Sie sind auf Seite 1von 86

LABORATORY EXPERIMENTS

FOR

PHYSICS 204C:

Sound, Heat, Light, Atomic and Nuclear Physics


SPRING 2011

Sections 1-3

Dr. Eric Dietz

Drs. C.K. Chau and R. Paulson


Department of Physics
California State University, Chico

© 2011 Eric R. Dietz


LAB TIPS ....................................................... 4

Standing Waves on a String ..................................... 6

Sound .......................................................... 9

Temperature, Heat Flow and Thermal Expansion .................. 12

The Ideal Gas Law ............................................. 16

Specific Heat and Heat of Fusion for H2O ...................... 19

Specific Heat Ratio for Gases: Rüchardt’s Method .............. 23

Geometrical Optics I: Reflection and Refraction of


Light .................................................... 27

Geometrical Optics II: Image Formation with Mirrors


and Lenses ............................................... 31

Physical Optics I: Microwave Interference and


Diffraction .............................................. 34

Physical Optics II: Interference and Diffraction of


Light .................................................... 38

Physical Optics III: Intensity and Polarization of


Light .................................................... 41

Atomic Spectroscopy and the Bohr Model ........................ 45

Matter Waves: Electron Diffraction ............................ 49

The Half-life of Radioactive Indium 116 ....................... 53

©2011 Eric R. Dietz


Appendix A: Measurements and Uncertainties .................... 56

Appendix B: Meter and Oscilloscope Uncertainties .............. 67

Appendix C: Reading a Vernier Scale ........................... 69

Appendix D: Vernier Hardware Spec Sheets ...................... 72

©2011 Eric R. Dietz


LAB TIPS
There are some important issues to keep in mind as you do these lab experiments and
record your observations. Following these suggestions is practically guaranteed to substantially
increase both your lab scores and your enjoyment of the experience.

1. Read the write-up of the experiment before coming to lab and try to visualize yourself doing
the experiment. Such preparation will help you and your group perform the measurements more
efficiently than coming in “cold.” This will also allow you to pace yourself during the lab, so that
you will finish, hopefully with time to spare. Be an asset to your team, not a liability.

2. Entries in the lab notebook should be in pencil; feel free to erase. Writing should be large and
dark enough to be read easily. Each lab should have a title at the top of the first page of the
entry. This should be followed by a brief description of the experiment: what will you do and
how will you do it? Write using complete sentences, so that someone who wasn’t there during
the experiment would be able to read your notebook and have a good idea of what happened.

In particular, avoid at all costs starting sentences (answers to questions) with the words “yes” or
“no.” Remember that your work will be read by others, so NEATNESS COUNTS!

3. UNITS: Don’t forget to indicate the units (where appropriate) on all numerical values. If these
are recorded in a chart, you can indicate the units at the top of a column. By far the most
common deficiency in lab entries is failure to give units the proper respect!

4. Show sample calculations to indicate how data is being combined to produce a calculated
result, and how uncertainties are being estimated. Results should be reported in a form (i.e. with
a number of significant figures) consistent with its uncertainty. Please note that it is almost never
appropriate to report an uncertainty with more than one significant figure!

 Always make a clear distinction between a calculated, accepted, or theoretical result and
a quantity that you measured to which it is being compared. This is especially easy to do
if your results are summarized in chart form.

5. GRAPHS: Graphed data should take up at least half a page if feasible. If you are graphing a
few data points (as opposed to the continuously recorded output of one of the sensors), make sure
to plot the data points clearly and indicate the curve that best fits them. Don’t forget to label the
axes with the quantity being plotted (e.g., “time,” “voltage,” etc.) along with the units. Always
give the graph a title (nothing fancy—just say what’s being plotted). When comparing
predictions and observations, you can use the same set of axes for both graphs—just be sure to
distinguish the two plots (with, e.g., solid and dashed lines, or with different colors). Please do
not put arrowheads on the ends of best-fit lines!

6. If your observations or calculations give results that are so far from those that you expect that
something must be wrong and you and your group can’t figure it out, ask the instructor for help.

7. Always write a conclusion in which you answer the indicated questions at the end of the write-
up. Try to base your statements on what you and your partners actually observed.

8. Always review graded labs for comments intended as feedback. Don’t make the same mistake
twice!

© 2011 Eric R. Dietz


9. Courtesy: As a courtesy to your group, to the class, and to the instructor, don’t come in late to
lab. Do not write on the tabletop. Backpacks must NOT be placed in the aisles (stow them under
the table or by the entrance). Before you leave: Clean up any cups, wrappers, etc. from your
work area as a courtesy to the next group and to the lab technician. Put everything back the way
it was when you started. Report broken apparatus immediately. The cell phone policy stated in
the syllabus applies in lab as well.

5
Standing Waves on a String

Introduction. By exciting a wire under tension that is fixed at both ends, a


standing wave pattern can be produced if the distance L between the ends corresponds to
an integer number of half-wavelengths. In the situation shown below, two
half-wavelengths comprise the standing wave pattern:

The wavelengths corresponding to the nth standing wave are determined by requiring that
there are an integer number (n) of half-wavelengths between the fixed points:
 
n  n   L . These wavelengths are always related to the frequencies fn and the wave
 2 
speed vs by the general kinematic relationship
vs  f n  n .
For a wire with linear mass density  that is maintained under tension T, the wave speed
is shown from Newton’s 2nd law to be
T
vs  .

In this experiment, you will determine if/how the wave speed depends on the excitation
frequency for constant tension. Then you will investigate how the wave speed depends
on the tension.

Procedure. You will need the following:

 Computer/Logger Pro  mass set


 clothed Cu wire (0.479 g/m)  digital oscilloscope
 permanent magnet  impedance matching transformer
 2-meter stick  audio oscillator
 pulleys and clamps

6
Standing Waves on a String Physics 204C Laboratory

Configure the apparatus as indicated schematically below:

Make sure that the ground of the scope is connected to the ground on the output of the
generator; the generator terminal not connected to ground should be connected to the
ch. 1 input of the scope and also to the other side of the matching transformer. Be sure
that the clothed wire is not snagged or otherwise encumbered on the apparatus so that the
tension in the horizontal portion between the pulleys is Mg.

When the signal generator is on, an alternating current flows through the clothed wire
under tension, and the magnetic field then excites the wire with an oscillating vertical
force FB.

Observe the standing wave patterns: With the generator on, adjust the frequency dial until
a standing wave pattern is observed on the wire. With the amplitude dial on the generator
set fairly low, you should try to maximize the response for each pattern by positioning the
magnet to coincide with an antinode of the pattern and carefully moving the fine
adjustment on the frequency dial. Read the frequency directly from the scope by pushing
the gray Time button and making the appropriate menu selection with the softkeys at the
bottom of the screen.

I. Measure the wave speed for different frequencies. With a 200-gram load1 producing a
fixed tension on the wire, adjust the frequency and observe the first four standing wave
patterns (harmonics). For each standing wave, sketch the pattern, measure the frequency
and the wavelength, and determine the wave speed vs  vs . Tabulate these values of vs
along with the uncertainties (see below).

According to your results (and considering the uncertainties), does the wave speed for the
wire depend on the frequency?

1
If this load results in a fundamental frequency that is too low to be produced by your signal generator, try
adding more mass.

7
Standing Waves on a String Physics 204C Laboratory

II. Measure the wave speed for different tensions. For four different tensions T (not to
exceed the weight of 600 g), measure the frequencies corresponding to the same standing
wave pattern. For each value of T, compute the value of the wave speed vs and tabulate
your results.

Plot2 the square of the wave speed vs. the tension in Logger Pro and determine the slope
of the best-fit line along with the uncertainty3. Sketch the resulting fit to your data. Is the
relationship linear (and why should it be)? Is the slope consistent (within uncertainties)
with the measured linear mass density of the wire, which is given4 as
(0.479 ± 0.005) g/m?

Uncertainties. The uncertainty in the measurement of a standing wave frequency (f) is


likely to have less to do with the accuracy of the digital scope display and more to do
with identifying the frequency for which you observe a maximum response on the wire.
For the wavelength, the uncertainty () will most likely have to do with the details of
your observation, rather than half the smallest division on your measuring stick.

Since the wave speed vs is the product of the two aforementioned uncertain quantities,
Eq. (8b) of Appendix A suggests its fractional uncertainty can be estimated as
2 2
vs  f    
  n   n  .
vs  fn   n 
Conclusion. Briefly summarize what you learned today about waves on a wire. In
particular:

What factors, according to your observations, influence the speed of a wave on a wire or
string under tension? How do your observations support your answer?

2
To plot a graph, always bring up Logger Pro from the desktop shortcut icon, not from the program menu.
3
To obtain the uncertainty in the best-fit slope in Logger Pro it may be necessary to double-click the box
with the slope value and check the appropriate box in the dialog that appears.
4
If the spool from which the wire was taken is available at the front bench, use the value of  printed on it
instead of the value given here.

8
Sound Physics 204C Laboratory

Sound

Introduction. A sound wave is a propagating longitudinal disturbance in an


elastic medium that can be described either in terms of pressure variations or molecular
displacements of the medium. The microphone we use in this lab is a transducer that
translates the pressure variations in a sound wave into voltage signals that can be
analyzed with Logger Pro.

The speed of sound in a medium depends on the ratio of its bulk modulus (“springiness”)
to its density. For air, a good approximation for this speed at temperatures near 0C is
given by
vs  331.5 m/s  ( 0.607 m/(sC) ) T, (1)
where T is the temperature in C.

In this lab you will determine the speed of sound by measuring the time it takes to travel
a known distance. You will also examine the waveforms and frequency characteristics of
sounds produced by various sources.

Procedure. You will need the following:

 computer  LabPro  meter stick


 Logger Pro  wood block
 microphone  2 steel balls
 temperature probe  2 tuning forks (same f , one on box)
 PVC tube  musical instrument.

I. Speed of sound in air: Connect the LabPro to the computer and the microphone5 to
Ch. 1 of the LabPro, and then bring up the Speed of Sound in Air 1 (.cmbl) file from the
“Dietz’s 204C LABS” folder on the desktop6. Block one end of the PVC tube and put the
microphone at the other end. Press “Collect” and make a sharp sound7 at the
microphone; this will trigger collection of a 10 ms segment of the waveform, on which
you should be able to identify the original sound pulse and its echo. Sketch this
waveform and use the AnalyzeExamine feature from the menu to measure the time
delay t  t between corresponding points on the pulse and its echo.

From this round trip time t and a measurement of the tube length L  L, deduce the
speed of sound vs along with its uncertainty. You should repeat this a few times and take
an average for t; you can estimate t from the standard deviation in these times.

5
For more information about the microphone and the rest of the sensors, see Appendix D.
6
Always be sure that the sensor and LabPro connections are complete and that the LabPro is powered-up
before bringing up any of the Logger Pro files.
7
The sound can be generated by clinking two metal balls together, or alternately (but less reproducibly)
with a finger snap. The pulse must be of short enough duration so that it does not overlap its echo.

