downwards, remember that the magnitude of their accelerations A is the same because an inextensible string
connects them. Thus you may use the same variable A for both the block and the glider.
By combining the force equations for the block and the glider, eliminate the variable T and solve for A. This will
allow you to predict the acceleration A simply from the masses m and M (and constants like g). Finally,
compare the predicted acceleration A to the experimental acceleration a, determined earlier from
measurements of D and t. These accelerations should match (within their uncertainties, of course) if Newton's
2nd Law is correct and if the assumptions made about the motion are reasonable. Notice that the experimental
acceleration a will have an associated uncertainty which stems from the uncertainties in D and t, and the
predicted acceleration A will have an associated uncertainty which stems from the uncertainties in m and M.
Equipment:
Air track, glider with flag, electronic timer with two photogates (t=1%), light string or monofilament fishing line,
pulley, collection of masses for dropping block, and a 1200g (m=.1g) digital balance.
Procedure:
Step 1:

Derive an expression for the acceleration A and its uncertainty A and show it to your instructor
before going on to Step 2. Include the derivation (with proper force diagrams) in your report.

Step 2:

Carefully level the air track and adjust the air supply so that the glider can move freely. With too little
airflow, the glider may drag on the air track. With too much airflow, the glider may wobble and/or
experience friction from individual air jets. Never slide a glider along the track with air supply off -you may scratch the track and/or the glider! If the air track is slightly bowed, level it as best you can,
then document where it has peaks and valleys. Try to perform this experiment to minimize the affect
of any curvature.

Step 3:

Attach the string between the glider and the dropping mass, as shown in Figure 5.1. Place the
photogates at the two ends of the accelerated motion (see Comments below). Set the photogate to 1
ms resolution. Deduce the range of masses over which you can safely perform the experiment, which
means catching the glider after it passes the second photogate but before it smashes into the end
stop. (Let x1 be the position of the first gate and x2 the position of the second.) Practice releasing the
glider smoothly from rest as close as possible to the first photo gate. Use the LED on the side of the
photo gate and the scale on the side of the air track to help you do this reproducibly. These two

36

Experiment 5: Newtons Second Law


devices will also allow you to determine the travel distance, D, accurately. Remember not to move
the photo gates once you have measured this distance! Don't forget to determine the mass of the
glider M as well.
Step 4:

Select five different mass values m for the dropping block, starting at 20g and adding 10g increments.
For each value of m, measure t 5 times to help characterize the variability in the timing. Use the
average t in your calculation to determine the experimental acceleration, a. (For the uncertainty on
the average time, take the average of the individual time uncertainties plus the standard deviation of
the time values.)

Step 5:

For each value of m, determine the experimental acceleration a +


acceleration A + A. Present your 5 pairs of results in tabular form.

a as well as the predicted

Comments:
Be sure that the glider is still accelerating when it reaches the second photo gate. In other words, the glider
must trip the second photo gate before the dropping block strikes the floor (at which point the string goes slack
and the glider proceeds with constant velocity).
You may use the accepted value of g = 9.8057 + 0.0005 m/s2 in your analysis, which applies in Morris, MN.
Questions:
(a)

One of the assumptions in the analysis was that there was no friction in this case. What effect would
friction have had on the results (the relationship between your experimental and theoretical values) you
found? Is there evidence of any friction in your calculated results? Where might friction arise in this
apparatus? List several possible sources.

(b)

One of the assumptions in the analysis was that the glider was at rest when the timing began. If you
were careful, this assumption should be quite good, though it cannot be perfectly accurate. If you were
not so careful and the glider had some modest velocity when it tripped the first photo gate, what would
be the affect on the relationship between your theoretical and experimental acceleration? How can this
possible error be distinguished from the possibility of friction in the apparatus?

(c) The larger the mass of the dropping block, for a given glider mass, the larger the acceleration of the
glider/block system. What is the limiting value of the magnitude of the acceleration that can be achieved in
this way? For example, what acceleration would you expect if you used your textbook as the dropping
block? How about tying your lab partner to the end of the (unbreakable) string and dropping him or her? If
your initial response was that unlimited accelerations could be achieved, think again! No calculations are
necessary.
Submission:
Completed electronic lab report by e-mail (spreadsheet)

37

Experiment 6: Centripetal Force

38

Experiment 6: Centripetal Force


Score: __________of 10

Name: ____________________________________

Pre-Lab Assignment for Experiment 6: Dealing with Random Errors


A.

In experiment 6, the centripetal force is calculated from the mass m, the radius r and the period T as

Fc

B.

4 2 mr
. Write the error equation for the relative error for Fc.
T2

Study this: In a certain measurement run, the period T has been measured 3 times using a photo gate,
see table. The systematic uncertainty of each time value is 1% of the time value, due to systematic
variations in the quality of the photo gates. We will need the average period with its respective
uncertainty in this experiment. However, the uncertainty of the period is not only determined by the
systematic error of the photo gate, but also by random factors during the execution of each single
experiment. Hence, the uncertainty of the average is larger than the individual uncertainties of the
measurement values. In such a measurement, the uncertainty of the average time is calculated as

Ti

1
Ni

1
Ni

Ti
1, N

Ti T

1, N

The first term is the average of the uncertainties of the individual measurements; the second term is the
standard deviation of the measurement values.
C.

Apply this step by step by filling in the white fields in the table below in order to find average time and its
uncertainty.
2
Time Ti (s)
Uncertainty Ti (s)

Ti T

(1% of Ti)
Measurement 1

0.695

Measurement 2

0.689

Measurement 3
Average

0.692

1
3

Ti

Ti
i 1

1
3i

Ti
1,3

1
Ni

Standard
deviation

Ti T
1, N

1
N

Uncertainty of
average

Ti T

i 1, N

Ti

In the spreadsheet, such tedious calculations are much simpler. Average and standard deviation can be
programmed as

AVERAGE

Ti
39

STDEV Ti

Experiment 6: Centripetal Force

40

Experiment 6: Centripetal Force

Laboratory Experiment #6: Centripetal Force


Objective:
To study how the centripetal force required to make an object undergo uniform circular motion varies with the
mass and speed of the object as well as with the radius of its circular path.
Introduction:
When an object moves under the influence of a net force that acts along the direction of motion, its speed
either increases or decreases. If the net force always acts at right angles to the direction of motion, it changes
the direction of motion of the object while leaving its speed unchanged. In both cases the objects velocity
vector is changing so the object is being accelerated.
An object moving in a circular path at a constant speed is said to be undergoing uniform circular motion. As
the object moves around the circular path, its direction changes continuously. Thus a net force whose
direction is continuously changing must be acting on it. The direction of this net force must be radial since the
velocity is tangential to the circular path and the net force must be perpendicular to the velocity if the speed is
to remain constant. A more careful analysis shows that the net force Fc must be directed toward the center of
the circle and must have a magnitude
mv 2
Fc
r
where m is the mass of the object, v is its speed, and r is the radius of the circle. A net force that always
points toward the center of a circle is called a centripetal force.
If the object takes time T to complete one revolution, its speed v is
2 r
v
T
so
2
m 2 r/ T
4 2mr
Fc
. (Eq.1)
r
T2
The experiment makes use of a centripetal-force apparatus in which a metal bob hanging is rotated in a
horizontal circle. A sketch of the apparatus with the bob rotating is shown in Figure 6.1(a).
cross arm
bob
m

axle

pointer

(a)

(b)
Figure 6.1: The centripetal force apparatus.
The metal bob of mass m hangs from a rod attached to a rotating axle. The bob is connected to the axle by a
horizontal spring. As the axle is spun by hand, the metal bob moves faster and faster, stretching the spring
as more and more horizontal force is needed to keep the bob moving in a circular path. The radius of the
circular path can be determined by moving a pointer horizontally until its tip is directly below the tip of the
metal bob. In actuality, the pointer is positioned first at some desired radius then the metal bob is spun till it
passes directly over the tip of the pointer. The person spinning the axle must watch the alignment between
41

Experiment 6: Centripetal Force


the metal bob and the pointer and continuously adjust the spinning rate, to compensate for slowing due to the
small amount of friction in the axle bearing.
A photogate timer is used to time the rotational period T of the metal bob while is it being spun at the desired
radius. The radius r is determined by measuring the distance from the pointer to the axle, then adding half
the diameter of the axle. The radius may be adjusted by moving the pointer in or out. The metal bob can be
taken off the apparatus to measure its mass m. Additional slotted masses can be piled on top of the metal
bob to adjust m.
The horizontal force exerted by the spring can also be determined when the metal bob is not rotating. A
second mass M is suspended from a string that runs over a pulley then horizontally to the side of the metal
bob opposite the spring (shown in Figure 6.1 (b)). The horizontal force exerted by the spring Fs must be
sufficient to support the weight of the hanging mass, so
Fs

(Eq. 2)

Mg.

The hanging mass M can be increased till the spring is extended so that the metal bob lines up with the
pointer. At this point the force exerted by the spring can be determined using Equation 2.
When the metal bob is rotating the only horizontal force acting on it is that of the spring, so in this case the
centripetal force required for circular motion is due to the spring. Thus Fc Fs in this situation, so

Mg

mr

(Eq. 3)

Notice that increasing M stretches the spring, thereby increasing r, but is completely independent of changes
in m. Similarly, m can be increased without affecting r (or M). Thus when M (and hence r) are fixed, the
spring force is a constant. Experimentally this means that changes in m must result in corresponding
2
changes in T to keep the right hand side of Equation 3 (more specifically, the ratio m / T ) a constant. On the
2
other hand, m / T is not fixed when r changes. This can be seen because r is not directly proportional to M.
As M approaches zero, r approaches the relaxed length of the spring, not zero. If the metal bob mass m is
held constant and the hanging mass M is varied, the radius r and the period T will both change in such a way
2
to ensure that the ratio r / T is directly proportional to M.
Equipment:
Centripetal force apparatus, weight hanger, weights including 500g sizes, ruler or meter stick, photogate timer
(t=1%), 10 kg capacity (m=1g) digital balance, and a piece of 3x5 card.
Procedure:
Place the centripetal force apparatus on the lab table so that the hanging mass M will clear the table edge
when it hangs down from the pulley. Adjust the apparatus' legs as necessary to ensure that the cross arm
rotates in a level horizontal plane.
Step 1:

Set the pointer location at its minimum distance from the axle. Disconnect the spring from the metal
bob and move the cross arm as necessary so that the freely hanging bob is lined up with the
pointer. Attach a strip of 3x5 cardstock horizontally to the top of the mass, extending out away from
the spring (see Figure 6.1(a)). Place the photo gate timer so the strip will trip the timer when
rotating at this radius. Select the appropriate timing mode to measure the period of rotation T. Set
the photo gate to 1 ms resolution.

Step 2:

Remove the metal bob from the apparatus and measure its mass. Also measure the radius r of the
circular path of the metal bob by measuring the distance from the center of the pointer to the center
of the axle.
42

Step 3:

Experiment 6: Centripetal Force


Reattach the bob to the apparatus. Attach the mass hanger and string to the metal bob using the
hook provided. Add mass to the hanger until the metal bob lines up with the pointer. Record the
value of M.

Step 4:

Changing mass: Remove the mass hanger, string, and hook from the bob. Twist the axle to start
the metal bob moving. Increase the rotational rate until the tip of the metal bob passes directly over
the pointer on every pass. Practice keeping the metal bob going at this rate for many revolutions.
Do not allow the rotating mass to strike the photo gate timer! Take 10 good measurements of
the period, clearing the photogate timer between measurements. (It helps to have the memory
function ON so that the timer does not restart on its own the next time the bob comes around.) The
measurements will not be during consecutive rotations because you need to write down each
period before taking the next.

Step 5:

Change m by adding a 100-gram slotted mass securely to the top of the metal bob using the
knurled nut on top of the bob to hold the added mass in place. (Do you need to measure M again?)
Measure the period 10 more times for this mass and radius combination.