© 2011 Eric R. Dietz


Sound Physics 204C Laboratory

For comparison, calculate the speed of sound (with its uncertainty) from the expression
for vs in Eq. (1). You will need the temperature of the air, which you can measure by
connecting the temperature probe (disconnect the mic) to the LabPro and bringing up the
Speed of Sound in Air 2 file.

How does your measured speed of sound compare with that given by the expression in
Eq. (1)?

II. Sound waveforms: With only the microphone connected to the LabPro, bring up the
Waveforms of Sound Waves file.

Qualitative: Examine the waveform produced by various sources such as your voice,
clapping hands, and both tuning forks8. Sketch one of these waveforms.

Do the waveforms for the two tuning forks appear to be different? Can you explain?

Quantitative: Capture a waveform from the box-mounted tuning fork and determine the
frequency of the wave (1) by counting cycles on the graph and (2) by fitting a sinusoidal
curve to the waveform9. How do your values of the frequency compare?

III. Frequency characteristics of sound: In this part you will measure the frequency
components of a few different waveforms by means of a technique known as the Fast
Fourier Transform (FFT). Bring up the Frequency Distribution for Sound Waves file.

1. Capture a sound waveform from the box-mounted tuning fork and sketch the FFT
graph. Do the same for the other fork. Does either fork produce a single frequency?
What appears to be the effect of the box?

2. Capture and sketch the FFT graph that results from one note of a musical instrument.
Identify the fundamental and higher harmonic frequencies.

Uncertainties.

I. For the measurement of the speed of sound, you will need to estimate t from either the
appearance of the pulses or from the standard deviation in the measured transit time
(whichever estimate is larger). The transit distance (twice the tube length) will have an
uncertainty estimated during the measurement process. The combined fractional
uncertainty for vs will then be
2 2
vs  t   L 
     .
vs  t   L 

8
For this file, Logger Pro is set to repeat the collection periodically, so when you see the waveform you
want to examine, press STOP immediately.
9
You might try y  A sin  Bt  C   D .

10
Sound Physics 204C Laboratory

The fractional uncertainty in the vs calculated from Eq. (1) will be the same as the
fractional uncertainty in the temperature, which you can estimate by observing any
fluctuations in the reading and noting the accuracy reported in Appendix D: Vernier
Hardware Spec Sheets.

II. The fractional uncertainty in the frequency calculated by the cyclecounting method
will be the same as the fractional uncertainty in the total time duration of the cycles
involved in the measurement.

Conclusion.

Summarize your results for the measurement of the speed of sound and the
comparison with the value of vs from Eq. (1).

What did you learn about the characteristics of the sounds produced by tuning
forks and musical instruments?

11
T, , Q Physics 204C Laboratory

Temperature, Heat Flow and Thermal Expansion

Introduction. In this set of investigations you will examine several aspects of


temperature and heat transfer. In particular, you will study (I) the rate at which an object
cools to the temperature of its surroundings, (II) the relationship between temperature
sensation and heat flow, and (III) the thermal expansion of metals.

Two bodies in thermal equilibrium have the same temperature. If an object has a
temperature difference Tdiff with respect to its surroundings, heat flow will occur until the
temperatures are equal (so that Tdiff  0). This rate of heating or cooling is often


dTdiff
proportional to Tdiff :   1 Tdiff . The temperature difference Tdiff at time t under
dt
these circumstances is given by
Tdiff  t   Tdiff (0) e t/  , (1)
where Tdiff (t) is the difference between the temperature To(t) of the object and that of the
surroundings at time t, and Tdiff(0) is the initial difference. This is Newton’s Law of
Cooling. The parameter  characterizes the rate of heat transfer between the two and so
depends on thermal conductivities and heat capacities of the materials under
consideration. You will measure To(t) for a warm thermometer in air.

Our subjective assessment of whether an object feels warm or cool to the touch can be
related to the temperature difference between our skin and the object and to the thermal
conductivity of the material of which the object is made. These factors affect the rate at
which heat is transferred during the touch. Again, the heat transfer rate is proportional to
the temperature difference.

An object will change its dimensions in response to a change in temperature (T). The
change in the length (L) of an object due to thermal expansion or contraction is given
empirically by
L   L 0 T , (2)
where L0 is the initial length of the object before the temperature change, and  is the
coefficient of linear expansion for the material of which the object is composed. You will
measure  for one metal.

Procedure. You will need the following:

 computer  LabPro  Logger Pro  melting blocks


 plastic bag  LARGE ice cubes (several trays in freezer)
 beaker/foam cup/tray  tweezers
 temperature probe  thermal expansion apparatus (3)
 hotplate + water pitcher  rubber tubing, ¼” I.D.
 DMM  cables

© 2011 Eric R. Dietz


T, Q,  Physics 204C Laboratory

I. Approach to thermal equilibrium: Place the temperature probe in a plastic bag and
connect it to Ch. 1 of LabPro. When LabPro is powered up and connected to the PC,
open the Heating and Cooling Rate file. Prepare a hot water bath and start data
collection. Place the probe (in bag) in the hot water and observe the heating curve. After
the probe reaches equilibrium with the bath, remove it and let it cool in the air without the
bag.

Sketch the heating and cooling curves and try a curve fit10 with Eq. (1) to determine if it
adequately describes these processes. Report the best-fit parameters for the heating and
cooling curves.

II. Heat flow, thermal conductivities and temperature difference: Touch the two melting
blocks (made of aluminum and plastic) and notice which feels cooler. Can you explain
the different sensations?

Predict: If an ice cube is placed in the center of each block, do they melt at different
rates? If so, which one melts more rapidly?

Test your prediction11. Using language and concepts discussed in class, give a careful
explanation of how and why the temperature sensation is related to the melting rate.

III. Thermal expansion: You will measure the coefficient of linear expansion , defined
by Eq. (2), for steel, copper or aluminum, using the thermal expansion apparatus pictured
below.

Small changes in the length L of the metal tube are measured by the dial gauge12, on
which the smallest division represents 0.01 mm. Temperatures are measured (to within
about 0.2 C) by reading the resistance on the DMM of a thermistor in thermal contact
with the tube.

For one of the three tubes:

10
Note that Eq. (1) involves Tdiff, the temperature difference between the object and the surroundings, while
your temperature probe reads the temperature T of the “object.” You may wish to enter Tdiff as a calculated
column derived from the values of T collected by LabPro, or alternatively add a fit constant (representing
the temperature of the “environment”) to the curve fit expression.
11
Use two ice cubes from the freezer. You may want to put an o-ring on each block to prevent the ice cube
from sliding off as it melts.
12
Note the dial gauge has two hands, and reads like a clock.

13
T, Q,  Physics 204C Laboratory

Verify that both ends of the expansion tube


are seated correctly. Connect the water hose
from the sink tap to the inlet of the apparatus
and put the outlet hose exhaust in the sink.
Let cold water flow through the apparatus
until the thermistor reading stabilizes. Note
carefully and record this initial thermistor
resistance reading on the DMM, which should
be set to the most sensitive resistance scale
that is not overloaded13. Record the initial dial gauge reading.

Start the flow of warm water (let it drain into the sink) and wait until the temperature
indicated by the DMM reading has maxed out and stabilized. Now take final readings of
both the dial gauge and the DMM. Compute L from the dial gauge readings and T
from the thermistor readings. Note that you will need to use the chart on the apparatus to
find the initial and final temperatures; these can be obtained to within 0.2 C by linear
interpolation.

Use Eq. (2) to compute   , the thermal expansion coefficient for the metal in the
expansion tube, given that its initial length is L0  700  1 mm at room temperature.

How does your value of  compare with the accepted one from the table above?

Uncertainties. Your result for the uncertainty in  will depend on (T)—the


uncertainty in the temperature change T—and on (L), the uncertainty in the length
change L. Although T from linear interpolation can be assigned an uncertainty 0.2 C
and L can be read to within half the smallest division on the dial gauge, the uncertainties
in the changes in these will be a bit greater because they involve the difference in two
uncertain quantities. A quantity z that is the difference between two measured quantities
z1 and z2 will have an uncertainty approximated by
 z1    z 2 
2 2
z  ,
according to Eq. (8a) of Appendix A.

For the fractional uncertainty in  calculated from Eq. (2), we have, from Eq. (8b) of
Appendix A,
 L     L      T  
2 2 2

  0      .
  L 0   L   T 

13
A reading of “1.” indicates an overload and that a change to a less sensitive scale is needed.

14
T, Q,  Physics 204C Laboratory

Conclusion.

Do your observations support Newton’s law of cooling? From your observations,


describe the rate at which an object comes to thermal equilibrium with its surroundings
(if the heat transfer process is dominated by conduction).

What are the factors that, according to your observations, determine whether an object
“feels” hot or cold?

Does your calculated coefficient of thermal expansion come within uncertainties of the
“accepted” value? State a couple of the assumptions you have made in calculating .

15
IGL Physics 204C Laboratory

The Ideal Gas Law

Introduction. The equation of state for n moles of an ideal gas relates its
pressure (P), its absolute temperature (T), and its volume (V):
PV  nRT , (1)
where R is the universal gas constant and has the value 8.314 J/molK. Eq. (1) is known
as the ideal-gas law (IGL). In this experiment you will investigate how well the IGL
describes pressure changes in a fixed amount of air (I) resulting from volume changes at
constant temperature (Boyle’s law) and (II) resulting from temperature changes at
constant volume. In connection with (II), you will also attempt to locate the absolute
zero of temperature on the Centigrade scale.

Procedure. You will need the following:

 computer/Logger Pro/LabPro  plastic tubing w/ 2 connectors


 gas pressure sensor  rubber stopper assembly w/valve
 20-ml syringe  125-mL Erlenmeyer flask
 temperature probe  ring stand  utility clamp
 hotplates  foam container in 500-mL beaker
 metal pitchers  French curve
 crushed ice  shovel

I. Pressure vs. volume at constant temperature:


With the LabPro powered up and the pressure
sensor connected to Ch. 1, bring up the P vs. V
(Boyle’s Law) file in Logger Pro. Before
connecting the syringe to the pressure sensor,
move its piston until the front edge of the inside
black ring is positioned at the 5.0 mL mark, as indicated by the arrow at the right. Attach
the syringe to the white stem on the pressure sensor with a gentle half-turn.

Start collecting P vs. V data; when the pressure stabilizes, press , then type the
14
piston position in the box that appears and press “enter.” Collect data for piston
positions from 5.0 mL to 20.0 mL in steps of 2.5 mL.15 Remember to click stop when
you are finished collecting data.

Use the curve-fit routine to fit your P vs. V data; choose an expression to try based on the
IGL of Eq. (1) for constant T. Sketch the resulting graph and record the values of any

14
Data entry and syringe adjustment should be performed by different students. Note that you can redo a
point by pressing ESC after clicking “keep” but before entering a value.
15
To adjust the piston position, stabilize the syringe with one hand and pull the piston with the other; do
not hold the sensor while pulling the piston!