Step 6:

Changing Radius: Leaving the amount of the mass on the bob the same, move the pointer to a new
location approximately 3/4 of the way out from the axle and determine the value of M needed to
stretch the spring that far. Be sure to adjust the cross arm when you move the pointer so that the
metal bob is lined up with the pointer when hanging freely. Measure the value of the radius
corresponding to this new pointer location. Measure the period 10 more times for this mass and
radius combination.

Step 7:

Calculate the average period T for each of the three data sets. Think about what uncertainty T to
associate with these average Ts. Beware! The variability in the periods is likely to exceed the
scale-limit of the photogate timer. Calculate the standard deviation of each set of periods to quantify
the reproducibility-uncertainty in this experiment.

Step 8:

Use Equation 1 to determine the centripetal force, complete with uncertainty, for each data set.

Step 9:

Use Equation 2 to determine the force exerted by the spring, complete with uncertainty, for each
data set.

Questions:
Do your data support the proposal that the centripetal force is independent of mass when the radius is held
fixed? Explain.

Submission:
Electronic spreadsheet.

43

Experiment 7: Collisions in one Dimension

44

Experiment 7: Collisions in one Dimension


Name: ____________________________________

Score: __________of 10

Pre-Lab Assignment for Experiment 7:


A. Write down the error equation for the momentum

L
m and the kinetic energy KE
t

1
L
m
2
t

B. For one measurement of an inelastic collision, an incident glider is colliding with a resting glider. The
momenta of all glicers before and after the values for the momentum of two gliders before and after
the collision are:
initial mom pi (kg m/s)
pi (kg m/s)
final mom pf (kg m/s)
pf (kg m/s)
incident glider
stationary
glider

0.0873

0.002

0.0316

0.0007

0.000

0.002

0.0466

0.0004

Total
momentum
pf pi
Find the total momentum of both gliders before and after the collision, together with their uncertainties.
Find the difference in momentum before and after, together with the uncertainty of that value. You will
need the Addition/Subtraction rule for uncertainty propagation.
C. Why is there an uncertainty associated with the initial speed of the resting glider? Explain.

D. The expected value for the difference in momentum before and after the collision is 0. Was this
experiment successful? Explain.

45

Experiment 7: Collisions in one Dimension

46

Experiment 7: Collisions in one Dimension

Laboratory Experiment #7: Collisions in One Dimension


Objectives:
To observe, qualitatively and quantitatively, the behavior of particles when they collide and to see whether
momentum and kinetic energy are conserved in such collisions.
Introduction:
When two or more particles collide, there is a transferring of momentum between them. It is an observed fact
that the total momentum of the particles is the same after the collision as before.
Momentum (p) is defined as the mass of the object multiplied by its velocity:
p = mv.
The conservation of momentum is then written
pf

pi

where the summation symbol indicates that the momenta of all particles are to be added. The subscripts "f"
and "i" denote final and initial values, respectively. For a collision involving only two particles,
p1f + p2f = p1i + p2i.
If, furthermore, the motion is restricted to one dimension, this becomes
p1f + p2f = p1i + p2i
or
m1v1f + m2v2f = m1v1i + m2v2i.
This conservation of momentum will hold true as long as there is no external force acting to change the
velocities (such as gravity).
Kinetic energy, on the other hand, is not always conserved in a collision. When it is conserved, the collision is
called "elastic". Then
Kf

Ki

or

m1
m2
m1
m2
2
2
2
2
v 1f
v2 f
v1i
v 2i .
2
2
2
2
When the kinetic energy is not conserved, the collision is "inelastic". In an inelastic collision, the kinetic energy
will decrease as a result of the collision (unless an explosion occurs during the collision). The lost kinetic
energy is converted into heat, sound, or some other form of energy.
The air track is a very convenient device for studying collisions in one dimension. If the track is properly
leveled, the downward force of gravity on the glider is completely negated by the normal force that the track
exerts on the glider. Then gravity will not influence the velocities of the gliders and momentum will be
conserved. Elastic collisions will result if two gliders are sent toward one another so that their contact is via the
band springs that are mounted on one end of each glider. (During a collision, the springs will temporarily store
some energy in potential form but, because they are conservative in nature, this energy will be returned to

47

Experiment 7: Collisions in one Dimension


kinetic form as the gliders separate. Thus kinetic energy is conserved and the collision is elastic.) On the other
hand , if the two gliders somehow stick together after the collision, it will be inelastic. Such a collision can be
achieved by using the needle and socket ends of the two gliders--in a collision, the socket will grip the needle,
holding the two gliders together.
Equipment:
Air track, two gliders with flags, additional glider-masses, appropriate bumpers for elastic and inelastic
collisions, two electronic timers with photo gates (t=1%), Vernier caliper, and a 1200g (m=.1g) digital
balance.
Procedure:
Carefully level the air track using a glider with no attachments on it (no flag, no bumpers). If the track is not
perfectly straight, try to characterize where it has dips and/or peaks so that you can do the experiment in such
as way as to minimize the effect of such imperfections. Load the gliders symmetrically and adjust the airflow as
necessary, to make sure that they never scrape on the air track.
Step 1:

Qualitatively study elastic collisions, in which the gliders bounce apart after the collision. Record your
observations on your data sheet. Limit your studies to collisions where a moving glider (call it glider
A) strikes a stationary glider (call it glider B). Look for trends in the collision outcome as the mass of
glider A is adjusted with respect to that of glider B. (It should be sufficient to check (1) a light glider
strikes a light glider, (2) a light glider strikes a heavy glider, (3) a heavy glider strikes a light glider,
and (4) a heavy glider strikes a heavy glider. To change a glider from "light" to "heavy", add two 50-g
masses to it, one on each side.) Your report should include a paragraph summarizing your
observations. Be sure to answer (at least) the following questions.
(1) Under what conditions (if any) does glider A continue forward/bounce backward/not move after
the collision?
(2) Under what conditions (if any) does glider B move forward/move backward/not move after the
collision?
(3) Under what conditions (if any) is the speed of glider B after the collision greater than/less
than/equal to the speed of glider A before the collision?

Step 2:

Qualitatively study totally inelastic collisions, in which the gliders stick together after the collision.
Record your observations on your data sheet. Limit your studies to collisions where one glider is
initially at rest and label the gliders as above. Again, let the relative masses of the gliders be the
primary adjustable parameter. Be sure to answer (at least) the following questions in your summaryof-observations paragraph.
(1) Under what conditions (if any) do the two gliders (now locked together) move forward/move
backward/not move after the collision?
(2) Under what conditions (if any) is the speed of the gliders after the collision greater than/equal
to/less than half the speed of the incident glider?

Step 3:

Now quantitatively study elastic collisions and inelastic collisions when a light incident glider strikes a
heavy stationary glider. Attach the elastic collision attachment on one end and the inelastic
attachment to the other end of each glider. Measure the length of the glider flags using the Vernier
caliper. Set the photo gate resolutions to 1 ms.
Perform two good trials of each type of collision. For each trial, determine the initial and final
velocities of both gliders by measuring the time required for each glider to traverse a pre-determined
distance (i.e. by measuring the time it takes for a glider flag to pass completely through a photogate).
No more than three different velocities will ever need to be determined, since the heavy glider
always begins at rest. Each trial is independent because the initial push given to the incident glider is

48

Experiment 7: Collisions in one Dimension


different every time. Thus it is not possible to compare different trials as a check of the accuracy of
the measurements. Good technique is vital! Estimate your uncertainties thoughtfully! Read the
comments below.
In each trial for each type of collision, compute the momentum, p

mL
, of each glider, both before
t

and after the collision. Remember that p's are vectors, so sometimes they can be negative. Also
compute the change of the total momentum: p
pf
pi . Next, compute the kinetic energy,
2

m L
, of each glider before and after the collision, as well as the change in total kinetic
2 t

energy:

Kf

Ki . Remember that K's should never be negative. Dont forget to compute

the uncertainties of each of these quantities. Remember that some (but not all!) of the final
differences are expected to equal zero (within uncertainty).
Comments on the timing systems:
If you are using two independent photogate timer systems:
In some cases, you will need to take two measurements with one of the timer systems. You may
either do this by quickly reading and resetting the timer in the middle of the trial or by making use of
the memory function of the timer system. If you use the latter approach, the time in the memory is
actually the sum of the two times. To retrieve the time of the first event from memory, first note the
time of the second event on the display, then press the toggle switch to see the value in memory. To
get the time of the first event, subtract the time of the second event from the displayed value.
If you are using one timer system with two photo gates connected to it:
Locate the photo gates so that only one photo gate is making a measurement at any one time. The
timer unit can only store 2 timing values simultaneously (using the memory function). Thus it will be
necessary to reset the timer in the middle of any run requiring 3 timing measurements (after noting
the initial time value(s), of course). Many practice trials may be necessary to arrange the photo gates
so that the gliders pass through them in the correct order.
When two events occur in succession without the timer being reset, the time in the memory is
actually the sum of the two times. To retrieve the time of the first event from memory, first note the
time of the second event on the display, then press the toggle switch to see the value in memory. To
get the time of the first event, subtract the time of the second event from the displayed value.
Submission:
Electronic spreadsheet

49

Experiment 8: Ballistic Pendulum

50

Experiment 8: Ballistic Pendulum


Name: ____________________________________

Score: __________of 10

Pre-Lab Assignment for Experiment 8: Logarithm Rule


A. Using the spring-loaded mechanism of the ballistic pendulum in order to propel a metal ball in a
horizontal projectile motion off the table, the initial velocity, v, of the projectile can be found from
measurements of the vertical fall distance, y, and the horizontal range, x, using
g
.
2y
Use the logarithm rule to find the error equation for the speed v.
v

B. If x = 2.81 .03 m, g = 9.806 .0005 m/s , and y = 1.02 .01 m, find velocity v and its relative error
as v v/v.

C. For the ballistic pendulum, the initial speed of the projectile is calculated as v o
Use the logarithm rule to derive the equation for the relative error
vo
M
M m
1 h 1 g
=
+
+
+
vo
m M m(m M) 2 h 2 g

51

m M
2gh .
m

Experiment 8: Ballistic Pendulum

52

Experiment 8: Ballistic Pendulum

Laboratory Experiment #8: Ballistic Pendulum


Objective:
To determine the velocity of a projectile using a ballistic pendulum and the laws of conservation of energy and
momentum.
Introduction:
In this experiment a metal ball of known mass m is shot horizontally using a spring gun. When the ball is
caught by the hollow bob of a pendulum, the pendulum and the ball swing together along the arc of a vertical
circle. The velocity of projection can be determined from the height of the swing and the laws of conservation
of energy and momentum.
The apparatus, shown in Figure 8.1, consists of a light pendulum rod that has a hollowed-out bob at its lower
end and is suspended at its upper end from a horizontal suspension rod. It swings freely in a vertical plane
about the suspension rod axis. The hollow interior of the bob has a rubber o-ring which allows the ball to enter
but prevents it from coming out once inside; it thereby is able to catch the ball. The metal ball (with a hole
along its diameter) is placed on the horizontal rod of the spring gun. By pushing on the ball with the palm of
your hand, the spring can be compressed and locked. A squeeze of the trigger shoots the ball horizontally so
that it hits the freely hanging pendulum bob where it is captured. The pendulum with the embedded ball
swings forward and upward until it stops at the highest point of its swing. A ratcheting lever attached to the
bottom of the bob engages the curved rack and holds the pendulum at that position . (The rack has a series of
horizontal ridges that are spaced about 1 mm apart. By engaging the notch nearest to the highest position of
the bob, the lever prevents the pendulum from swinging back.) An index attached to the pendulum bob marks
the center of gravity of the loaded pendulum.

Figure 8.1: Ballistic Pendulum


Method 1:
To determine the initial velocity of the ball, measure the height h through which the center of mass of the
loaded pendulum rises as the ball collides with the bob at speed vo. The combined pendulum and ball move
with a speed v just after the collision. Let m be the mass of the ball and M be the mass of the pendulum. Then
using the principle of momentum conservation
m vo = (m + M ) v .