© 2011 Eric R. Dietz


IGL Physics 204C Laboratory

parameters that result from the fitting. Based on the initial data point, compare your fit
parameter to its expected value.

Is the appearance of the best-fit curve consistent with the IGL? Explain.

II. Pressure vs. temperature at constant volume: For a fixed volume and number of moles
of air, you will measure the pressure as the temperature is varied. Since you will measure
the temperature in degrees C, it is convenient to rewrite the IGL of Eq. (1) as
 nR 
P    TC  T0C  , (2)
 V 
where TC is the temperature measured by your probe in C, and T0C is the absolute zero
 nR 
of temperature, also in C. Note that the factor   is held constant, and at STP
 V 
conditions (0C and a pressure of 1 atm or 1.013  105 Pa)--under which your flask will
be sealed--this factor evaluates to 0.371 kPa/C.

Power up the LabPro interface and connect it to the


computer. Plug the pressure sensor into Ch. 1 of the
interface, the temperature probe into Ch. 2, and bring up
the P vs. T file. You will use the rubber-stopper
assembly pictured at the right: the tubing should be
connected to the stopper at one end and the stem of the
pressure sensor at the other with a clockwise turn. Leave the two-way valve on the
stopper open (lined up with the valve stem as shown). The assembly should be twisted
into the neck of the flask to ensure a tight fit. Complete the procedure described below as
quickly as possible to minimize the effects of leaks.

Place the flask in the empty foam container in the beaker


and use the ring stand clamp to secure the rubber stopper.
Pour in ice-water so that the entire flask is covered and
insert the temperature probe into the bath, as shown at the
right. When the pressure and temperature stop changing
(as equilibrium is attained), close the two-way valve (rotate
it ¼ turn so that the handle is perpendicular to the stem).

Click to begin data collection. When the TC and P readouts have stabilized, click
to record the (P,TC) data pair at/near 1 atm, 0C. Replace the ice-water bath with
water at room temperature and collect another data point when the readings have
stabilized. For additional points, use hot tap water and , finally, boiling hot water.

Don’t forget to press when you have finished collecting data!

Do a curve fit to your P vs. TC data, using Eq. (2) as a guide. LoggerPro should give a
 nR 
value of the parameter   as well as a value for T0C, the absolute zero of temperature
 V 

17
IGL Physics 204C Laboratory

on the Kelvin scale. Sketch the P vs. TC data with the best-fit line, and record the value
of these parameters. Leave enough room on both sides of the origin of your graph so you
can locate T0C graphically by extrapolating the line to the zero of pressure.

Uncertainties. For this experiment, most of the calculation of the uncertainties is done
for you by LoggerPro during the curve-fit process.

Conclusion. Decide whether the ideal gas law correctly predicts the observations you’ve
made with air. In particular:

Examine your P vs. V (at constant T) data and decide whether Boyle’s law holds. Is the
fit parameter you obtained in part I consistent with the initial values you typed in for P
and V?

For part II, is your data consistent with the prediction of the ideal gas law—that the
relationship between P vs. T (with n and V held constant) should be linear? How does
your curve-fit value for (nR/V) compare with the number corresponding to your sealed
flask at 1 atm, 0C? How does the value of T0C, the absolute zero of temperature,
compare with the accepted value of 273.15 C?

18
cp and Lf for H2O Physics 204C Laboratory

Specific Heat and Heat of Fusion for H2O

Introduction. In this lab you will measure two important thermodynamic


constants for water: the specific heat cp, and the latent heat of fusion Lf—both measured
at constant pressure.

For a substance that is not undergoing a phase change, the heat input Q is related to its
specific heat cp and its temperature change T:
Q  mc p T , (1)
assuming that the specific heat cp is not a function of temperature over the range T. The
specific heat is then the amount of heat required to raise the temperature of one gram of
the substance by 1 C.

The amount of heat required to melt a mass m of substance (without changing its
temperature) is given as
Q  mLf , (2)
where Lf is the latent heat of fusion for the substance.

Procedure. You will need the following:

 computer/Logger Pro/LabPro  250 mL beakers (2)


 stainless steel temperature probe  100 mL graduated cylinder
 current probe  small blue plastic tongs
 DMM  2 Styrofoam drinking cups
 6V lantern battery  ring standutility clamp
 heater/foam container/600 ml beaker  large (melting) ice cubes
 stirring rod  hotplate
 triple-beam balance  voltage probe (direct to LabPro)
 4 alligator clips  Deionized water
 SPST switch

I. Specific heat of water: Connect the temperature probe to Ch. 1, the voltage probe to
Ch. 2, and the current probe to Ch. 3 of the LabPro. Configure the heater, the battery and
the data acquisition system as indicated schematically below. Set the DMM so it reads
the current through the heater, but leave the switch open.

19
cp and Lf for H2O Physics 204C Laboratory

With the LabPro powered up, open the file Specific Heat of Water. Note that this file
gives simultaneous readouts of the temperature, the current through the probe, and the
voltage across the heater. Momentarily clip the voltage probe wires together. Click Zero
and select both probes. Reconnect the voltage probe clips to the heater terminals.

Fill the foam cup about half full with deionized water and cover it with the heater cap.
Put the assembly into the glass beaker for added insulation. Insert the temperature probe
in the center hole so that it is just above the bottom of the cup and record the temperature
when it stabilizes. Hit Collect and close the switch. Let the system collect data for
5-10 min while gently swirling the beaker to agitate the water. Open the switch and
allow the temperature to stabilize again while agitating. Record the final temperature
before it starts to decrease and hit Stop. Please leave the switch open when not taking
data.

Calibrate the current probe: While the water is heating, calculate and record the ratio of
the heater current read on the DMM to the live readout from the current probe. After the
collection is stopped, enter this correction factor in the formula for the calculated
“Power” column16.

Compute the specific heat of water: Compute T from your initial and final water
temperatures. Carefully determine the mass m of water in the cup. On one of your
graphs, click on the y-axis label and choose to plot “Power,” then select the data on the
graph. Use the AnalyzeIntegral utility to compute the total heat delivered to the water:
Q   I V dt . Finally, use Eq. (1) to compute cp, along with its estimated
heating time

uncertainty.

Compare this with the “book” value for water, which is 4.19 J/gK.

II. Latent heat of fusion for water: Disconnect everything


from the LabPro interface except the temperature probe and
open the file Heat of Fusion. Set up the probe, Styrofoam
cup and beaker with the ring stand and clamp as shown at
the right. Determine the mass of the empty cupbeaker.

Obtain several (5 or 6) large ice cubes and allow them to sit


at room temperature until they have reached 0 C and have
begun to melt (they should be wet). Obtain about 100 mL
of warm water (around 50-60 C), pour it into the cup and
quickly (but carefully) determine the mass of the cupbeakerwater with the balance.

16
Double click on the “Power” column heading and enter your calibration factor in place of the
approximate one already there.

20
cp and Lf for H2O Physics 204C Laboratory

Calculate the mass of warm water (mww). Hit Collect and lower the temperature probe
into the cup so it is about 1 cm from the bottom. Allow the temperature indication to
max out at the water temperature. Record this maximum temperature Tmax and then
quickly add some ice cubes to the water (shake off any water clinging to the ice).

Stirring gently and adding more ice as necessary, allow the mixture to cool to about 4 C
and then carefully fish out all the unmelted ice with the tongs, shaking any excess water
back into the cup. Allow the temperature indication to reach a minimum before it starts
to warm (keep stirring). Record the minimum temperature Tmin (before it starts to
increase) and click Stop.

Use the balance again to determine the mass mim of the melted ice. Compute the
temperature change T for the initially warm water. (You can use the
AnalyzeStatistics feature to confirm Tmax and Tmin.)

Compute the heat of fusion Lf: Applying energy conservation to the warm water and the
ice gives Qice  Qwater  0. Using Eqs. (1) and (2) for the Q’s gives
 mim Lf  mimcpTmin   mww cp  Tmin  Tmax   0 , (3)
Where Tmin (C) represents the change in the temperature of the melted ice when it
reaches the final temperature Tmin of the mix. From Eq. (3) you can compute Lf for H2O,
along with its uncertainty. This is to be compared with the “book” value for Lf of
334 J/g.

Uncertainties.

I. For the specific heat measurement, the dominant uncertainty will likely be that in the
measured temperature increase T  Tfinal  Tinitial . Note that the values of T are subject to
fluctuations for various reasons, but the subtraction to get T compounds this; the
uncertainty in T (represented as u(T) ) is17
 u  T     u Tinitial     u Tfinal 
2 2
.
The fractional uncertainty in cp is then roughly the same as the fractional uncertainty in
T.

The accuracy (uT) for the Vernier temperature probe is rated at 0.2C at 0C and
0.5C at 100C.

II. For the heat of fusion measurement, you should consider u(T) (see ¶ above) along
with the uncertainties in the masses mim and mww and uTmin. If you use the “book value”
of cp for water in this calculation, then neglect its uncertainty. The uncertainty in Lf can
then be estimated by computing its “range,” i.e., its high and low values determined by
the range in the uncertain quantities (masses and temperatures) from which it is

17
In an attempt to minimize confusion here, I have used “” to mean “change in,” and “u” to mean
“uncertainty in.” So u(T) would mean “the uncertainty in the change in T.”

21
cp and Lf for H2O Physics 204C Laboratory

calculated. The uncertainty Lf can be taken as half the difference between its highest
possible and its lowest possible value18.

Conclusion. How do your experimental values of specific heat and latent heat of fusion
for H2O compare with the “book” values for these quantities, considering your
uncertainties?

Can you identify some factors that contribute to systematic errors (as opposed to random
fluctuations and instrumental inaccuracies) in these measurements?

18
The alternative is to solve Eq. (3) for Lf and then use Eq. 8 in Appendix A to estimate Lf.

22
Gamma for Gases Physics 204C Laboratory

Specific Heat Ratio for Gases: Rüchardt’s Method

Introduction.
cp
An important thermodynamic quantity for a gas is its specific heat ratio   , where cp
cv
is the specific heat at constant pressure and cv is the specific heat at constant volume.
This constant appears in the pressure-volume relationship for adiabatic changes in an
ideal gas:
PV   constant . (1)
Rüchardt’s method for measuring  involves measuring the frequency with which a
piston undergoes rapid (adiabatic) oscillations due to the
“springiness” of the gas beneath it. In this lab you will
measure  for air and for He by observing these oscillations.

In the apparatus shown at the right, when a piston of mass m


is in equilibrium, the pressure P0 of the gas it confines is
slightly above the atmospheric pressure to support its weight
mg. If the piston is displaced a small distance x from
equilibrium (x  0), then the gas pressure increases19
according to Eq. (1), restoring the piston back toward
equilibrium. The result is simple harmonic motion of the
piston.