53

(1)

Experiment 8: Ballistic Pendulum


The speed v can be related to the height using the principle of conservation of energy. At the start of the
swing the energy is completely kinetic ((m+M)v2/2) and at the end it is completely potential ((m+M)gh). Hence
we have
Ei = Ef
(m+M) v2 / 2 = (m+M)gh
Then

2gh

(2)

m M
2gh
m

(3)

v
Combining equations (1) and (2) one finds

vo

The corresponding error equation is given by

vo
M
M m
1 h 1 g
=
+
+
+
vo
m M m(m M) 2 h 2 g

(4)

Method 2:
The velocity vo can also be determined from the range of the ball. With the pendulum removed, the ball is
fired so that it hits the floor. The distance x along the horizontal from the gun to the point on the floor as well
as the vertical distance y directly below the gun to the floor are measured. See Figure 8.2. If t is the time of
flight of the ball as it travels from the gun to the floor ,
x = vot
and

gt2
.
2

By eliminating t from these two equations one obtains the trajectory:

vo

g
2y

(5)

A knowledge of x and y thus enables us to determine vo.

Equipment:
Ballistic pendulum apparatus, 150 g (m=.01g) digital mass balance, meter stick, ruler, paper and a thin piece
of wood or cork.

54

Experiment 8: Ballistic Pendulum


Procedure:
Check to see that the pendulum hangs freely and can oscillate without friction. Load the gun and then release
the trigger to ensure that the rubber 0-ring operates effectively as a catch and that the lever engages with the
notched rack.
Method 1:
Step 1: Measure the mass of the ball, m. The mass of the pendulum, M, is noted on the apparatus.
Step 2:

Make sure the pendulum hangs freely. Determine the height h 1 of the center of mass from the base
of the apparatus using a ruler.

Step 3:

Place the metal ball on the rod of the spring gun and by pushing on it, compress and lock the
spring. Pull the trigger to release the ball.

Step 4:

Note the notch number on the rack that the lever engaged.

Step 5:

Lift the lever out of the notch and move the pendulum back to the vertical position. Remove the ball
carefully from the pendulum bob and place it back on the spring gun. With the pendulum stationary
in the vertical position, fire the gun again and note the notch number.

Step 6:

Repeat the above process for a total of ten trials. (Repeat any unsuccessful trials.)

Step 7:

Determine the average value of the notch numbers. Round it upward to the next integer.
(Understand why.) Move the pendulum up the rack until the lever is in the notch corresponding to
that rounded average value. With the pendulum in this position determine the height h 2 of the
center of mass from the base of the apparatus. The difference h2 - h1 equals the height, h. Estimate
h from the highest and lowest notch values and your own judgment. A second contribution to h
arises from the spread in notch values. Calculate this using the standard deviation. The total h is
the sum of these two contributions.

Step 8:

Determine vo from equation (3) and the error vo from equation (4).

Method 2:
Step 1: Prepare to shoot the ball onto the floor. Place the pendulum high on the rack so that it is out of the
path of the ball. Align the base of the apparatus with a corner of the table in a way that can be
easily and precisely reproduced before each firing.
Step 2:

Locate the place on the floor where the ball lands by shooting the gun a couple of times. Tape a
sheet of paper (with a thin wood or cork board underneath) to the floor at this location. See Figure
2. One member of the group could stand behind the landing zone with a coat to stop the ball.

Step 3:

Shoot the ball ten times. Number the marks on the paper from 1 to 10. Check the location of the
base before each shot.

Step 4:

Determine the values of the horizontal distances from the point where the ball leaves the gun to the

marks on the paper. Call them x1 x1, x2 x2, x3 x3, etc. Find the average range x x .

Step 5:

Determine the height above the floor of the point where the ball leaves the gun. In estimating
uncertainty, consider the levelness of the floor.

Step 6:

Use equation ( 5 ) to obtain vo vo. (Hand in the target piece of paper with your report.)

55

Experiment 8: Ballistic Pendulum

Figure 8.2: Set-up for Method 2 .


Comments:
Compare the values of vo as obtained in method 1 and method 2. Discuss whether they agree within their
uncertainties. What is the percentage difference between them? What sources of error were not taken into
account by the uncertainties?
The acceleration of gravity in Morris, MN is 9.8057 .0005 m/s 2.
Question:
Explain why the average notch number in Step 7 of Method 1 should be rounded up to the next integer.
Submission:
electronic spreadsheet

56

Experiment 9: Static Equilibrium


Name: ____________________________________

Score: __________of 10

Pre-Lab Assignment for Experiment 9:

The torque balance for the calculation of an unknown mass leads us to the expression M 2

m1

x1 x pp
x2

x pp

in which all variables are measured quantities. Show that the correct error equation for M2 is:

M2
M2

m1
m1

x1
x pp

x2
x1

x2

x pp

x2
x pp

x1

x1 x2

x pp

x pp

You can either use the basic equation for uncertainty propagation or the logarithm rule. In both cases, much
care is needed to place the absolute-value brackets into the correct place.

57

Experiment 9: Static Equilibrium

58

Experiment 9: Static Equilibrium

Laboratory Experiment #9: Static Equilibrium


Objective:
To study the conditions of static equilibrium.
Introduction:
Newton's third law states:
Fnet =

(1)

F = ma

where F is an external force acting on the system. Note that this vector equation can be rewritten as three separate scalar
equations, one along each of the x, y and z directions:
(Fnet )x =
(Fnet )y =

Fx = m a x
Fy = m a y

(Fnet )z =

Fz = m a z

(2)

For a system of particles, equation (1) implies that if the center of mass of the system is to remain stationary (i.e. v = a = 0), then
F 0
or
Fx
Fy
Fz 0 .
We could imagine a situation in which the center of mass of a system of particles is stationary while many or all of the particles
comprising the body are not (as in the case of a rigid body rotating about its center of mass. There must be a different "force-like"
quantity acting here. This quantity which causes a rigid body to change its rotational motion is called a torque, . A torque with
respect to an origin O due to a force F at a point P whose position is specified by the vector r with respect to O is given by
=rxF

(3)

See Figure 9.1.

->

->
O

r
y

x
Figure 9.1
The magnitude of

is given by
= r F sin = ( r sin ) F = L F

where L is called the lever arm of the force F.

59

(4)

Experiment 9: Static Equilibrium

Consider now forces in a plane. (See Figure 9.2) Let the origin be at O. Suppose there is a force F (parallel
to the x-axis for simplicity) acting on a rigid body at the point P whose position is given by r with respect to
O.

rigid body
y

line of action of the


force F

P
r

L = r sin

x
O
Figure 9.2
The torque is again given by equation (4) above as
=
) F = L F
We can imagine that the rigid body will then rotate clockwise about an axis passing perpendicular to O.
From Figure 9.2, we can interpret the lever arm L = r sin
rotation (in this case the axis of rotation is passing perpendicular to O ) to the line of action of the force. The
point through which the axis of rotation passes is also known as the pivot point. Hence for simplicity, for
forces on a plane,
=
) F = L F
(5)
where L is the perpendicular distance from the axis of rotation (or pivot point) to the line of action of the
force F.
Since the rotational acceleration resulting from the torque can be either clockwise or counter clockwise,
we have to assign to it a sign convention. A common convention is to let > 0 if the force (acting alone)
would cause the system to rotate counterclockwise and let < 0 if the force (acting alone) would cause the
system to rotate clockwise.
Static equilibrium is defined as the condition in which a system is not moving at all. (Note that a system may
be in equilibrium but still be moving if its center of mass moves at a constant velocity). The condition F = 0
is insufficient for static equilibrium since a system's center of mass may be stationary while the system
rotates. Hence, the sum of the torques acting on the system must also equal zero. In summary, for static
equilibrium to occur
F

0 and

(6)

For forces on a plane (like the x-y plane), we can reduce the six equations (three vector components for
each equation) resulting from equations (6) into three equations,
F x = 0,

F y = 0 and

ccw

cw

0 .

(7)

where the last of these conditions has been rewritten in terms of clockwise and counterclockwise torques.

60

Experiment 9: Static Equilibrium


Equipment:
Digital (1200g) mass balance (m=.1g), meter stick, set of weights, weight hangers, pivoting meter stick
holder, vertical rod mounted on table edge, object of unknown mass
Procedure:
Part A:
Step 1:

Determine and record the mass of the meter stick. Throughout the experiment measure the total
masses suspended, including the clamps, on the digital balance.

Step 2:

Find the center of mass of the meter stick by hanging it at the 50 cm mark using the lever holder
and adjusting the location until it hangs approximately parallel to the ground. Record this center of
mass location. Hang the meter stick by the center of mass. The center of mass is now the socalled pivot point, xpp.

Step 3:

Hang a known mass m1 at some location x1 to the left of the pivot point. By trial and error, put
another mass m 2 of a different value at some location x2 to the right of the pivot point until the
system is in static equilibrium. Record m 1, m2, x1 and x2 .

Step 4:

Calculate the sum of the torques in the system with respect to the pivot point.

Step 5:

Repeat steps 3 and 4 for two more different sets of values of m 1, m2, x1 and x2.

Step 6:

Using the last set of data (i.e. trial 3), also calculate the sum of the torques on the system about
the 0.00 cm and 75.00 cm marks. (Do not forget the torques due to the vertical forces acting on
the meter stick at the hinge since the hinge is no longer the pivot point!) As before, the sum of the
torques should be zero. Be very careful to re-identify which torques are cw and which are ccw for
each change of pivot points. (Hint: you should have fours torques in each of these two trials.)

Part B:
Step 1:

Step 2:

Using the set-up, find the mass of the object of unknown mass by balancing it against a known
mass. Be sure your object hangs freely so that the center of mass is directly below the point of
suspension.
Find the mass of the unknown, M2, using the mass balance. Compare to the value found in Step
1 of Part 2.

Submission:
completed spreadsheet

61

Experiment 10: Simple Harmonic Motion

62

Experiment 10: Simple Harmonic Motion


Name: ____________________________________

Score: __________of 10

Pre-Lab Assignment for Experiment 10:


A stop watch is used to measure the period of an oscillating mass-spring system, precision 1/100 of a
second. For both series of measurements:
a) Assign the uncertainty of each individual time measurement, considering the precision of the stop
watch.(1)
b) Find the experimental result for the period of oscillation, using the average of the data. (3)
c) Find the uncertainty for this result. Remember that random errors are considered using the
standard deviation of the data and the systematic precision errors. Consider the difference in
technique (measuring 1 versus 10 oscillations) properly. (4)
Student A measures one period per data point.
Exp result for period

Data
T-1 (s)

Tavg (s)

T-1 (s)

Tavg (s)

1.38

1.47
1.30
Student B measures 10 periods per data point.
Exp result for period

Data
T-10 (s)

T-10 (s)

T-10avg (s)

T-10avg (s)

13.25

13.44
13.37
d) Why is student B doing a better experiment? (2)

63

Tavg (s)

Tavg (s)

Experiment 10: Simple Harmonic Motion

64

Experiment 10: Simple Harmonic Motion

Laboratory Experiment #10: Hooke's Law and Simple Harmonic Motion


Objectives:
To observe whether a rubber band and a helical spring obey Hooke's Law. To measure the period of
simple harmonic motion.
Introduction:
All material substances will deform in response to applied forces but the manner in which they deform
varies from substance to substance. A material is said to deform "elastically" under the application of a
force if, when the force is removed, the material resumes its former shape. On the other hand, a material
is said to deform "plastically" if it deforms permanently and without fractures in response to the applied
forces. A third possibility is that the material will break or fracture. For small forces, the deformation will be
elastic. Thus the material will return to its former shape when the force is removed. The most stress a
material object can sustain without the deformation becoming permanent is called the "elastic limit". Once
the force exceeds that limit, the deformation will be plastic. Finally, at some even greater force, the object
will break under the stress.
During an elastic deformation, internal forces develop which act so as to restore the object to its original
shape. These restoring forces are generated by the molecular bonds in the object. (Bonds resist being
stretched or compressed; they prefer their equilibrium length.) For many materials the magnitude of the
restoring force is proportional to the amount of deformation. Such materials are said to obey Hooke's
Law:
F = - k (r - ro)