To obtain an expression for the natural frequency of these oscillations, we note that the
net force on the piston created by the change P in pressure due to the vertical
displacement x is Fy  A P . Now P can be related to the volume change V of the gas
and hence to the piston displacement x by implicitly differentiating Eq. (1):
P P
P   0  V   0 A x . Thus the restoring force is given by
V0  A x V0
P P
Fx  A P   0 A 2 x   0 A x , where we have used V0  H 0 A for this apparatus.
V0 H0
d2 x
Putting this into Newton’s second law (  Fx  m ) gives for the equation of motion
dt 2
for the piston:
d2x  P A
2
  0  x . (2)
dt m H0 
 
02

19
The oscillations in this case are so rapid that we assume the process is adiabatic, so Eq. (1) applies.

© 2011 Eric R. Dietz


Gamma for Gases Physics 204C Laboratory

As in the discussion of Sec. 14.4 of the text (Physics 204A), we see that this equation
represents simple harmonic motion in x; the natural angular frequency 0 is given in
terms of the constant on the RHS between the minus sign and the displacement:
P A
0 2  0 . (3)
m H0
We see that (i) measuring the frequency of the free oscillations of the piston can give us
the value of , and (ii) the frequency decreases for larger values of H0—larger volumes of
gas are less “springy.”

Procedure. You will need the following:

 computer/Logger Pro/LabPro  helium balloon w/clamp


 heat engine/gas law apparatus  tubes w/ connectors
 gas pressure sensor  air

I. Specific Heat Ratio for Air

Connect the gas pressure sensor to one port of the gas law apparatus and plug the cable
into the LabPro. The valves on both ports should be open. When the LabPro is powered
up and connected to the computer, bring up the file Specific Heat Ratio for Gases. The
file is set so that data collection is triggered when the sensor sees pressure changes.

For the first data point, draw the piston up so that the lower edge is even with a graduated
reading of 90-95 mm (this is H0) and completely shut the valve on the open port. Hit
Collect and push down the piston platform with your thumbs and then release it by
drawing your thumbs off the edge of the platform. Logger Pro should collect the
pressure vs. time data for several oscillations of the piston. To determine 0, do a curve
fit (AnalyzeCurve Fit) to this data using the damped simple harmonic expression20 for
P(t):
P  P0  A cos  0 t    e  t /  ,
and record the best-fit values for P0, 0, and . (Tips: (1) Select for the curve fit only the
portion of the data that exhibits oscillations with significant amplitude and (2) If the
curve fit appears poor, hit “Try Fit” again.)

Repeat this for a few (3 or 4) additional well-separated values of H0. (Change H0 by


opening the valve on the open port, moving the piston (or allowing it to move), and then
resealing the valve.) Sketch one of the P vs. t curves. Record 0 for each of the trials (P0
should remain constant) and tabulate your data in a table like the one below:

20
Note that since the oscillations are damped, the fit parameter 0 slightly underestimates the free-
oscillation frequency in Eq. (3). A more accurate value for this frequency (to be used in the data tables and
2
1
the plots) would be 0,corr  0    , where 0 and  are the best-fit parameters in the fit equation
2


for P(t) above.

24
Gamma for Gases Physics 204C Laboratory

AIR
P0  Pa
1
H0 (m) 0 (s ) T0 2  4 2 (s2)
02
(Logger Pro can do this)

Compute  for air: Although it would certainly be possible to determine  from one value
of 0, there are systematic errors due to the extra volume in the system not accounted for
by H0—the volume of air in the connecting tubes, for example. Looking again at Eq. (3),
we see that it implies a linear relationship between the square of the oscillation period T0
(  2/0) and the equilibrium column height H0:
 4 2 m 
T0 2    H0 . (4)
  P0 A 
4 2 m
So a plot of T02 vs. H0 should be a straight line with slope , and in terms of this
 P0 A
slope,  is then given by
4 2 m
 .
P0 A  (slope)

Bring up the file LoggerProFile (from the desktop 204C lab folder) and plot T02 vs. H0
for your data. Using the data for the piston printed on the apparatus to get values for A
and m, deduce a value of air by doing a linear fit (AnalyzeLinear Fit). Be sure to
record both the slope of the line and its uncertainty. Sketch the best-fit line with the data
points.

II. Specific Heat Ratio for Helium

Repeat these measurements for He gas. To load the cylinder21 with He: Open the valve
on the free port of the apparatus and bring the piston down to the bottom of the cylinder.
Connect the piece of tubing with the check valve to the free port. Hold the stem of the
He balloon on the end of the tubing and slowly open the balloon’s clamp, allowing the
piston to reach the 90 mm mark. Close the valve completely.

Retake the 0 vs. H0 data and graph T02 vs. H0 as you did for air. Again, determine He
from the slope of the best-fit line and sketch the line on the graph you drew for the air
data.

21
Before loading He into the cylinder, it is a good idea to flush the tubes to get the air out. This is done
with both plastic clamp valves open and the piston held to the bottom of the chamber while He gas is
allowed to flow through the tubes (with the pressure sensor momentarily disconnected).

25
Gamma for Gases Physics 204C Laboratory

Uncertainties.

The uncertainty in the slope ((slope)) of the line is given with the linear fit22 in Logger
Pro. The mass m and diameter D of the piston also have uncertainties that are printed on
the apparatus. The uncertainty in P0 is either the extent of the real-time fluctuations or
the overall variation during the data taking, whichever is larger. In terms of these, the
fractional uncertainty in  can be estimated as

2 2 2 2
  m   P0   D   (slope) 
     2    .
  m   P0   D   slope 

Conclusion.

From your measurements and calculations, report the values of    for air and for He
gas. Compare these with the “table” values of 1.40 for dry air and 1.66 for He (at 20 C).

22
If the uncertainty does not appear in the fit box, double click the box and select “Show Uncertainty.”

26
Reflection and Refraction Physics 204C Laboratory

Geometrical Optics I: Reflection and Refraction of Light

Introduction. The way in which light reflects


and refracts at the interface between two
media is at the root of our discussion of image
formation (geometrical optics). A light ray in
medium 1 incident on a flat interface is, in
general, partially reflected and partially
refracted into medium 2, as depicted at the
right.

For a smooth (specular) reflecting surface, the law of reflection states that the incident
ray, reflected ray, and the normal to the surface all lie in the same plane (called the plane
of incidence) and that the angle of reflection (r) is equal to the angle of incidence (1):
r  1 . (1)
If a reflecting surface is curved rather than flat, then Eq. (1) can be applied to each small
portion by considering it to be approximately flat and tangent to the curve at that point.
For a concave mirror with radius of curvature R, the application of this law shows that
rays parallel to the mirror axis converge at a point F on this axis that is located a distance
f from the mirror, where
f  12 R . (2)
The point F is the focal point of the mirror; the distance f is its focal length.

The law of refraction (Snel’s law) states that the ray transmitted into medium 2 is
refracted (“bent”) so that it makes an angle 2 with the normal that depends on 1 and the
indices of refraction n1 and n2:
n1 sin 1  n 2 sin 2 . (3)
This ray also lies in the plane of incidence.

In the case where light enters a medium with a smaller index of refraction (e.g., from
water to air so that n2  n1), there is a critical angle of incidence C beyond which the
light is not transmitted into the lighter medium but rather is completely reflected from the
interface:
n
sin C  2 . (4)
n1
This is known as total internal reflection (TIR).

In today’s lab you will investigate the law of reflection for a flat mirror surface and
observe how a concave cylindrical mirror brings light rays to a focus in front of it. You
will also investigate Snel’s law and observe TIR.

© 2011 Eric R. Dietz


Reflection and Refraction Physics 204C Laboratory

Procedure. You will need the following:

 computer/Logger Pro  three-surface mirror


 ray table  base  Vernier caliper
 cylindrical lens  straight edge with hole
 optics bench  white paper
 PASCO light source  desk lamp

Note: Make no marks on the ray table.

I. Reflection:

A. Law of Reflection (Plane Mirror):

Mount the light source and the ray table base on the optics bench and put the ray table on
the base with the Cartesian grid (mm scale) facing up. Position the light source near the
edge of the ray table and adjust its slit mask so that it projects one ray across the middle
of the top of the table. Rotate the table so that the light shines along the zero degree line
of the table.

Place the plane surface of the three-surface mirror on the


90 line so that it faces the light source. As shown at the
right, you can vary the angle of incidence by rotating the
table.

Measure the angle of reflection for six different angles of


incidence. Use Logger Pro to graph r vs. 1 and do a
curve fit and sketch the result. Is your result consistent with the law of reflection?

B. Focal Point of a Concave Mirror: Adjust the slit mask on the front of the light source
so that five light rays shine across the top of the ray table. Turn the three-surface mirror
so that the concave surface faces the light source with the center ray along the
zero-degree line, hitting the center of the concave surface.

The point where the rays intersect on the mirror axis is the
focal point F of the mirror. Measure the focal length
f  f, which is the distance between F and the center of
the mirror.

Compute f from the geometry of the mirror using Eq. (2).


You will need to deduce R, the radius of curvature, by
measuring X and Y, as in the diagram at the right, using a
straight edge with a hole in it (to measure 2X) and a

28
Reflection and Refraction Physics 204C Laboratory

Vernier caliper23 (to measure Y through the hole). Using the indicated right triangle,
compute the radius as
X2  Y2
R . (5)
2Y
How does the focal length derived from your measurements of the curvature of the mirror
surface compare with the focal length measured on the ray table?

II. Refraction (Snel’s law and TIR):

A. Snel’s Law: Adjust the slit mask and the ray table so
that a single ray is projected across the table along the
zero-degree line. Place the acrylic half-cylinder (avoid
touching the plastic) on the table so that the flat surface
faces the light source with the ray striking its center and
with its edge along the 90 line. The setup is illustrated
at the right.

Rotate the ray table to set six different values of the angle of incidence 1 and measure
the corresponding angle of refraction 2 for each. Plot sin 1 vs. sin 2 in Logger Pro24
and do a linear curve fit to determine the value of n for acrylic, along with its uncertainty,
assuming Eq. (3) is correct. Sketch the fit with the data points.

Observe: Project a refracted ray on a piece of white paper and notice the color separation
for high 1. What causes this?

B: TIR: Rotate the ray table so that the ray is incident on the curved surface of the
half-cylinder. Find the minimum angle (C  C) such that total internal reflection
occurs at the flat surface. Compute n  n from this measurement, using Eq. (4).

How do these two values of n compare?

Uncertainties.

For the focal length f derived from the dimensions of the curved mirror, we have
f  12 R , but for R we have from Eq. (5) above and Eq. (8) in Appendix A,
2
 1 X  
2 2
X
R   X    1  2  Y  .
Y  2 Y  

23
If you are uncertain as to how to read a Vernier scale, see Appendix C: Reading a Vernier Scale. Note
Ex. C at the end, which shows how to make a depth measurement like the one you will make for Y.
24
If you use the Calculated Column feature to compute the sines, be sure to set the appropriate angle units
in FileSettings.

29
Reflection and Refraction Physics 204C Laboratory

For the index of refraction n from the Snel’s law data, n will be just the uncertainty in
the slope of the best-fit line. For the uncertainty in n from the TIR measurement, we
have from Eq. (4):
n  C  n cot C  ,
where C is in radians.

Conclusion.

I. Are your plane mirror results consistent with the law of reflection (explain)? For the
curved mirror: How does the focal length measured on the ray table compare with the one
you computed from the mirror’s dimensions, considering the uncertainties in both values?

II. Do your observations confirm Snel’s law? Explain. How does the index of refraction
n for acrylic computed from the sin 1 vs. sin 2 data compare with that estimated from
the TIR observation, considering uncertainties? Report your best value of n for acrylic.

30
Image Formation Physics 204C Laboratory

Geometrical Optics II: Image Formation with Mirrors and Lenses

Introduction. A mirror or lens with curved surfaces can redirect rays from an object to
form an image. As we discuss in class, application of the laws of reflection to a thin
mirror surface or the laws of refraction to the surfaces of a thin lens yields the
relationship between object distance s, image distance s, and focal length f:
1 1 1
  , (1)
s s f
where s and s are both measured from the surface of the lens or mirror. Note that f is
positive for converging elements (concave mirrors and convex lenses), and negative for
diverging elements (convex mirrors and concave lenses). Additionally, the lateral
magnification m, defined as the ratio of the linear sizes of the image and object, also has
the same expression for both mirrors and lenses:
s
m . (2)
s
Note that a negative value of m signifies an inversion of the image with respect to the
object.

In this lab you will have the opportunity to observe the properties of some real images
formed by mirrors and lenses. You will also investigate Eqs. (1) and (2) by making
quantitative measurements of the locations and sizes of images formed by these elements.

Procedure. You will need the following:

 computer/Logger Pro  Vernier calipers and straight edge


 concave mirror  converging lens (f  100 mm)
 half-screen  diverging lens (f  150 mm)
 optics bench  screen
 PASCO light source  aperture accessory disk (rotatable)
 desk lamp

I. Images from a concave (converging) mirror: Configure the illuminated object (crossed
arrows) and the mirror, with the half-screen between these two, as shown below.

© 2011 Eric R. Dietz


Image Formation Physics 204C Laboratory

NOTE: With the light box mounted correctly, the white object plate should be vertically
above the notch at the bottom of the mounting bracket. The distances s and s are then
measured using the tape along the track and the notches on the mounting brackets.

A. With the room lights off, adjust s and/or s (measured from the mirror) so that a clear
image is obtained on the screen25. Measure s and s using the tape along the track, and
repeat this for three more values of s. Record the manufacturer’s value of the focal
length f.

Note: Please turn off the light source when it’s not in use.

Open LoggerProFile.cmbl from the desktop 204C lab folder and plot 1/s vs. 1/s.
Perform a linear fit (AnalyzeLinear Fit)26. Record the slope and intercept with their
respective uncertainties. Is the best-fit slope what you expect? Does the y-intercept
agree your expectations from Eq. (1), given the value of the focal length printed on the
mirror mount? Sketch the data with the best-fit line.

B. Draw a ray diagram depicting the formation of an image by this mirror,27 showing its
size and location.

C. Predict: What will be the effect on the image if part of the mirror is covered?

Observe: Snap the rotatable aperture disk into the mirror mount. Observe and record the
effect of using smaller portions of the mirror to form the image. Do your observations
match your predictions?

II. Images from a convex (converging) lens: Configure the illuminated object (crossed
arrows), a converging lens, and the (full) screen as shown below.

25
If you have trouble locating an image, try calculating s for your chosen value of s  2f and putting the
screen there.
26
Enter only decimal numbers (e.g., 0.067) in the Logger Pro data columns, no fractions (e.g., 1/15).
27
A ray diagram consists of at least two rays that emanate from at least one non-axial object point.

32
Image Formation Physics 204C Laboratory

A. As you did for the mirror, find the image locations s for four different object distances
s (s and s are measured from the lens). Additionally, measure and record the object
height and the height of one of these images (with their uncertainties).

Open LoggerProFile.cmbl from the desktop 204C lab folder to plot 1/s vs. 1/s and
record the slope and intercept with their respective uncertainties. Compare the slope and
intercept of the best-fit line to your expectations from Eq. (1) and the value of f printed
on the mount. Sketch the data with the best-fit line.

B. Draw a ray diagram depicting the formation of one of the images.

C. For one image, compute the magnification as the ratio of the linear sizes of image and
object and compare it to the prediction of Eq. (2). Don’t forget the uncertainties.

D. Predict: What will happen to the image if a concave (diverging) lens is placed
between the convex lens and the screen?

Observe: Try this and locate the new image. Explain what happened.

Uncertainties.

The uncertainties in the slope and the intercept of a best-fit line will be given by Logger
Pro, but it may be necessary to double-click the fit result box and request this explicitly.

The uncertainties in the ratios of image-to-object heights or distances can be estimated by


noting, e.g., that, from Eq. (8(b)) of Appendix A,
2 2
 s   s   s 
       , etc.
s  s   s 
Conclusion.

Do your measurements of image location and size for the mirror and lens appear to be
consistent with Eqs. (1) and (2) for thin elements? Explain completely.

What effect does covering a portion of a lens or mirror have on the image? Explain this,
based on your understanding.

33
Microwave Interference Physics 204C Laboratory

Physical Optics I: Microwave Interference and Diffraction

Introduction.

All wave phenomena can be made to exhibit constructive and destructive interference and
diffraction. In this experiment, we investigate these effects for electromagnetic waves in
the microwave region of the spectrum with a wavelength of about 3 cm. In the next lab,
we will see that visible light, which has a wavelength on the order of 105 times smaller,
exhibits the same type of behavior.

In Part I of this lab, you will measure the wavelength of microwaves in two different
ways: A. The Michelson Interferometer and B. Standing Waves. Your microwaves are
produced by a Gunn diode and have a frequency of 10.5 GHz.

A. The interferometer is described in your text (read Sec. 22.6) for light, but the same
principle applies to microwaves. In this device (shown below under “Procedure”) when
one of the reflectors is moved by one-half wavelength, the path length difference (PLD)
for the two paths changes by a full wavelength (), and the amplitude at the receiver is
the same as it was before moving the reflector. The reflector displacement for successive
peaks in the signal thus represents /2.

B. A standing wave can be set up between the transmitter and the full reflector with a
probe mounted in the middle to “sample” the wave by reflecting part of it to the receiver.
As the reflector is moved, standing waves are set up for certain positions of the reflector,
thus causing a maximum indication in the receiver. The distance between reflector
positions for standing waves (peak signals) is /2.

In Part II, you will record the two-slit intensity pattern produced when microwaves are
incident on a mask with two slits. The intensity is a maximum (constructive interference
occurs) when the PLD between waves from the two slits (spacing d) is either zero or an
integer multiple of . If we approximate the PLD as d sin , then
d sin   m  (1)
gives the values of  where the mth maximum is expected. Likewise, if the PLD is an
odd number of half-wavelengths, then a minimum should occur:
 1
d sin    m    (2)
 2
You will record the two-slit intensity pattern and compare the peak positions to those
predicted by Eq. (1).

34
Microwave Interference Physics 204C Laboratory

Procedure. You will need the following:

 computer/Logger Pro  half reflector


 microwave transmitter (Gunn diode)  probe
 microwave receiver  Vernier calipers
 goniometer assembly  double slit plate
 full reflectors (2)  French curve

Tips for working with the microwave apparatus:

 Make sure there are no obstructions in the path of the microwaves before taking a
reading.

 Make sure the transmitter and receiver are aligned with the long axes of the rectangular
horns horizontal (the microwaves are polarized).

 Set the gain so that there is a nearly full scale indication (if possible) when the receiver
is at the position of maximum signal for any configuration.

I. Measurement of the Wavelength:

A. Michelson Interferometer. Set up the Michelson Interferometer as in the diagram


below and connect the receiver (R) to the DMM. Angle the half-reflector to maximize
your signal.

Move the 270 reflector in 1 mm increments and record the relative power indication
(arbitrary units) of the receiver vs. the reflector displacement directly in Logger Pro
(open from the desktop) or in your notebook. You should collect enough data to show at
least two peaks of the voltage vs. displacement curve. Sketch the graph.

35
Microwave Interference Physics 204C Laboratory

By observing the relative power variations as you move the reflector, determine the
locations of two peaks in the signal and compute the wavelength    for the
microwaves. Compare  with the wavelength computed from the frequency, which is
10.5 GHz.

B. Standing Waves. Starting with the interferometer setup of part A,


remove the 270 full reflector and replace the half-reflector with the
probe, consisting of the narrow aluminum strip pictured at the right.
Angle the probe so that it reflects some of the signal directly into the
receiver.

Move the 180 full reflector in 1 mm increments and again record the
displacement vs. the signal strength directly in Logger Pro (started from the desktop) or
in your notebook. Collect enough data to show at least two peaks in the signal intensity.
Sketch the graph.

Again, determine the wavelength    from the relative power variations as the
reflector is moved and compare it to the one computed from the microwave generator
frequency in part (A).

II. Double-Slit Interference:

Set up the apparatus with the double-slit plate, pictured below in the   0 position.

Align the transmitter (T) symmetrically with respect to the slits so that the same
microwave power passes through each. Rotate the receiver to record the signal strength
at 5 increments for 60    60, or over a range large enough so that you can locate
two peaks on both sides of center, and enter the data directly into Logger Pro (opened
from the desktop) or in your notebook. Sketch the plot of signal strength vs. angle,
identifying the orders of interference maxima. Estimate    for the observed
interference maxima by slowly sweeping the receiver through the angular range given
above and recording the peak positions.

36
Microwave Interference Physics 204C Laboratory

Measure the center-to-center slit spacing on the double-slit mask28 and use it in Eq. (1) to
compute the expected values of  for the maxima you observed. Compare these with the
angles estimated from your data.

Note that the condition for an interference maximum stated in Eq. (1) is derived with the
assumption that the slit spacing is very small compared to the distance between the slit
mask and the receiver. The worksheet file:
Microwave Double Slit Intensity Pattern.mcd,
located in the 204C lab folder, calculates the intensity pattern from scratch without
making this assumption. To calculate this worksheet, enter the parameters for your
apparatus and click MathCalculate Worksheet.

Uncertainties.

You should estimate uncertainties in your measurements of the wavelength and the
double-slit pattern peak positions from the graphs used to determine them.

The uncertainties  (in radians) in the predictions of Eq. (1) for the angles  at the peaks
can be estimated from
2 2
    d 
  sec      .
    d 

Conclusion.

What are your measured values of the wavelength of your microwaves, as derived from
both the interferometer and the standing wave methods? Are these consistent with each
other and with the calculated wavelength for 10.5 GHz radiation?

Do your predicted angular positions for the two-slit maxima coincide, within
uncertainties, with those that you measured?

28
Use the Vernier calipers to measure the width and the edge to edge distance. Then subtract the width to
get the distance d from center to center.