(1)

where (r-r0) indicates the vector displacement from equilibrium of some point in the object and "k" is a
constant that characterizes the particular object.
Once an object is known to obey Hooke's Law, one can write Newton's Second Law to describe the
dynamics of that object:
m ax = - k (x - x0)
or
d 2x
m 2
k(x x o ) .
dt

(2)

Solving this differential equation, one finds that the object will oscillate sinusoidally with a period that does
not depend on the amplitude. (The period is the time taken for a single oscillation.) From the above
differential equation, one can show the period (T) to be
T

m
k

A good example of such an oscillator is a mass "M" suspended from a spring of force constant "k" and
mass "m". If the spring obeys Hooke's Law, then the mass can be set into an up-and-down motion whose
period can be found from

m /3
.
k

65

(3)

Experiment 10: Simple Harmonic Motion


Notice that one-third of the spring's mass is added to the mass of the suspended object. This occurs
because the spring also participates in the motion but only in a partial fashion (one-third part, in fact).
Equipment:
Helical spring and rubber band mounted on a stand, weights, timer, meter stick, and 1200g (m=.1g)
digital balance.
Procedure:
Part 1
Suspend a 50g* weight hanger from the rubber band and record the location of the lower end of the
rubber band with respect to some fixed reference point. Then add 50g and record the new location.
Continue adding mass in 50g increments and recording locations until you sense that the elastic limit is
being approached. Then stop. (Rubber bands have no region of plastic behavior--they go from elastic
right into breaking.) Measure the hanging mass each time using the digital balance.
Compute the force exerted by the rubber band on the mass. Plot the force as a function of the location of
the weight hanger. Notice whether the points form a straight line.
Now repeat this procedure for the spring. The elastic limit of the spring may be much lower than that of
the rubber band so be careful. Notice whether the data points form a straight line. If they do, the slope of
the best linear fit gives the magnitude of the value of the spring constant k. (Spring constants are always
positive but, depending on whether you measured locations upward or downward, you might have a
positive or negative slope on the graph.)
(*The instructor may suggest different increment values to suit the equipment being used.)
Part 2:
Suspend a M=100g from the spring and set the system into oscillation. (Use a fairly small amplitude,
perhaps 1 cm.) Measure the time for 20 cycles, T 20. Repeat the measurement for two more trials to obtain
three values for T20. Average the three values and divide by 20 to get T avg.
Add 50 g to the mass and conduct three trials for a second case. Likewise adding 50g increments,
measure times for third, fourth, fifth, and sixth cases. For each case divide the average time by 20 to
obtain the period, Tavg. don't forget to compute Tavg. as well.
Plot the square of the values found for T avg. as a function of the mass hung on the spring. Determine the
slope of this graph. Since
4 2
4 2m
T2
M
k
3k
the slope of the graph should equal

4
k

Using this relation, find the value of k from the slope of your graph. Compare to the value found in Part 1.

66

Experiment 10: Simple Harmonic Motion


Comments:
You may use Excel for the graphs and the slope calculation.
For the slope and its uncertainty: include a linear trend line in the graph, display the equation; read the
slope from the equation. Since we can not use the graphical techniques learned in earlier experiments on
the computer, you may estimate the uncertainty of the slope by using the data points furthest apart from
each other and using the slope uncertainty as

s
s

y2
y2

y1

x2
x2

y1

x1
x1

The acceleration of gravity in Morris, MN is 9.8057 .0005 m/s 2.


Question:
With a 200-g mass attached to the spring, measure the time for 10 oscillations for different amplitudes.
Start the first trial by lifting the mass by approximately 1.0 cm, and the second trial by lifting by 2.0 cm.
What is your observation?
Submission:
electronic spreadsheet

67

Experiment 11: Standing waves on a string

68

Experiment 11: Standing waves on a string


Name: ____________________________________

Score: __________of 10

Pre-Lab Assignment for Experiment 11:


The frequency f and the half wavelength d of a standing wave on a string are related by

1 F
2d

using the tension F and the linear mass density . In this experiment, you will vary the tension in the string,
and measure d for several different standing waves, with the goal of finding the frequency f. While one could
find a value for f from each single data point, it is often a better strategy to use the ensemble of all your data
and transform them into a linear plot. Thus, the random errors made in each single point become diminished,
and the experimental result is representative of all of your data points.
The equation above can be written as
a straight line of slope

4 f 2 d 2 . In a plot of F versus d2, data points should lie along

4 f 2 as in the example below. Find slope and uncertainty of the slope for the best-

fit line in the graph below. Use techniques discussed in earlier experiments. Find the respective frequency f,
if =0.315 g/m.
d(m)
0.570
0.423
0.338
0.283

d(m)
0.015
0.015
0.015
0.015

d (m )
0.325
0.179
0.114
0.080

d (m )
0.017
0.013
0.010
0.009

F (N)
5.47
3.04
1.96
1.34

F (N)
0.34
0.10
0.10
0.03

7.0

6.0

tension F (N)

5.0

4.0

3.0

2.0

1.0

0.0
0.00

0.05

0.10

0.15

0.20
d2 (m 2 )

69

0.25

0.30

0.35

0.40

Experiment 11: Standing waves on a string

70

Experiment 11: Standing waves on a string

Laboratory Experiment # 11: Standing Waves in a String


Objective:
To study transverse waves in a taut string.

Introduction:
A mechanical wave is a disturbance in a material medium that propagates through the medium at a speed
characteristic of that medium. Such a wave transports energy from one place to another without any transfer
of matter. Examples of mechanical waves are sound (including ultrasound), vibrating strings (guitar, piano,
etc), water waves, and seismic waves. Electromagnetic waves are not included -- they do not require any
material medium for propagation.
Sound is an example of a longitudinal wave. The medium is (generally) air and the disturbance is a deviation
from the equilibrium air pressure. Such a disturbance will move across a room (for example, from a lecturer's
mouth to a student's ear) without any net movement of air. Sound is "longitudinal" because the particles
comprising the medium oscillate along the direction of the wave's propagation.
When a stretched string is plucked or bowed, transverse waves propagate along the string. The medium is
the string and the disturbance is a deviation from the equilibrium (straight line) configuration. Such a wave is
transverse because the particles comprising the medium move perpendicular to the direction of the wave
propagation. In the case of musical instruments, a significant portion of the vibrating strings energy is
transferred to the air creating sound waves. (This transfer is facilitated by a resonating cavity such as a
guitar or violin body or by a sounding board such as in a piano.)
Water waves are neither longitudinal nor transverse. The medium is water and the disturbance is a deviation
from the (equilibrium) flat surface. The particles of the medium actually move in circles as the wave moves.
By observing the motion of a leaf floating in a lake, one can verify that the wave does not carry the leaf (or,
therefore, the water) along.
Both longitudinal and transverse seismic waves exist. The medium is (generally) rock and the disturbance is
a displacement of rock particles away from equilibrium positions. Longitudinal (or "compressional") seismic
waves move faster than transverse (or "shear") seismic waves. The fact that shear waves (created by
earthquakes) cannot pass through the earth's core provides evidence that the core is liquid.
Let us now consider the stretched string in greater detail. By applying Newton's laws to the string we can
mathematically describe the motion of the string. Consider an infinitesimal segment of the string when a
disturbance is present as in Figure 11.1.

y(x)

Fo

x x+dx

Figure 11.1: Analysis of a segment of the string.

71

Experiment 11: Standing waves on a string


The segment, for which we shall write Newton's second law, has mass dm and length dx. (Assume only a
slight deviation from equilibrium.) Although the string tension, F o, is applied equally to both ends of the
segment there is a non-zero net force because of a difference (albeit infinitesimal) in the two angles. The net
force on the segment is
dFy = Fosin

- Fo sin

(Only the y-component is significant.) For small angles we can rewrite this as
dFy = Fo(tan

y
xx

Fo

- tan 1)

dx

Fo

y
xx

dx

The last step follows from the definition of a derivative.

dg(x)
dx

g(x dx) g(x)


; with g
dx

y
x

According to Newton's second law


Fne t

ma

Then

dFy = dm

or
2

Fo

Then

dx = dm

y
2

Fo

dm
where = dx is the linear mass density of the string. This differential equation is of the same form as the
"wave equation":
2
2
y
y
2
=
v
2
2
t
x

72

Experiment 11: Standing waves on a string


Its solutions are waves:

y (x,t) = yocos(kx- t)
where the speed v of the wave is given by v =

/ k.

Substitution of this solution back into the wave equation results in

yo cos(kx- t) = - k 2

Fo

yo cos(kx- t)

Thus the given function is a solution provided the relation


2

k2

Fo

is satisfied. Hence we require


Fo
k

Since the quantity on the left is equal to the wave velocity, we obtain an expression relating the speed of
wave propagation to the tension and mass density:
Fo

for one wavelength to pass a given point is called the period (T).

Then

Since T is the inverse of f,

v=
Then
f

Fo

or
f

1 Fo

(1)

Unfortunately, measuring the wavelength of a moving wave can be rather difficult. One way to circumvent
this difficulty is to utilize standing wave patterns. These can be created in strings that have both ends fixed.
(Children playing with ropes often discover standing waves.) The wave reflects back when it encounters an

73

Experiment 11: Standing waves on a string


end. If the string length and the wavelength are properly related, the reflected wave reinforces the incoming
wave resulting in a large-amplitude quasi-stationary pattern called a standing wave.
The proper relationship between string length (L) and wavelength ( ) necessary for standing waves can be
discerned from a few sketches:

Figure 11.2: Standing waves.


Nodes are denoted by "N". These are points on the string that do not move. The fact that a node must occur
at each end results in standing waves occurring only for
wavelength of a standing wave is easily inferred from the distances between nodes.

Equipment:
String vibrator (Sargent Welch #3256A), white cord about 2m long, a roughly 30cm segment of identical
cord, support post and c-clamp, pulley with table-edge mount, weights, weight holder, meter-stick, a 1200g
capacity (m=.1 g) digital balance, and a 150 g (m=.01 g) digital balance.

Procedure:
Step 1: The string should run from the vibrator (mounted at one end of the table) over the pulley (mounted at
the other end of the table about 1.5 m away ). On the dangling end, a weight hanger (50g) should be
suspended. Plug in the vibrator. It vibrates at 120 .2 Hz (twice the line frequency).
Step 2: By adding mass to the weight hanger, create standing waves in which the string vibrates in three,
four, five, six, and seven segments (loops). For each, look for the range of mass values that cause a stable
wave pattern of maximum amplitude and optimal stability. For your measured value, record the midpoint of
the range. (Use the digital mass balance to determine the mass hanging on the string.) The net uncertainty is
then half the range plus the uncertainty of the balance reading. Measure the node-node distance, d, and
record it. The wavelength will be twice this value.
Step 3: Find the mass per unit length of string using the roughly 80 cm segment. When measuring the length
consider that the string stretches under tension. (Assign an uncertainty to cover a reasonable range of
lengths.)

74

Experiment 11: Standing waves on a string


Step 4: To obtain the frequency of oscillation of the string, plot the string tension as a function of d 2. Since
1 F
f
and
= 2d,
we can write
F = (4 f2)d2

(2)

Then a plot of F as a function of d2 should yield a straight line whose slope equals 4 f2. From the
slope of the line that best fits your data, solve for f and compare to the known frequency. Be careful to use
the proper units on your graph.

Comments:
The end of the string at the vibrator is not exactly a node. Do not make that assumption.
Submission:
electronic spreadsheet

75

Experiment 12: Calorimetry

76

Experiment 12: Calorimetry


Name: ____________________________________

Score: __________of 10

Pre-Lab Assignment for Experiment 12:


A. In the calorimetry experiment, a basic term in the heat-balance equation is Q = m c (Tf - Ti).
Show that the corresponding error equation is

Q
m
c
=
+
Q
m
c

Tf
Tf

Ti
Ti

o
o
B. In which mass ratio mCu : mW would you have to mix 100 C water and 0 C copper in order to

achieve a mixture at a final temperature of 50 C? The specific heat of water is 1 cal/g/K, the
3
3
specific heat of copper is 0.09 cal/g/K. Considering the density of 1 g/cm for water and 9 g/cm
for copper, find the volume ratio VCu : VW (Of course, this experiment assumes no losses into the
environment or any other parts used in the experiment.)