37
Interference of Light Physics 204C Laboratory

Physical Optics II: Interference and Diffraction of Light

Introduction. In this lab you will study the Fraunhofer


(far-field) patterns produced by a beam of monochromatic
laser light of wavelength   670  10 nm when it passes
through small apertures and onto a distant screen, as
depicted at the right.

For a single slit of width a, light passing through different parts of the slit interfere and
produce a single-slit diffraction pattern with minima found at angles  given by
a sin   m  , (1)
where m is a non-zero integer and is referred to as the order of the minimum. For a line
obstacle of width a, the pattern is similar and has minima at the same locations as for the
slit.

For multiple (two or more) narrow parallel slits with their centers spaced a distance d
apart, the maxima (bright spots) appear where light from the individual slits
constructively interfere. This happens for angles  such that
d sin   m  , (2)
where m is any integer. For the case of two slits, the angular positions of the minima are
given by
d sin    m  12   , (3)
where again m is any integer.

The multiple slit pattern is complicated by the fact that the slits themselves are not
infinitely narrow. If they each have a width a, then the pattern has interference minima
and maxima as described above, but the overall intensity is “multiplied” by that of the
single slit diffraction pattern. So the brightness of the two-slit maximum will vary
according to the wider single-slit diffraction “envelope.” This means, for example, that if
an interference (two-slit) maximum is at the same location as a single-slit minimum, then
the bright spot is “suppressed” by the single-slit dark spot.

Instead of measuring the angular positions for minima and maxima directly, it is more
convenient to measure their linear displacements (y) from the center of the pattern, as
indicated in the diagram above. In this case the angles  are simply related to y and L:
y
tan   . (4)
L
If y  L , the small angle approximation that says tan   sin    ( in radians) can
simplify the use of the above relationships.

© 2011 Eric R. Dietz


Interference of Light Physics 204C Laboratory

Procedure. You will need the following:

 laser diode (  670  10 nm)  multiple slit set


 single slit set  screen + blank sheet of paper
 optics bench  laser mount
 straight edge  desk lamp

Set up the experiment as indicated above, with the laser diode and the screen at opposite
ends of the optics bench. Switch on the laser so the light beam is near the center of the
screen.

I. Single apertures: Install the single slit set on its holder and adjust the dial so that a
single slit intercepts the laser beam and a diffraction pattern is visible on the screen.
Attach a blank piece of paper to the screen.

Note that (1) You can adjust the horizontal and vertical beam alignment with the knobs
on the back of the laser to get the best pattern and (2) Larger separations between the
slits and the screen result in patterns that are more spread out and easier to observe and
measure.

A. Qualitative: Observe the diffraction patterns for (1) a single slit, (2) a circular hole,
and (3) a line (obstacle). For each of these, sketch the pattern and the intensity vs.
position curve to which you think it corresponds.

B. Quantitative: Select a single slit and record the indicated width (given in mm) that is
printed on the mask. Position the slit dial and project the diffraction pattern on the
screen. Mark the locations ym of the minima (located symmetrically between bright
spots) on the screen for some value of m and measure the distance L  L between the
screen and the mask. Record ym  ym.

Use Eqs. (1) and (4) to compute the width (a) of the slit, along with its uncertainty.
Compare this with the value printed on the mask.

II. Multiple Slit patterns: Replace the single-slit mask with the one with n slits (n  1).

A. Qualitative:

(1) Double Slits. Select two double-slit apertures and record the values on the mask for
the width (a) and separation (d) for each pair. Observe their patterns on the screen and
sketch them along with the corresponding intensity vs. position profiles. How are these
two patterns different? How are these differences related to the values of a and d?

39
Interference of Light Physics 204C Laboratory

(2) Multiple slits: Select a multi-slit (n  2) aperture and sketch the pattern and the
intensity profile. How is it different from the double-slit patterns?

B. Quantitative Double Slit: Select one of the double-slit apertures and record the values
of a and d printed on the mask. On the paper where the pattern is projected, mark the
locations ym of the interference maxima and minima, along with their order (the m-value
in Eqs. (2) and (3)). Also mark the locations of the diffraction minima along with their
order (m-value from Eq. (1)). Be sure to include an estimate of the uncertainties in these
distances.

Based on these measurements and Eqs. (1)-(4), compute the slit width a  a and the slit
spacing d  d and compare these to the values printed on the mask.

Uncertainties. The uncertainties in the locations ym of the pattern features will be at


least half of the smallest scale division of your measuring device. These values of ym,
in turn, can be used to estimate the uncertainty in the calculated slit dimensions. Eq. (8b)
in Appendix A along with Eqs. (1)-(4) suggest that, for small angles, the fractional
uncertainty in a or d will be approximately
2 2 2
 y m      L 
      .
 ym      L 

Conclusion.

How do your calculated widths for the single and double slit compare with those printed
on the mask? How does your calculated double-slit separation compare?

40
Intensity and Polarization Physics 204C Laboratory

Physical Optics III: Intensity and Polarization of Light

Introduction. In this set of experiments you will explore:


(I) the way in which the intensity of light from a source diminishes with distance and
(II) the polarization of light from various sources along with the action of a linear
polarizing filter as both a polarizer and an analyzer.

I. The intensity (or irradiance) of a light beam at some point refers to the time-average
rate, per unit area, at which energy is transported across a surface perpendicular to the
direction of propagation. It is often expressed in SI “radiometric” units of W/m2. The
sensor and software you will use in this set of measurements displays the intensity in
“photometric” units29 of lux; 1 lux  1 lumen/m2. For a 555-nm monochromatic light
source, 1 lumen is equivalent to 683
1
W.

The way in which the intensity (I) diminishes as one moves further away from a light
source depends somewhat on the geometry of the source. For a point source, the rate of
energy flow (power P) through a sphere of radius R centered on the source is
just P  I  A , where A  4R2 is the area of the sphere. If no absorption is taking place
between the source and the observer, then this power must be the same for all such
spheres: I(4R2)  power. Thus for a point source, we expect
1
I 2 (1)
R
to represent the way in which intensity falls off with distance R from the source.

II. A Polaroid sheet (or filter) has a


characteristic direction in the plane of the sheet
called the polarizing axis: light incident on the
sheet will be transmitted with its electric field
polarized along this axis, as illustrated at the
right. The filter works by simply absorbing all
the light polarized perpendicular to the axis.
The intensity I of light transmitted through an ideal Polaroid filter thus follows two rules:

(i) For unpolarized light (as in the illustration), the transmitted intensity I is half of the
original incident intensity I0: I  12 I0 , and

(ii) For light that is completely polarized with its electric field at an angle  with respect
to the polarizing axis, the intensity is given by the Law of Malus:
I  I0 cos 2  . (2)

A single Polaroid sheet can be used as a polarizer to create polarized light or to change
the direction of polarization. It can also be used as an analyzer to determine the state of

29
For further discussion of photometric units, see Greene, N., The Physics Teacher, V. 41, p. 409 (2003)

© 2011 Eric R. Dietz


Intensity and Polarization Physics 204C Laboratory

polarization of an incident beam by rotating it and noting the variations in the transmitted
intensity.

Procedure. You will need the following:

 computer/Logger Pro/LabPro  French curve  triangle


 light sensor  laser diode  mounting block
 light sensor mount  two linear polarizers
 optics bench  polarizer holder
 PASCO light source  straightedge
 black cloth  desk lamp

I. Intensity vs. Distance: Set the switch on the light sensor box to the 0-600 lux range and
plug the cable into Ch. 1 of the LabPro. With the LabPro powered up and connected to
the computer, bring up the file Brightness vs. Distance. Zero the light sensor by
covering the photodiode element in the probe head, clicking Zero and waiting until the
intensity reading is zero.

With the light sensor holder block on the bench, mount the probe head with the rim of the
blue plastic cap against the edge of the metal slot; the photodiode should face the light
source and the probe tube should be aligned along the bench. The photodiode is situated
about 3 mm back from the front of the cap. If you determine the distance between this
element and the front surface of the wood block, you can use the front edge of the block
as a reference as the sensor position is varied along the bench.

The point source should be mounted on the bench and aligned with the sensor tube. Note
that the light box should be mounted in the bracket so that the position of the point source
along the bench is indicated by the notch at the bottom.

The room lights should be out and the black cloth should be used to eliminate stray light.
Click Collect and record the intensity for nine source-detector separations between 15 cm
and 45 cm. For each measurement, click Keep to record the intensity and enter the
separation as calculated from the readings on the bench tape. Click Stop to terminate the
measurements.

Note: Please unplug the point source when it’s not in use.

Do a curve fit30 to your I vs. R data to determine the power of the separation that best
describes the variation in intensity (note also the uncertainty). Sketch the result (the
French curve may help).

30
You might try, e.g., I(R)  AR B to determine fit parameters A and B.

42
Intensity and Polarization Physics 204C Laboratory

II. Polarization:

A. Qualitative: Put a single polarizing filter disk in the dual filter holder and observe light
from various sources through the filter as you rotate it. Interesting sources include light
from the sky, light reflected from the floor (glare from overhead lights or desk lamps),
direct light from desk lamp or overheads, light from your computer monitor screen.
Record your observations and note which of these appears to be at least partially
polarized.

What happens as you look at, say, your desk lamp through two filters and rotate one of
them? Explain.

B. Quantitative: With the light sensor on the bench, set up the laser diode in the bracket
and align it (with the bracket screw and the knobs on the back) so that the beam hits dead
center on the photodiode sensing element. Set the switch on the sensor box to the
6000 lux range and bring up the Polarization of Light file. Cover the sensor and Zero
the reading.

Mount the filter holder on the bench with a single filter in it facing the laser; rotate the
filter until the reading is maximum and keep it in this position31. Put the second filter in
on the other side (away from the diode) and rotate it also until the reading is again
maximized. (You should end up with both polarizer axes nearly vertical and parallel.)

Click Collect and then Keep to record the first intensity, then enter 0 for the angle
reading (the two axes are parallel for max intensity). Rotate the second filter in 10
increments, recording the intensities and the angles (in degrees with respect to the initial
starting angle). When the rotation is at least 200, click Stop to terminate the
measurements.

Try to fit your data either manually32 or with autofit to an expression in the form of Eq.
(2). You can guess the value of the constant (I0) by looking at your maximum (initial)
value for I(). Sketch the fit with your data.

Conclusion.

Is your Intensity vs. Position data consistent with what would be expected from
conservation of energy, as expressed by Eq. (1)?

Which of the light sources you observed exhibited some degree of linear polarization?
Do any of your observations surprise you?

31
Again, the room lights should be off and the black cloth used to eliminate stray light.
32
For a manual fit: Use AnalyzeCurve Fit and select “Manual” from the upper right portion of the dialog
box and type in the equation and the parameter values. For an autofit: Use an expression of the form
D*(cos(xC))^2B.

43
Intensity and Polarization Physics 204C Laboratory

Is your Intensity vs. Angle data from the double polarizer consistent with the Law of
Malus, as expressed by Eq. (2)?

44
Uncertainties Physics 204C Laboratory

Atomic Spectroscopy and the Bohr Model

Introduction. The Bohr model was the first description of the atom that led to
quantitative agreement between predicted and measured wavelengths emitted and
absorbed by single-electron atoms. In this experiment you will measure the wavelengths
of light emitted from hydrogen atoms and compare them with those calculated using the
Bohr model.

The Bohr Model: The Bohr model depicts a single electron in a circular orbit about a
nearly stationary positively charged nucleus. Only certain discrete orbits are allowed
with corresponding integer quantum numbers (n), radii (rn), and energies (En). A
transition in which an electron in an orbit (state) with energy Ei and quantum number ni
moves to an orbit (nf) of lower energy Ef results in the emission of a single photon with
an energy equal to the difference in energy of these two orbits, as pictured in Fig. 1.

Figure 1
The energy Ephoton ( hf) of a photon emitted in this process would be just the difference
Ei  Ef, so that the corresponding wavelength  would be
c hc
  . (1)
f E ph
The Bohr model applied to the hydrogen atom gives the value of the orbit energies33 as
13.61 eV 2.180 1018 J
En   . The emitted photon energies are thus calculated as
n2 n2
 1 1 
E photon  E i  E f  2.180  1018 J  2  2  . (2)
n
 f n i 

The emitted wavelengths can then be calculated from Eq. (1).

33
Recall that 1 eV = 1.602  1019 J.

© 2011 Eric R. Dietz


Electron Diffraction Physics 204C Laboratory

The only visible34 photons emitted from the hydrogen atom, according to this model,
result from transitions to the n  2 orbit from higher energy states. This set of transitions
(shown in Fig. 1) is known as the Balmer series, and only the four longest wavelengths in
this series are considered to be in the visible region. You may only be able to observe
three.

Grating Spectrometer: The grating spectrometer is a precision instrument that you will
use to measure the wavelengths of light emitted from a hydrogen discharge tube.

Figure 2
Light from the source, comprising several wavelengths, is incident normally on the plane
of the diffraction grating, as in Fig. 2. The grating consists of ruled lines separated by a
distance d, which must be measured to “calibrate” the spectrometer. Diffracted light is
observed through a telescope, which can be adjusted continuously to any angle  with
respect to the incident beam. The interference condition states that maxima for light of
wavelength  will be observed for angles m such that
d sin  m   m   m  0,  1,  2 , (3)
where m is the order of the maximum. Measuring  for any line thus gives  directly.

Procedure. You will need the following:

 grating spectrometer  high voltage power supply


 Hg discharge tube  black cloth
 H discharge tube  desk lamp

Adjust the Grating Spectrometer:

34
Visible light has wavelengths in the range 400 nm    700 nm with corresponding photon energies in
the range 1.8 eV  Ephoton  3.1 eV.

46
Electron Diffraction Physics 204C Laboratory

Note: Handle the discharge tubes by the ends (don’t touch the middle) and the grating by
the edges (avoid touching the ruled surface). Avoid touching the high voltage electrodes
of the power supply. Be aware that the tubes can become hot.

1. Familiarize yourself with the spectrometer. Do not adjust the platform and leveling
knobs of the collimator or the telescope. If you are unsure as to how to read the angle
scale, refer to Appendix C: Reading a Vernier Scale. Note that on the main (outer) angle
scale, the smallest division represents half a degree; the Vernier (inner) scale runs from 0
to 30 minutes of arc (1 min  601 0 ) so you can interpolate between smallest divisions on

the main scale.

2. Install the Hg discharge tube in the power supply and place the center of the tube
directly in front of the collimator slit. If the vertical center of the tube does not coincide
with the horizontal axis of the collimator, carefully adjust things so that it does. Switch
on the unit. Looking through the telescope in the   0 position, gently position the
source in front of the slit and rotate the black knurled end cap on the end of the collimator
so that a bright thin vertical line is seen. Gently move the spectrometer from side to side
to maximize the intensity of the light from the slit. This is the m  0 maximum for all
wavelengths and is the same color as the directly viewed source.

3. Adjust the black eyepiece (in and out) until the crosshairs are in focus.

4. Adjust the knurled silver ring in back of the eyepiece until the illuminated slit is in
focus. The slit should be centered vertically.

I. Calibrate the grating: Before the spectrometer can be used to measure unknown
wavelengths, you must determine the value of the grating spacing d with a known
wavelength. For this purpose you will use the green line from the Hg source, which has a
wavelength of 546.1 nm.

With the grating in the holder, view the Hg source directly and then notice several
colored lines to either side of the center telescope position (  0). Locate the first-order
(m  1) bright green line on both sides of center, and record the corresponding angular
separation with the uncertainty. The angle 1 in Eq. (3) will then be one-half of this
angular separation.

From Eq. (3) and the value of 1, compute the average slit spacing d for your grating,
along with its uncertainty. Compare this with the value calculated from the
manufacturer’s specification of 7500 lines/in (1 in  2.540 cm).

II. Measure the wavelengths of the hydrogen spectrum: Replace the mercury discharge
tube with a hydrogen tube and confirm that the collimator looks directly at the vertical
center of the tube. Use the grating spectrometer to measure 1 for the visible lines in first
order and record your data in a table like the one below.
d  d 

47
Electron Diffraction Physics 204C Laboratory

Line ID (color)       (meas) Ephoton  (Bohr) AGREE?

Compute the wavelengths from Eq. (3) using your calibration value of d and enter them
in the chart, along with their uncertainties.

Use Eqs. (1) and (2) to compute the visible photon energies and wavelengths predicted by
the Bohr model that correspond to transitions of the electron to the n  2 state from n  2,
and enter them in your table alongside the values obtained from the spectroscope.

Finally, decide, for each of your measured wavelengths, whether there is agreement
(considering uncertainties) with those predicted by the Bohr model.

Draw an energy level diagram like the one at the right in Fig. 1 and label the transitions
with wavelengths and photon energies.

Uncertainties.

For the calibration, if we take the Hg green wavelength to be known exactly, then the
fractional uncertainty in the spacing d for the slits on the grating will be approximately
d
  cot   ,
d
where  is in radians.

For the measurement of wavelengths, the uncertainty  should take into account both
the uncertainty in d (above) and the uncertainty  (in radians):
2
  d 
   cot    .
2
 
  d 
.
Conclusion.

Based on your observations and measurements, does the Bohr model adequately predict
the wavelengths of visible light emitted by the hydrogen atom?

48
Uncertainties Physics 204C Laboratory

Matter Waves: Electron Diffraction

Introduction. In this experiment you will observe the wavelike properties of electrons as
evidenced by their diffraction by a polycrystalline lattice of atoms. From these
observations you will determine the spacings that characterize the regular periodic
arrangements of atoms in a thin layer of carbon in the form of graphite.

h
Consider the scattering of electrons (with de Broglie wavelength   ) from a crystal
p
lattice of atoms arranged in planes spaced a distance d apart, as illustrated below.

Figure 1
A beam of electrons incident on the crystal at an angle  (measured with respect to the
rows of atoms) will be strongly reflected at the same angle  if this angle is also one at
which constructive interference occurs. This happens when the path length difference
(PLD) between waves reflected from successive rows of atoms is an integer multiple (m)
of the wavelength . Geometrically, this PLD can be seen to be 2d sin  , so the condition
on  for strong reflections is
m  2d sin  , (1)
where m is a positive integer. This is known as Bragg’s law and was originally applied to
x-ray diffraction. Using your apparatus, you will not measure , but rather  , shown in
Fig. 1 as the angle between the reflected beam and the unscattered incident beam. Since
  2 and you will confine your observation to first-order (m  1) reflections, Eq. (1)
reads
  
  2d sin   . (2)
2
Since it will be possible to independently control  by adjusting the speed of the
  
electrons, plotting  vs. sin   should yield a straight line with slope 2d, twice the
2
spacing between atomic rows (actually planes).

The Graphite Target: The structure of the graphite in


your target is a bit more complex than that depicted by
Fig. 1. The carbon atoms are arranged in a layered

© 2011 Eric
Figure 2 R. Dietz
Electron Diffraction Physics 204C Laboratory

hexagonal lattice as pictured in Fig. 2 at the right. In this structure, there are two regular
interatomic spacings d that will correspond to two different diffraction maxima. Because
the graphite is in the form of randomly oriented micro-crystallites deposited on the nickel
grid, the diffracted beam will emerge in all directions consistent with Eq. (2), thus
resulting in rings (as shown in Fig. 3) instead of dots35. Note that the larger d-spacing
(d2) corresponds to the smaller (inner) diffraction ring and the smaller spacing (d1)
corresponds to the outer ring.36

The Apparatus: The apparatus you will use to observe electron diffraction, shown
schematically below in Fig. 3, comprises an electron “gun” that emits a narrow beam of
electrons (that have been accelerated by a variable high voltage Va) within an evacuated
clear glass bulb, on the front surface of which is deposited a luminescent screen.

Figure 3

The electrons are diffracted by the target at two angles  with respect to the incident
beam, corresponding to the two different spacings of the atoms in the graphite lattice.
The diffracted beam strikes the glass a distance L from the target, forming a set of two
glowing rings (diameters Din and Dout) which can then be measured to deduce the values
of  , and then the interatomic spacings d.

Electron wavelength in terms of Va: The voltage Va through which the electrons are
accelerated determines the speed of the electrons hitting the target. The speed, in turn,
determines the electron momentum and hence the wavelength. The wavelength is given
h h
by    , where v is given from the Work/Energy theorem in terms of Va:
p me v
2Va e h
U  K  Wext  v  . This gives the wavelength as   Va  2 .
1

  
V e 1
a
2
me v 0
me 2 e me
2

Putting in the values of the constants gives the working expression for  in nm in terms
of Va (in volts):

35
This is referred to by crystallographers as a “powder pattern,” in contrast to that produced by a single
crystalline array like that pictured in Fig. 1.
36
This is analogous to what we observed in connection with the diffraction of light: narrower slits and
smaller slit spacings correspond to wider patterns, and vice versa.

50
Electron Diffraction Physics 204C Laboratory

1.226
 (nm)  . (3)
Va

Note that for any wavelength, the values of  for the rings are deduced from their
D
measured diameters Din and Dout: tan   . The length L is given by the manufacturer
2L
as L  0.140  0.003 m.

Procedure. You will need the following:

 electron diffraction apparatus  Vernier calipers


 computer/LoggerPro  desk lamp

The apparatus should already be wired for operation. Do not alter the wiring or touch the
glass envelope unless specifically asked to do so by the instructor.

Preparation: With the slider set to the minimum position, turn on the high voltage
supply, which supplies the accelerating voltage Va as well as the heater voltage for the
cathode from which the electrons are emitted. Allow one minute for the heater to warm
up so that the operating conditions are stabilized. An external bias (low-voltage) supply
is provided to control the focus and intensity of the electron beam. This control should
be adjusted for optimal viewing of the diffraction rings, but the current (monitored by the
DMM set to read DC mA) should never be allowed to exceed 0.15 mA! Note in this
connection that increasing the external bias voltage (by turning the knob on the little box
clockwise) decreases the current. Note also that the voltage and current controls are not,
unfortunately, completely independent of each other.