77

Experiment 12: Calorimetry

78

Experiment 12: Calorimetry

Laboratory Experiment # 12: Calorimetry - Specific Heat


Objective:
To measure the specific heats of a couple of substances using calorimetric techniques.
Introduction:
When a substance is heated, its temperature rises. The temperature change (T = Tf -Ti) produced is
proportional to the heat input. The constant of proportionality is the heat capacity, C. Thus
Q = C T.
But the heat capacity is proportional to the size of the object being investigated. A more useful quantity is
the specific heat, which is the heat capacity per unit mass: c = C / m. Then
Q = m c T.
The value of c is a function only of the composition of an object. (It does also vary with temperature but
for small temperature ranges, it can be considered nearly constant.)
To determine the specific heat of a substance, calorimetry is employed. A heated (or cooled) sample of
that material is added to water inside a thermally isolated container (the calorimeter). The unknown
specific heat can be found from the heat-balance equation.
Q = 0.
which can be written in more detail as
(m c T)sample +(m c T)water + (m c T)cup = 0

(1)

Equipment:
A calorimeter (inner and outer aluminum cups separated by a non-conducting gasket, no stirring rod), a
digital thermometer, an electric steam generator, a digital mass balance, and two metal samples one of
which is aluminum.
Procedure:
Step 1:

Since the inner cup of the calorimeter will play a role in the heat flow, it will contribute a term to
the heat balance equation. Determine its mass. (It is composed of aluminum.)

Step 2:

Measure the mass of the empty dipper (the cup that fits inside the steam generator). Fill the
dipper roughly 3/4 full with aluminum bits and measure the total mass. Determine the mass of
the bits.

Step 3:

Fill the steam generator to about 2/3 full. (Never allow the water level to drop below 1/3 full.)
Plug it in and insert the dipper. Monitor the sample temperature by very carefully jiggling the
thermometer into the metal sample. (Lift and tip the dipper to facilitate this, if necessary.)

79

Experiment 12: Calorimetry


Step 4:

While the sample is heating, fill the inner calorimeter cup about 2/3 full with cool water.
Determine the mass of the water.

Step 5:

When the temperature of the sample has peaked, record that temperature. Move the
thermometer into the water. Stir gently for one minute, read the temperature, and then add the
heated metal bits quickly but without splashing. Stir or gently swirl the water and observe the
temperature. Record the "equilibrium" temperature.

Step 6:

Compute the specific heat using Eq. 1.

Step 7:

Pour off the water and place the metal bits on the appropriate towel to drain.

Step 8:

Fill the dipper approximately 1/3 full with a second, more dense, type of metal and repeat the
experiment.

Comments:
Please try not to spill the metal shot. Use a colander when discarding fluid into the sink.
Lab reports should include some discussion of errors caused by uncontrolled heat losses or gains.
Report uncertainties on all measured quantities, as usual, but only propagate errors forward in your
calculations that arise because of the temperature uncertainties. (These ought to be the dominant
sources of error.) Use the following expression:

Tf
c
=
c
Tf

T1
T1

Sample values:
Water (at

1.00 kcal/kg C

20oC)
Aluminum

.215 kcal/kg C

Copper

.0923 kcal/kg C

Lead

.0305 kcal/kg C

80

Tf
Tf

T2
T2

Appendix A: Forms of Lab Reports

Appendix A: Forms of Lab Reports


For paper lab reports:

Each student turns in a report of his/her own.

For electronic lab reports:


Individual reports containing group data and individual responses
are fine. If you submit one spreadsheet as a work group, everyone will receive the same grade.
Generic form of a paper lab report:
[Your name]
{Course][Section#]
[Semester and Year]

Partners:
[]
Title of Experiment

Preamble:
Briefly describe in your own words the objective and method of the lab by basically answering the
following questions:
- Which principles have you studied?
- What sort of experiment did you do to accomplish the objective?
Be concise but thorough, and limit the preamble to about 70 words.
Data and results:
Present your data in a neat and organized fashion, often as a spreadsheet. If the lab report is on
paper, the spreadsheet must be included on paper as well. Every measured quantity must include
a reasonable uncertainty. Results may also include graphs or diagrams.
Every experimental result must be having proper significant figures and be accompanied by a
proper uncertainty. You should, however, carry extra digits along in intermediate calculations, and
only round for the end result.
Include uncertainties in graphs in the form of error bars. Show the slope uncertainty directly on
the graph. Do not draw a straight line through data points that do not, at least theoretically, form
a straight line. Graphs should be drawn large on graph paper with only one graph on each sheet.
Label the axes with appropriate variable names and units and put a meaningful title on each
graph.
Error equations:
Exhibit each different equation used in your data analysis. Write each equation used followed by
the corresponding error equation. Do not show samples of trivial calculations such as unit
conversions.
Conclusions:
Respond to each (a) through (d) for each part of the experiment:
(a) Summarize and evaluate your findings. This includes comparing your result to
expectations and indicating whether or not they agree within uncertainty.
(b) Explain which of your measured quantities caused the most uncertainty in the
experimental results.
(c) Describe sources of error not quantitatively included in your calculations.
(d) What one realistic improvement would most improve your experimental result?
Questions:
Respond to the questions at the end of the lab experiment write-up in the manual.

- 81 -

Appendix B: Sample Lab Report

Appendix B: Sample Lab Report


Al Einstein
Phys 1101
Section 3
Spring, 2003

Partner:
Marie Curie

Hooke's Law and Simple Harmonic Motion


Preamble:
The purpose of this experiment was to study Hooke's Law and Simple Harmonic Motion. A rubber
band and a spring were checked for compliance with Hooke's Law by examining whether their
lengths increased linearly with the load they supported. The period of oscillation for the spring
was also measured under two different loads, in order to check for agreement with the theoretical
prediction of Simple Harmonic Motion.

Error Equations:
(1)

F = mg

(2)

t
N

(3)

F
F

m
m

T
T

m
k

T
T

g
g

t
t

1
2

m
m

N
N

1
2

- 82 -

k
k

Appendix B: Sample Lab Report


Experiment # 9
Hookes Law and Simple Harmonic Motion
Data & Results Sheet

Name
Partner

Al Einstein
Marie Curie

Part 1:
Rubber Band
mass (g)
50 0.2

L (m)
.609 0.0005

F (N)
.49 0.4%

position (cm)
60.9 0.05

150 0.4

.614 0.0005

1.47 0.3%

61.4 0.05

250 0.6

.620 0.0005

2.45 0.2 %

62.0 0.05

350 0.8

.626 0.0005

3.43 0.2%

62.6 0.05

450 1.0

.635 0.0005

4.41 0.2%

63.5 0.05

550 1.2

.644 0.0005

5.39 0.2%

64.4 0.05

650 1.4

.654 0.0005

6.37 0.2%

65.4 0.05

750 1.6

.666 0.0005

7.35 0.2%

66.6 0.05

850 1.8

.680 0.0005

8.33 0.2%

68.0 0.05

950 2.0

.694 0.0005

9.31 0.2%

69.4 0.05

1050 0.6

.710 0.0005

10.29 0.1%

71.0 0.05

1150 0.8

.726 0.0005

11.27 0.1%

72.6 0.05

1250 1.0

.742 0.0005

12.25 0.1%

74.2 0.05

Spring
mass (g)
50 0.2

position (cm)
59.5 0.05

L (m)
.595 0.0005

F (N)
.49 0.4%

100 0.4

61.2 0.05

.612 0.0005

.98 0.4%

150 0.4

63.1 0.05

.631 0.0005

1.47 0.3%

200 0.6

64.9 0.05

.649 0.0005

1.96 0.3%

250 0.6

66.7 0.05

.667 0.0005

2.45 0.2 %

300 0.8

68.6 0.05

.686 0.0005

2.94 0.3%

Part 2:
Oscillating Spring:
Number of oscillations, N = 10
Trial #

M (g)

M (kg)

Observed T (s)

Theoretical T
(s)

300 0.8

6.55 0.05

.300 0.0008

.655 1%

.660 1%

500 1.2

8.60 0.05

.500 .0012

.860 1%

.852 1%

- 83 -

Appendix B: Sample Lab Report


Conclusions:
(a)

Part 1: As can be seen from the graphs, the rubber band data points do not form a straight
line while the spring data points do. Thus the spring obeys Hooke's Law while the rubber
band does not. By drawing a straight-line fit to the spring graph, the force constant of the
spring was found to be k = 26.9 .3 N/m.
Part 2: The period of oscillation of the spring was experimentally measured to be 0.655 s
1% with the 300 g mass and 0.860 1% with the 500 g mass. (The uncertainty in these
periods arises entirely from the time measurements.) The theoretical periods for these two
cases were calculated to be 0.660 s 1% and 0.852 1% respectively. The experimental
values and the theoretical values agree within the indicated uncertainties as expected.

(b)

Part 1: Because

m
m

y
y

is larger than either

or

g
g

, the largest single contributor to

uncertainty in the results of Part 1 was the measurement of masses.


Part 2: Because

t
t

is larger than either

N
N

or

m
m

, the largest single contributor to

uncertainty in the results of Part 2 was the measurement of oscillation times,


(c)

t.

In Part 1, the support system flexed slightly as more mass was suspended. We did not take
this into account in the error calculations but the effect was very small.
In Part 2, the oscillating mass often started swinging side-to-side as it moved up and down.
By keeping oscillation amplitudes small, we minimized this but it may have had some effect
on the results.

(d)

In Part 1, the uncertainty could be reduced by using digital mass balances of greater
precision.
In Part 2,the uncertainty could be reduced by increasing the number oscillations for the
period timing in part two. This would make the denominators larger in Error Equation #2,
reducing the major experimental uncertainty.

- 84 -

Appendix B: Sample Lab Report

Graph #1: Force as Function of Position for Rubber Band


14

12

Force (N)

10

0
0.60

0.62

0.64

0.66

0.68

0.70

0.72

0.74

0.76

Position (m)

Graph#2: Force as a Function of Position for Spring

3.5
.685.0005, 3.05.005
3.0

Force (N)

2.5

Linear Fit:
Force = (26.9N/m) Position - 15.5N

2.0

1.5

slope/slope = rise/rise + run/run


=(.005+.005)/(3.05-.25) + (.0005+.0005)/(.685-.585)
=.01/2.80 + .001/.100
=1.3 %

1.0

0.5

.585.0005, .25.005
0.0
0.58

0.60

0.62

0.64
Position (m)

- 85 -

0.66

0.68

0.70

Appendix C: Significant Figures

Appendix C: Significant Figures


Objectives:
To learn the proper way of writing numbers that represent measured quantities. To become aware of
the uncertainty that such numbers always have. To learn to recognize the sources of that uncertainty.
To learn how to estimate the magnitude of that uncertainty.
Introduction:
In our everyday lives, we are all aware (to some degree) of the limited precision of measurements.
When we weigh ourselves on a bathroom scale, we recognize that the result is precise, at best, only
to the nearest pound (assuming the scale is properly calibrated). If you need a pane of glass to
replace a broken window, measurements precise to the nearest sixteenth of an inch will certainly
suffice. When a cookie recipe calls for a baking time of fourteen minutes, you need not feel
compelled to measure that time to the nearest second--ten or twenty seconds will not make much
difference. As you drive a car down the highway, a glance at the speedometer will inform you of your
speed to the nearest mile-per-hour. No reasonable person would claim to know the car's speed with
much greater precision. Further examples are all around us.
This concept of limited precision is very important in physics because all of the laws and principles of
physics are based upon experimental measurements--measurements that have limited precision.
The rough ideas of precision that suffice in our everyday lives are not adequate in physics. Precise
ways of describing the degree of precision of a number must be used. This exercise introduces two
such ways: significant figures and "plus-or-minus values".
Significant Figures
Once we realize that all numbers representing physical quantities are the result of measurements
having limited precision, we ought to express those numbers in a way that also conveys some idea of
their precision. This can be accomplished in a simple but approximate fashion by writing only the
significant figures (the significant digits).
The significant figures in a decimal number are those digits whose values are certain. For example,
the distance from Morris, MN to Washington, D.C. is about 1200 miles. In this number, the digit "1" is
significant--it is certainly "1" and not "3" or "5" or anything else. The digit "2" is also significant--it is
fairly certain. The two zeroes are not significant--they are not certain. (We do not really know what
those digits should be.) The two zeroes merely serve to locate the decimal point. Thus "1200" has
two significant figures.
In another example, a steel ball bearing is measured to have a diameter of 10.6 mm. Since the "1",
the "0", and the "6" are all certain, this number has three significant figures. This is a case where
zero is significant because the measurement indicates that it certainly is a zero and not a one or a
three or anything else. Now suppose we write the same number in a different form: 1.06 cm. Of
course, it still has three significant figures. Rewrite the number again: .0106 m. Since we already
know the number has three significant figures, we conclude that the leftmost zero is not significant. It
serves only to hold the decimal place.
As another example, consider the speed with which light travels through space: 300,000,000 meters
per second. How many significant digits does this number have? Checking a physics textbook, one
finds that the answer is three. The three and the two leftmost zeroes are significant while the other
six zeroes are only place-holders. This example illustrates how zeroes can be a problem--significant
zeroes and zeroes that are not significant look exactly alike!
To deal with this problem, a set of rules have been devised. These rules describe the proper way to
interpret a number so that the significant figures are unambiguous. Those rules are:
(1)

Non-zero digits are always significant.

86

Appendix C: Significant Figures

(2)

Zeroes that are located between non-zero digits are always significant. (e.g. 10.6 has
three significant figures.)

(3)

Zeroes to the left of all non-zero digits are not significant. (e.g. .00314 has three
significant figures.)

(4)

Zeroes to the right of all non-zero digits are significant if there is an explicitly written
decimal point. (e.g. 40.00 has four significant figures.)

(5)

Zeroes to the right of all non-zero digits are not significant if no decimal point is
explicitly written. (e.g. 1200 has two significant figures.) Exceptions to this rule can be
made by underlining the rightmost significant figures.

While these rules may seem complicated at first, they will seem more natural to you if you think about
them, understand why they are necessary, and practice using them. If you can recognize that all five
rules are derived from the question, "Is it a reliably known digit or is it merely a place-holder?", you
will understand much. Actually, the first three rules are rather obvious, while the fourth and fifth rules
are mainly concerned with defining proper usage of the decimal point and the underline.
Of course if we are to use these rules to interpret numbers, we must also write numbers in
accordance with the rules. (Numbers not written with these rules in mind cannot be correctly
interpreted by them.) For example, consider the first rule. To be consistent with this rule, we must
write non-zero digits only when they are significant. The temptation exists, particularly among
students with electronic calculators, to write non-zero digits that are not significant. Suppose you
wanted to convert 88.0 feet into meters. Dividing by 3.28 ft/m, a calculator gives 26.829268 m. Can
all these digits be significant? No. The original length is accurate only to the nearest tenth of a foot.
Converting units does not improve the precision of the number. It is therefore erroneous to write so
many digits. Your calculator does not understand significant figures--a bit of human thought is
required here. Incidentally, the correct conversion of 88.0 feet into meters is 26.8 m.
Another way of expressing numbers that you must become familiar with is scientific notation,
sometimes called "powers-of-ten notation". In this notation, each number is written as a number
between one and ten which is multiplied by a power of ten. Not only does this allow very large and
very small numbers to be written concisely but it allows numbers to be written entirely without placeholding zeroes. Thus a number properly written in scientific notation has only significant digits. Digits
that are not significant are not written.
As long as you are a student of physics you will be expected to write numbers properly so as to
convey, not only the number itself, but also its degree of precision. You may write numbers either in
decimal form (adhering to the five rules of significant figures) or in scientific notation. Whichever you
favor, you must understand both. The following table contains a list of properly-written numbers. The
second column shows the number of significant figures in each number. The third column shows
each number as written in scientific notation. Study these examples until every one is clearly
understood.
Number
1200 miles
1200. miles
1.50 kg
.0020 m
5.0020 m
34 km
340 sec
3400 miles

SF
2
4
3
2
5
2
2
2

Scientific Notation
1.2 x 103 miles
1.200 x 103 miles
1.50 kg
2.0 x 10-3 m
5.0020 m
3.4 x 101 km
3.4 x 102 sec
3.4 x 103 miles

87

Appendix C: Significant Figures


3400 miles
1605 yards
75.44 sec
705 cm
.06 m
.006 m
4.050 kg

3
4
4
3
1
1
4

3.40 x 103 miles


1.605 x 103 yards
7.544 x 101 sec
7.05 x 102 cm
6 x 10-2 m
6 x 10-3 m
4.050 kg

As soon as one begins to calculate using numbers of limited precision as input, one is faced with the
problem of determining the precision of the result of that calculation. Clearly the limited precision of
the result is caused by the limited precision of the input. The process by which this occurs is called
"the propagation of errors". A fairly rigorous mathematical analysis of this process is the subject of
Part 2 of this exercise. But often a simpler, less precise procedure is adequate--a procedure that is
summarized by two rules:
(1)

The result of multiplying or dividing numbers that represent physical measurements will have
the same number of SF (significant figures) as the measurement having the least number of
SF.

(2)

To add or subtract numbers representing physical measurements, write the numbers in a


column with decimal points aligned. The least (i.e. rightmost) significant digit of the result
occurs in the same column as the leftmost of the least significant digits of the
measurements.

These rules may seem complicated but the ideas they represent are simple. Two illustrations will
help:
Illustration of Rule #1
Multiply 3.2 by 1.80. Notice that one factor has two SF while the other has three.
According to Rule #1, we expect the product to have only two SF. To see why this is so, let
us write the factors as "3.2?" and "1.80?". The question marks denote the first unknown
digits. Perform the multiplication by hand treating the question marks like digits. Logically,
we consider "?" times anything to be "?".
1.80?
3.2?
.0????
.360?
5.40?
5.8????
(We have rounded the second digit up to "8" because the third digit was "6+?"). Thus the
product has two SF as we expected. Furthermore, this little game with the question marks
shows why Rule #1 works.
Illustration of Rule #2
Add 3.2, 1.80, and 14.225. Again, to monitor the propagation of uncertainty, let us write
these numbers as "3.2?", "1.80?", and "14.225?". Then add:
3.2?
1.80?
14.225?

88

Appendix C: Significant Figures


19.2???
The rightmost SF of the sum is the tenths digit because that is as far to the right as every
number being added is known. Observing how the uncertain digits line up relative to the
decimal point is what Rule #2 is all about.
Illustration Combining Rules
Computations often involve combinations of multiplication and division with addition and
subtraction. Let's look at a sample. Subtract 3438 from 3519 and divide the difference by
9.377. Notice that each number has four SF. In proper mathematical notation, we write
(3519-3438) 9.377. First we compute the subtraction:
3519-3438 = 81
which has only two SF according to Rule #2.
Then
81 9.377 = 8.6
which has only two SF according to Rule#1.
Thus
(3519-3438) 9.377 = 8.6
Study the following examples thoroughly:
76.6 + 84 = 161
37.44 6.36 = 5.89
76.3 + 6.4 + 12.4 + 23.2 = 118.3
3.14159 x 6.34 x 19 = 380
7.4 16.23 x 3.44 = 1.6
14.7 + 6.9 + 27 - 4.63 = 44
3.7 x .034 187.3 x 4470 = 3.0
6.44 x 5.23 - 3.67 x 4.84 = 15.9
(97.6 - 83.2) 38.4 = .375
100. - 14.3 x 6.87 = 2
It will often be necessary for you to compute from equations involving mathematical constants such
as integers, functions of integers, or pi. Since these are precisely defined numbers, they must be
considered to have an infinite number of SF.
In essence:
2 = 2.0000000 . . .
2/3 = .6666666 . . .
ln 4 = 1.38629436 . . .
= 3.1415926535 . . .
Then for a circle of radius r = 32.47 cm, the diameter is d = 2r = 64.94 cm, the circumference is C =
2r = 204.0 cm and the area is A = r2 = 3312 cm2. (In accordance with the rules, each of these
quantities has four SF because "r" has four SF.)
You are expected to utilize care with significant figures in all of your work, including exams and labs.

89

Appendix E: slopes

Appendix D: Uncertainty
The word "error" is used in many different ways. In non-technical usage, it usually indicates a
mistake has been made. In scientific usage, however, it does not necessarily carry that same
meaning. If we use the word error, we do not mean mistake, but actually the uncertainty in a
measurement value. In the following, the words error and uncertainty are interchangeable.
When one measures (or calculates from experimental measurements) a quantity, X, for which
there is a generally accepted value, the difference between the experimental value and the
accepted value is sometimes called "the error". We shall call this difference the "deviation from
accepted value"
dev = Xexpt - Xacc
When this deviation is expressed as a percentage of the accepted value, we shall call it the
"percent deviation from accepted value"

%dev

Xexp t Xacc
x100%
X acc

As an example, consider an experiment in which the gravitational acceleration, g, is measured to


be 9.2 m/s2. Since the accepted value is 9.8 m/s2, the deviation from accepted is -.6 m/s2 while
the percent deviation is -6% Thus the experimental value is 6% low.
Students often consider the deviation from accepted value to be the only indicator of a well-done
experiment (namely that a small deviation means good work while a large deviation means poor
work). While there is some truth to this, it oversimplifies a complex situation.
The deviation is not equal to the uncertainty or error of the experiment!
Experimental investigations are, by their nature, of limited precision. The methods and
instruments used all contribute various degrees of imprecision. More expensive equipment or
more careful methods might improve the precision of the experiment but the need for greater
precision may not justify the increase in costs (both time and money).
The most important question to be asked about an experiment is not "What is the accuracy of the
result? (How close is the result to the true value?)" but rather "What is the degree of precision of
this result?" While these two questions might seem the same, they are not. (This confusion arises
from incorrectly thinking of the deviation from accepted as the measure of precision.) As an
illustration, let us return to the experiment in which "g", is measured. Suppose four students each
perform the experiment.
Student A finds
Student B finds
Student C finds
Student D finds

g = 9.2 .8 m/s2
g = 9.2 2.0 m/s2
g = 9.5 .9 m/s2
g = 9.2 .2 m/s2

Each student has assigned an "uncertainty" or "plus-or-minus" value to his or her result. (Those
are carefully calculated numbers, not guesses. The goal of the remainder of this exercise is to
show you how to determine uncertainties.) For example, when A finds g = 9.2 .8 m/s 2, she
means that she has determined that the true value of g lies somewhere in the range of 8.4 m/s 2
to 10.0 m/s2. Her equipment and experimental method do not allow her to be more accurate.

90

Appendix E: slopes
Since the accepted value (g = 9.8 m/s2) does lie within A's range, we conclude that she has done
well. Next, Student B finds g = 9.2 2.0 m/s2. Notice that B gives the same value as A but a
much larger uncertainty. Thus A and B have the same deviation from accepted (namely -.6 m/s2)
but A's result is more precise. Evidently student B had poorer equipment or a poorer experimental
technique. Student C's result is g = 9.5 .9 m/s2. While this is closer to the accepted value than
A's result, the uncertainty is larger. A's work is more precise (because it defines a narrower
range for g). Still, C has done good work also. Finally, Student D claims to have obtained g = 9.2
.2 m/s2. By writing such a small uncertainty, D is claiming to have done the experiment much
more precisely than others. Unfortunately, the accepted value lies well outside the range 9.0
m/s2 to 9.4 m/s2 that D proposes. Something is seriously amiss and we reject D's experiment as
faulty. (D's grade will reflect this.) In summary, after rejecting student D's result, we conclude
that Students A and C have done the best work and Student B has probably done OK.
The methods and instruments used in an experiment always contribute something to the
imprecision of the result. The uncertainty can never be reduced completely to zero--the
experimenter must be content with keeping it reasonably small.
Uncertainty in a measurement arises mainly from three sources:
1. Scale-limited uncertainty or Precision error: measurements use scales to read values. The
nature of the scale limits the number of significant figures you can confidently measure.
For example, the odometer of your car will allow you to measure the length of a trip to a
precision of 1 mile, but not 1.239 miles. The use of more precise scales in
instrumentation can minimize this error. For estimation of precision error, see below.
2. "Systematic errors" or Accuracy errors are those that arise when a measuring device or
method is flawed. Some examples are: a clock that runs too fast, a meter stick whose
end is worn down, a voltmeter whose needle is not properly zeroed, or a tape measure
with a kink in it. A well-designed and well-performed experiment eliminates this class of
errors as much as possible.
3. "Random errors" constitute the third class of errors that occur in an experiment. These
are present in every measurement. Random errors are usually due primarily to the
limitations of the measuring device ("scale-limited") though sometimes they are also
inherent in the quantity being measured. Examples of scale-limited random errors are:
(1)
When measuring a time interval (t) with the second-hand of a watch (which is
marked off in seconds), the interval can be measured only to the nearest second (or
perhaps, half-second). Thus the measurement is of limited precision because of the
scale. The imprecision or uncertainty can be characterized by assigning a t of .5
seconds. The reading is then: t .5 s.
(2)
When measuring a length (L) with a meter stick (which is marked off in
millimeters), the length can be measured only to the nearest millimeter (or perhaps, halfmillimeter). The uncertainty of such a measurement might be about L = .5 mm. The
reading would then be: L .5 mm.
An example of a random error inherent in the quantity being measured is: When
measuring the thickness of a brick with a meter stick, it is noticed that the thickness
varies over the length of the brick.
Random errors are assessed and minimized by repetition and averaging. We will learn
methods of estimating the random error in later experiments.
The uncertainty of a measurement is then:
X = X(precision) + X (accuracy) + X (random)

91

Appendix E: slopes
Another class of errors are "blunders". Blunders are those mistakes that occur due to the
experimenter's negligence or ignorance or due to accident. Examples of blunders are: misreading
a stopwatch, over tightening a micrometer, miscounting the oscillations of a pendulum, or writing
"18.3 cm" when "13.8 cm" was intended. Blunders ought not to occur. If they do, the affected
portion of the experiment must be redone.
The hallmark of a quality experiment is a small uncertainty in the result. If the uncertainty
(X) is determined correctly and honestly, it ought to be the case that the accepted value (X acc)
lies within the range defined by the experimental value (Xexpt) plus-or-minus the uncertainty.
Xexpt - X < Xacc < Xexpt + X
X

Xexpt

Xacc

Then the difference between the experimental and accepted values, which is the deviation from
accepted, must be less than the uncertainty.
|dev| < X
If one strives to reduce the uncertainty, the deviation from accepted is naturally reduced as well.
Uncertainties are written as plus-or-minus values immediately after the number whose
imprecision they describe. The uncertainty in a measurement of the quantity X is denoted X.
Then the measurement is properly written X X. X is sometimes called the "absolute
uncertainty". The absolute uncertainty has the unit of the measurement value. The relative
uncertainty of X is X/X and is usually expressed as a percentage. This is often more useful
than the absolute uncertainty. Consider two examples that each have the same absolute
uncertainty.
(1)

Suppose the thickness of a chemistry textbook is measured to be 4 1 cm. The


relative uncertainty is 1/4 or 25%. That is a poor measurement.

(2)

Suppose the height of the science building was measured to be 1842 1 cm. Now
the relative uncertainty is 1/1842 or .05%. That is an outstanding degree of
precision.

This shows how the relative uncertainty is sometimes a better indicator of good work than the
absolute uncertainty.
Measurements:

92

Appendix E: slopes
A: 6.4 .05 cm
B: 6.6 .05 cm
C: 6.5 .05 cm
Thus, although each measurement is accurate to .05 cm
(because of scale-limited error), the thickness of the brick
cannot be defined that precisely. A reasonable statement
would seem to be: thickness = 6.5 .15 cm. This range
includes all the possible thicknesses indicated by
measurements A, B, and C. Notice that both kinds of random
error are at work here.
The meter stick imposes an
uncertainty by virtue of its scale, namely .05 cm, while the
brick itself imposes an additional uncertainty of .1 cm due to
non-uniformity.
Estimating Scale-Limited Uncertainties (precision error)
Every time a measurement is made, the uncertainty must be estimated. Once systematic errors
and blunders are eliminated, only random errors need be considered. In most cases, the error
inherent in the quantity being measured is less than the scale-limited error. When that is so, the
uncertainty is determined by the scale-limited error alone.
To estimate scale-limited uncertainty, the experimenter must ask himself or herself the question
"Given the scale on this measuring device, how accurately can I take a reading?" If, for example,
the device is a meter stick, the experimenter might be able to read to the nearest half-millimeter.
The uncertainty is then half of that or 1/4 mm. Another experimenter, however, might be less
confident in his or her ability to estimate a reading that falls between two of the black lines on the
meter stick. Perhaps readings are then considered accurate only to the nearest millimeter. The
uncertainty is then 1/2 mm. Either of these estimates is acceptable as long as it is honestly
made.
The scale-limited uncertainty of a particular device is never greater than one-half of the smallest
subdivision or "least count" of its scale. For an experimenter with a sharp eye, some devices
allow estimates of a measurement accurate to one-tenth of the least count while other devices do
not allow any estimating beyond the least count.
In particular:
For a meter stick with least count of 1 mm
estimates to a fraction of a mm may be possible.
For a Vernier caliper with least count of .02 mm
further estimating is unlikely.
For a micrometer caliper with least count of .01 mm
estimates to .001 mm may be possible.
In each case, the uncertainty is no less than one-half of the least amount that can be estimated.
Estimating Inherent Variabilities
Sometimes the quantity being measured has inherent variability greater than the least count of
the measuring device. For example, the height above the floor of a tabletop may be uncertain
because of unevenness of the table or roundness of the tables edge. In another example,
because of friction in the pivot, a seesaw may be balanced by any weight within some range, not
just by one specific value.

93

Appendix E: slopes

Every effort should be made to account for both the scale-limited uncertainty and the inherent
variability when assigning uncertainties.
A simple illustration:

Precise and accurate measurement

Precise, but not accurate, measurement

Not precise, but accurate, measurement

Repetition can minimize the influence of random errors

94

Appendix E: Propagation of Uncertainty

Appendix E: Propagation of Uncertainties


As was already discussed and illustrated in Part 1 of this exercise, any calculation having numbers of
limited accuracy as input will yield a result of limited accuracy. We also examined a crude method for
determining the accuracy of that result from the accuracy of the input numbers. Now we are ready for
a more rigorous method.
An "error equation" stipulates how the uncertainly in a calculated quantity is related to the uncertainty
in the input values.
The general method of deriving error equations is based upon differential calculus. Consider, first, a
function of just one variable: f(x). Now suppose we wish to evaluate this function for a certain value of
will be able to estimate.
Recall the definition of a derivative:

df
dx

df
dx

or

f x

lim
x

fx

x f x
x

f x

x f x
x
fx

df
x
dx

But the left

df
x
dx
Graphically, one can visualize this as follows: Knowing the local slope of f(x) allows to estimate the
variation in F if x has a slight variation, see figure.
f

f(x)

y
f+ f
f
f- f

x- x

95

x+ x

Appendix E: Propagation of Uncertainty


The generalization to functions of many variables is straightforward. One adds together the
uncertainties in the function due to the uncertainty of each of the independent variables. Consider a
function f(a,b,z). Then
f
f
f
f
a
b
z
a
b
z
These derivatives, written with a rather than a d, are called partial derivatives. This notation is
used when a function of several variables is to be differentiated with respect to one variable while the
others are considered constant.
2

For example, if f=a /b+3z, then


f
a

2a
,
b

a2

f
b

, and

f
z

The determinate error equation for f is then

2a
b

a2
b2

3 z.

e positive or negative.
When one is dealing with uncertainties, it is incorrect to use the determinate error equation. Instead,
one must use the indeterminate error equation that is obtained from the determinate one by taking
the absolute value of the quantity multiplying each uncertainty. In general form, this is written:

f
a
a

f
b
b

f
z
z

Example #1
Find the equivalent resistance of three resistors R 1=165

5%, R2=274

3=312

For resistors in series: R = R1 + R2 + R3


= 165

+274

Then consider R as a function of R1, R2, and R3, take the derivatives, and calculate the
uncertainty (indeterminate error) in R:

R
R1
R1

R
R2
R2

R
R3
R3

= 1.R1 + 1.R2 + 1.R3


= (.05)165
= 8.25

. 274
+ 13.7

+ 15.6
37.55

96

. 312

Appendix E: Propagation of Uncertainty

Given the size of the uncertainty, the ones digit of the resistance value is completely
uncertain and should not be written. Also, since the initial uncertainties had only one
significant digit, the rules of Part 1 suggest that the final uncertainty also has only one digit.
Thus the result would be better written
40
where 40

is called the "absolute uncertainty", R, and has appropriate units

or
5%
where 5% is called the "relative uncertainty", R/R, and is unitless.
Important note!
In this class relative uncertainties should always be reported as
percentages with a % sign. Absolute uncertainties are always reported with the appropriate
units. Uncertainties without either units or a % sign are ambiguous and hence incorrect.
Example #2
Find the density of a rectangular block of length 6.3 .05 cm, width 4.2 .05 cm, height 2.7
.05 cm, and mass 143 .5 g.
mass
volume

Density:

M
LWH

2.002

g
cm3

Then

d
M
M

M
M
LWH M

d
L
L

d
W
W

M
L
LWH L

d
H
H

M
W
LWH W

M
H
LWH H

After some additional algebra, we find the simple-looking expression

d
d

.5g
143g

M
M

L
L

.05cm
6.3cm

W
W

.05cm
4.2cm

H
H

.05cm
2.7cm

=.04 = 4%
Considering the size of this uncertainty, the value of d should be written
d = 2.00 g/cm3 4%
= 2.00 g/cm3 .08 g/cm3
(Notice that d turns out to have approximately two significant figures, consistent with the
rules in Part 1.)

97

Appendix E: Propagation of Uncertainty


A short-cut derivation of error equations that makes use of differentials will sometimes be helpful. A
differential is an infinitesimal change in a variable. For example, the differential of x is dx. (A
derivative is the ratio of two differentials such as df/dx.)
The differential of

x
4x
3
x
x
e
lnx

is
is
is
is
is

dx
4dx
2
3x dx
x
e dx
dx/x

For a function f(x,y,z),


f
dx
x

df

f
dy
y

f
dz .
z

If we approximate df by f, dx by x, etc. we will have rederived the determinant error equation. To


get to the indeterminate error equation, take the absolute values of the factors multiplying the x, y,
and z.
A slight variant of this differential technique for computing error equations involves taking the
logarithm of the function before finding the differential:
ln(f)=ln(f(x,y,z))

df
f

ln f
dx
x

ln f
dy
y

ln f
dz .
z

This will sometimes yield a short-cut to the error equation. It is especially useful when the equation
being used contains mostly multiplications, divisions, and powers. For use as an error equation,
absolute values must be applied to the appropriate terms.

f
f

ln f
x

ln f
y

ln f
z

Example #3
An object falls freely from rest. Compute its displacement after a time 1.4 .1 seconds.
2
Use g = 9.806 0.0005 m/s .
Displacement: x

gt2
= 9.61 m
2

Because the expression for x has only multiplications, divisions, and powers, it will be most
efficient to use the log differential method for finding the error equation. First write
ln(x)=ln(g)+2ln(t)+ln(1/2)
The differential of this equation is

dx dg 2dt
=
+
0
x
g
t

98

Appendix E: Propagation of Uncertainty


Next, change infinitesimal differences to finite differences and take the absolute value of
each of the factors multiplying the different uncertainties. Then plug in values.
x
g 2 t .0005 m/ s2
0.1 s
=
+
=
= 0.14
2 +2
x
g
t
1.4 s
9.806 m/ s

Thus

x = 9.6 m 14%
= 9.6 m 1.4 m

(In this case, the rules of Part 1 suggest that the final uncertainty should have only one
significant digit. But because the leading digit of the uncertainty is "1", it is acceptable keep
an extra digit.)
Error-Equation Derivations in Special Cases
When only multiplication, division, addition, and subtraction are involved in an equation, the related
error equation can often (but not always) be derived with the careful application of two special rules:
Multiplication/Division/Powers Rule:
If

A .B .C
. .

D EF

then by the log-differential method

and

dX
X

dA
A

dB
B

dC
C

dD
D

dE
E

X
X

A
A

B
B

C
C

D
D

E
E

dF
F
F
.
F

The Multiplication/Division Rule can be stated:


For any number of multiplications and/or divisions, the relative uncertainty of the calculated
quantity is equal to the sum of the relative uncertainties of the terms being multiplied or
divided.
The density calculation in Example #2 could have been done with this rule.
Since

M
LWH

the rule immediately yields the correct expression:

d
d

M
M

L
L

W
W

H
.
H

This rule also handles powers, including fractional powers such as square roots, since a power is
essentially multiplication of a quantity by itself a certain number of times.
Powers Rule:

99

Appendix E: Propagation of Uncertainty

If

An

then by the log-differential rule


X
X

A
.
A

The Powers Rule can be stated:


For any power, the relative uncertainty of the calculated quantity is equal to the relative
uncertainties of the term being raised to a power multiplied by the absolute value of the
power.
Example #3 can be re-done using these two rules
For

gt 2
2

first apply the multiplication rule to the equation


x
x
x

Then

1
gX .
2
g
g

X
X

But, using the powers rule,we can determine that

X
X

So
x
x

g
t
2
.
g
t

Notice that constant mulitpliers, such as the 1/2 in the example above and the 2 in the example
below, do not appear in relative error equations because they have no uncertainty.
Example #5
If
t

then

t
t

2x
a

1 x
2 x

100

1/ 2 1/ 2

1 a
2 a

1/ 2

1 x
2 x

1 a
.
2 a

Appendix E: Propagation of Uncertainty


Addition/Subtraction Rule:
If
X=A-B+C-D-E+F
then by the differential method
dX = dA - dB + dC - dD - dE + dF
and
X = A + B + C + D + E + F.
The Addition/Subtraction Rule can be stated:
For any number of additions or subtraction, the absolute uncertainty of the calculated
quantity is equal to the sum of the absolute uncertainties of the terms being added or
subtracted.
The equivalent resistance calculation in Example #1 could have been done with this rule.
Since
R = R 1 + R2 + R 3
the rule immediately yields
R = R1 + R2 + R3.
Notice that this rule handles multiples of a variable too, since multiples are essentially the
addition of a quantity to itself a certain number of times.
For example, if
f = 4D = D + D + D + D
then
f = D + D + D + D = 4D.
These special rules can be used in combination but it must be done very carefully. The next example
illustrates this.
Example #5
Find v v where

xf

xi
t

and xi = 1.3 .05 m, xf = 22.3 .05 m, and t = 3.7 .1 s. Substituting values, we find v
= (22.3 m - 1.3 m) / 3.7 s = 5.7 m/s.
To get the uncertainty in v, we can view this calculation first as a division (the xf
quantity divided by t). Applying the multiplication/division rule gives
v
v

(x f xi )
xf x i

t
t

xi

Notice how the subtraction was treated as a single term with respect to the division and
absolute value signs had to be assigned to the difference in the denominator. Next,
apply the addition/subtraction rule to the numerator of the first term on the right side:

101

Appendix E: Propagation of Uncertainty

(xf - xi) = xf +xi


Substituting this back yields the final expression:

v
v

xf
(xf

xi
x i)

t
t

Putting in the values yields

v
v

.05m .05m
22.3m 1.3m

Then

v
v

.032

0.1s
.
3.7s

3.2%

Notice how the units all cancelled, as they must when calculating a relative uncertainty.
The final result can be written
v = 5.7 m/s 3%
Important note!
These special rules should only be applied to equations in which the variables
each occur only once. With experience, a person can figure out how to correctly use these rules on
equations that contain multiple occurrences of variables as well. Consult with your instructor if you are
unsure.

102

Appendix F: slopes

Appendix F: Finding the Slope and Slope of Linear Graphs


Graphs must be on graph paper and only one per page. Make the large so as to fill the page. If
the points form a straight line, or if you have theoretical reasons for expecting the data to form a
straight line, draw the best linear fit to the data.
Draw a large right triangle whose vertices intersect the best linear fit near the largest and smallest
data points but do not use actual data points for the vertices. (See the sample lab report for an
example.) Label those vertices with values for both the horizontal and vertical variables and
assign reasonable uncertainties to all those values. The uncertainties in the rise and run can be
determined by examining the values of the uncertainties of data points near the vertices of your
right triangle and by taking into consideration how well or poorly the line actually fits the data.
(You should look back at your data to select appropriate values for the uncertainties.)
Calculate the slope using:

slope =

rise
run

y2
x2

y1
,
x1

where y represents the vertical variable and x represents the horizontal variable.
Calculate the uncertainly of the slope using the error equation:

slope
slope

rise
rise

run
run

y2
y2

y1
y1

x2
x2

x1
x1

You should show, on the graph, your calculations of the slope and slope. But, if your computer
software does the linear fit and calculates the slope for you, you only need to show the slope
calculation. See Graph #2 of the sample lab report for an example of a slope calculation.

Appendix G: Excel tips

Appendix G: Excel Tips for Physics Labs


Recording Data
Change the format of cell data: Sometimes Excel changes the format of data against your will as you
enter it. Click FORMAT
CELLS, choose the number tab, and designate the format specifically. This also
allows you to set the displayed precision of numbers. (The precision can also be changed from the
toolbar by clicking the .00 buttons.) Highlight many cells simultaneously to format all the same.
Resize columns to fit the data: at the top of the screen in the row of letters labeling the columns, place
the cursor on the line separating two columns. When the cursor changes to a line with arrows pointing
left and right, double-click to automatically resize, or click-n-drag to manually resize. These and other
options are also available from the FORMAT
ROW or COLUMN menus.
Make the data fit horizontally on one printed page: Click file
print preview to check. If you are
close, adjust margins and/or close the preview and adjust column widths. Once you have previewed,
dotted lines in the spreadsheet show page boundaries. If necessary, change to landscape printing under
FILE
PAGE SETUP.
Include labels and notes with data: Describe data tables, label columns with names and units, etc. so
they make sense to other readers. To make cells that span more than one column and/or row, under the
FORMAT
CELLS choose the ALIGNMENT tab and select MERGE CELLS and/or WRAP TEXT.
Inserting extra rows and columns: If you need to add 6 extra rows between rows 9 and 10, click on
row label 10 to highlight the entire row, then drag down to highlight six rows (rows 10 through 15). Click
INSERT
ROWS and six empty rows will appear. New rows (or columns) will be re-numbered
automatically, and generally Excel will be able to adjust equations in other cells correctly, but always
check to be sure.
Manipulating Data
Calculations: To put the value of 7*3/2 into a cell, type =7*3/2 (without the quote marks) and hit
ENTER. (Format the cell to an appropriate precision!) Any calculations you want Excel to do should start
with =. (If you want the cell to actually display =7*3/2 rather than calculate it, first format the cell as
text. If you dont want to do that, you can trick Excel by starting with a space: =7*3/2.)
Calculations using cell values: Cells are defined by column letter and row number: K7. So =K7*3
will multiply the number in cell K7 by 3. You can type the cell name in by hand, or click on the cell you
want while typing the equation and Excel will insert the correct name at the cursor location. As before,
the calculation is performed when you hit ENTER.
Repeating calculations: Say column A contains a list of 10 time measurements ( t), column B
contains a list of 10 distance measurements ( x), and you want column C to contain the velocities
( x/ t). In the first row of column C (lets assume thats C1), type in the equation =B1/A1 and ENTER.
Click on C1 again and notice the cell has a thick border with a small box in the lower right corner. Place
your cursor over the box so that the cursor turns from an outlined + to a smaller solid +. Click and drag
to highlight the next 9 rows of column C. When you release, cells C2 through C10 should contain the
proper values for their rows. Notice that Excel has automatically changed the equation so that in row 2 it
says =B2/A2, in row 3 it says =B3/A3 and so on.
Using functions in calculations: To calculate the square root of the value in C1, type =sqrt(C1).
Always enclose the argument of a function in parentheses. You can browse through the available
functions by clicking on the fx button on the toolbar. Note that the number 3.14159 is a function, so
type =pi().
Preventing Excel from changing the cell names in repeated calculations: Say you wanted to
multiply each of the values in column C by the same factor, which you will record in cell K1, and put the
results in column D. If you type =C1*K1 in cell D1 and drag this equation down, cell D2 will contain
=C2*K2. To keep K1 from changing, write $K$1 instead. The equation =C1*$K$2 will update the
C values for each row, but always use the value in K1. Dollar signs must be inserted manually.

Appendix G: Excel tips


Graphing
Getting started: To open the chart wizard, click the toolbar button that looks like a colored bar graph.
Step 1 of 4: In this class, you will always use the XY scatter graph without any connecting lines.
(Not the line graph!)
Step 2 of 4: Choose the SERIES tab and ADD a series. It may be helpful to type an informative name in
the top white box. For the x-values, click on the red-white-and-blue box to the right of the middle white
box. The window will change. Click and drag to select the cells containing the x-values. The Excel
description will appear in the chart wizard window. Click the funny little box at the right of the chart
wizard window to return to the Step 2 window. Repeat for the y-values. If you want to plot multiple
datasets on the same graph, ADD another series and repeat.
Step 3 of 4: Add titles, axis labels (units are good!), remove gridlines, remove the legend unless it
conveys useful information and generally clean up the appearance of the graph.
Step 4 of 4: Place chart as object in the current sheet and FINISH. (This should be the default.)
Tidy things up: Right-click in the gray area and FORMAT PLOT AREA. Change AREA to NONE to cut back
on unnecessary use of ink. Most other features can also be adjusted in this way. (E.g., right-click on
axes to change their ranges.) Click in the graph to highlight the outermost border and drag the graph
to the desired location on the spreadsheet.
Adding error bars to graphed data points: Right-click on a data point and FORMAT DATA SERIES.
Under the tabs for X ERROR BARS or Y ERROR BARS, chose the appropriate options. Generally you will
want the BOTH display option. If all the points have the same uncertainty you can input the value under
FIXED VALUE. If each point has a different calculated uncertainty, choose CUSTOM and input the columns
containing the uncertainties in the same way you input the data columns. (Generally, the + and values
will be the same.) When the error bars are small, decrease the data marker SIZE and/or the line WEIGHT
from the same window under the PATTERNS tab.
Adding a trendline to graphed data: Warning do not add a trendline if the data do not look linear!
Right-click on a data point or pull down the CHART menu and select ADD TRENDLINE. You will always use
a LINEAR trendline in this course. Click the OPTIONS tab. If you have a good reason for believing the
intercept must be zero, then select SET INTERCEPT =0; otherwise, dont. Always choose DISPLAY EQUATION
ON CHART. When youre done, move the equation around on the graph to make the graph more legible.
Printing: To print just the graph filling a full page (often a good idea), click in the graph to select it then
preview and print (or just print). To print the graph on the same page as the relevant data table, make space
for the graph by adding rows if necessary and move the graph into place. Highlight the data and the cells
under the graph. Choose FILE
PRINT AREA
SET PRINT AREA, then preview and print. Remember to
choose FILE
PRINT AREA
CLEAR PRINT AREA when you are done. (Unfortunately, there is no printer in the
physics lab.)
Save Your Work! Save to a file with your name in it on the desktop early and often!!! At the end of
lab, email the file(s) to yourself and your lab partners. Files left on the desktop may be overwritten by
other groups and will be irregularly cleaned off without warning.

You might also like