Take the data: For six values37 of the high voltage Va (spread out as much as possible
over the usable range from 05000 V), measure the diameters Din and Dout of the
diffraction rings on the luminescent surface on the front of the tube. Use the Vernier
calipers38 for this. Since the rings are actually thick, fuzzy annuli, take the diameters as
averages of the inner and outer diameter of each annulus.

37
The voltage is read on the top scale of the meter on the high voltage supply when it is set to output
voltages in the range 05000 V.
38
If you are unsure about how to read the calipers, Appendix C may help.

51
Electron Diffraction Physics 204C Laboratory

Tabulate the results for Va, Din, and Dout in your notebook, using a table like the one
below:
L  0.140  0.003 m (manufacturer’s value)

Va (Volts) Din (m) Dout (m) 1.226 D  D 


 ( ) (nm) in ( tan 1  in ) out ( tan 1  out )
Va  2L   2L 

Complete the table by computing  for each voltage setting Va, along with the values of
 for each set of rings for which you measured the diameters.

Use Eq. (2) to determine the interatomic spacings d for your graphite target by plotting 
  
vs. sin   for the inner rings and for the outer rings (two plots) in LoggerPro39. Note
2
that the slope of each line will be 2d, i.e., twice the corresponding spacing. Sketch both
graphs on the same set of axes.

Determine the two values of d  d for your sample. The uncertainties d can be taken
as half of the uncertainties of the slopes given by LoggerPro. Compare your results to
the established values of these spacings: 0.213 nm and 0.123 nm.
.
Conclusion.

Are your measured interatomic spacings for graphite consistent with the established
values given above?

In what way do your observations provide evidence of the wave nature of particles?
What do you think your luminescent screen would look like if electrons behaved only as
particles?

39
You can use the “calculated column” feature in LoggerPro to compute the plotted quantities, but make
sure your angle units correspond to the setting in the dialog box in FileSettings for…

52
Uncertainties Physics 204C Laboratory

The Half-life of Radioactive Indium 116

Introduction. You will use a Geiger counter to determine the half-life of indium-116
(116
49 In ) , which does not occur naturally. Just before the measurement, we produce In-116
nuclei by bombarding an indium foil, which is 95.7% In-115 (and 4.3% In-113), with
neutrons--a process referred to as neutron activation.

The Neutron Source. The neutron source, comprising a mixture of americium-241 and
beryllium-9, is kept in the center of a large shielded cylindrical tank. Americium-241 has
a half-life of 458 years and decays by alpha emission. The equation for this reaction is:
241
95 Am 237
93 Np  2  .
4

The 5.6 MeV alpha particles which are emitted penetrate into the nucleus of beryllium-9,
forming a compound nucleus which is vibrating so violently that it immediately breaks
up into a neutron and a stable nucleus. The equation for this reaction, which is our source
of neutrons, is:

9
4 Be  42 126 C  01 n .

Decay of In-116: Naturally occurring indium is 95.7% indium-115, with a half-life of


5.1 x 1014 years. This is such a long half-life that essentially none of this indium is
decaying, so we measure almost no activity above normal background when we place a
foil of indium which has not been irradiated by neutrons under our Geiger counter.

When we irradiate our indium foil with neutrons, some of the neutrons are captured by
the In-115. The equation for this neutron activation reaction is:

115
49 In  01 n116
49 In
.

This new isotope of indium is radioactive and has a half-life which you will measure. It
decays by beta (e) emission which includes the ejection of an antineutrino (  ) also:
116
49 In116
50 Sn  1    .
0

Note that by irradiating indium in the neutron source, we change a very tiny fraction of
the indium atoms in the foil into tin.

The rate (R) with which a radioactive isotope decays by means of a single process is
given quite generally as a function of time by

R ( t )  R 0 e  t ,

© 2011 Eric R. Dietz


Half-life of In-116 Physics 204C Laboratory

where R0 is the rate of decay at time t  0 (when the first count rate is measured), and  is
the decay constant of the isotope. We can see how the decay constant may be determined
from a plot of n  R  vs t by taking the natural log of both sides of the above expression
for R(t):

 n  R   n  R 0    t

The slope of the n R vs t plot is thus .

The decay constant is related to the half-life T1 :


2

0.693
T1  .
2

Procedure. You will need the following:

 computer/LoggerPro  12 In foils (front bench)


 Geiger counter/ratemeter  stopwatch

To quantitatively assess the effect of the background radiation, you will need to make
several counts of the background radiation (i.e., with no sample in the chamber). Take
five 60-s counts of the background; compute and record the average background count
rate.

Immediately after the instructor gives you some irradiated In foils, insert the sample into
the counting chamber as close as possible to the detector and begin to record the
activity40 R (in counts per minute) vs. time by counting for 60 s during every 5th minute
for a total time of one hour. Tabulate these values of R in your notebook along with time
t (taken in the middle of the count period).

Enter the values of R (corrected by subtracting the average background rate) and t in
LoggerPro data columns and create a calculated column that gives n  R  . Display and
sketch the best-fit line for a plot of n  R  vs. t and record its slope, which should be ,
along with its uncertainty.

Produce a value of the lifetime of 116 In : T1/2  T1/2 and compare this with the
established value, which is 54.2 min.

Uncertainties.

40
Note that we have neglected slight corrections which could be made by taking into account the
"deadtime" of the Geiger counter.

54
Half-life of In-116 Physics 204C Laboratory

Once again, LoggerPro calculates the uncertainty in the slope for you; this is . The
fractional uncertainty in the half-life is the same as that for the decay constant .

Conclusion.
116
How does your measurement of the half-life of In compare with the established value?

Describe, in your own words, how the decay rate of a radioactive isotope varies with
time. What are some other processes that behave in a similar way?

55
Uncertainties Physics 204C Laboratory

Appendix A: Measurements and Uncertainties


The establishment or verification of a physical law or the experimental determination
of a physical quantity usually involves measurement. A reading taken from the scale on a
stopwatch, a meter stick or a voltmeter, for example, may be directly related by a chain of
analysis to the quantity or law under study; any uncertainty in these readings would thus
be directly reflected in an uncertainty in the final result. A measurement by itself,
without at least a rough quantitative statement as to the uncertainty involved, produces a
result of limited usefulness. It is therefore essential that any introductory physics
laboratory course include a discussion of the nature of experimental uncertainties and the
manner in which they should be assigned to experimental results.

1. UNCERTAINTY vs. DISCREPANCY

When you report the result of a measurement of a quantity x, you should also give the
uncertainty x, e.g.,

5.0 m ± 0.1 m
 
x x

The uncertainty tells you how precise you think your measurement is. It is also often
useful to compare your result with a "true" or accepted value; the difference between
these is the discrepancy and is a reflection of the overall accuracy of the measurement.

An estimate of the uncertainty of a measurement should always be made; a calculation


of the discrepancy can be made only if an accepted value or result happens to be available
for comparison. The conclusion for an experiment should, whenever possible, address
the question: Do the uncertainties involved in my measurements account for the
discrepancies between my result and the accepted one?

As an example, suppose you do an experiment to measure the acceleration of


m
gravity and you report the experimental value ( gexp  g exp ) to be 9.75  0.08 2 where
s
m
the “accepted” value is gacc  9.81 2 . As you can see from the graphic representation
s
below, the uncertainty g exp in the measurement accounts nicely for the discrepancy
between gexp and gacc.

gexp g acc
 ]
9.81
9.75

gexp

© 2011 Eric R. Dietz


Uncertainties Physics 204C Laboratory

2. ORIGINS OF UNCERTAINTIES

Problems which lead to discrepancies between experimental values and "true" values
can be placed in two general categories:

I. Systematic Errors are inaccuracies due to identifiable causes and can, at least in
principle, be eliminated. Errors of this kind result in values for the measured
quantity which are consistently either too high or too low. Such errors can be

a) Theoretical - due to simplifications of the model system or approximations in


the equations describing it.

b) Instrumental - e.g., a poorly calibrated instrument.

c) Environmental - e.g., factors such as inadequately controlled temperature and


pressure.

d) Observational - e.g., parallax in reading a meter scale.

II. Random Uncertainties are the result of small fluctuating disturbances which cause
about half the measurements of any quantity to be too high and half to be too low.
It is often not possible in practice to identify all the sources of such errors, to
evaluate their effects individually or to completely eliminate them. They can,
however, often be well characterized mathematically.

To illustrate the difference between systematic errors and random uncertainties,


we consider the measurement of the length of time Tf taken for a ball to fall some fixed
distance, say 10 m. Suppose we drop the ball about 1000 times and obtain as many
values of Tf, rounding each time to the nearest .01 s. If N(Tf) is a function of Tf that
represents the number of times a particular measured value of Tf occurs, then a histogram
of N versus Tf might look like

The "best" value for Tf is just the average (mean) value Tf :

57
Uncertainties Physics 204C Laboratory

N T i fi
Tf = i=1
(1)
n
where Ni is the number of times the value Tfi appears as a measurement and n is the total
number of times the ball is dropped. In this example we see that Tf = 1.43 s; the
discrepancy between this number and the "true" value of Tf, if known, would provide a
measure of the accuracy of the measurement. Systematic errors (e.g. the reaction time
involved in starting and stopping the clock) will affect Tf and hence the accuracy. The
spread of Tf values, indicated by the width of the curve, is a reflection of the precision
with which Tf is being measured. Random uncertainties contribute to this width and may
be attributable, to e.g., small fluctuations of the height from which the ball was dropped
or to the difficulty in determining the exact moment of impact with the ground. In what
follows we discuss the mathematical treatment of such random uncertainties and the role
they play in reporting laboratory measurements.

3. CHARACTERIZING A SET OF DATA: THE NORMAL DISTRIBUTION

It is most often the case that repeated measurements of the same quantity will, as
in the timing experiment example described above, exhibit a spread about the average
value related to the random errors associated with the measurement process. If we make
"many" (say 106) measurements of a quantity x and plot the frequency of occurrence
N(x), we quite often obtain a curve that approximates a Gaussian, or normal distribution,
as pictured below.

This curve N(x) represents the relative probability with which values of x are obtained as
the result of any single measurement of this quantity, which may be, for example, the
reading of a meter or stopwatch. The analytical expression for such a curve is
2
N 0 - (x 2)
N(x) = e 2 (2)
2

where the parameters  and  determine the position and width of the peak, respectively.
The "normalization" parameter NO would correspond, in our timing example, to the total
number of readings taken from the stopwatch.

58
Uncertainties Physics 204C Laboratory

The curve representing Eq. (2) is of importance for the analysis of experimental
data because in many cases this data is distributed normally and thus has a frequency
distribution that can be "fit" to this curve. The more data taken, the better the fit will be.
But for any set of data, regardless of the number of data points or the nature of the
distribution, we can define quantities which characterize the "best value" and "spread" of
the data.

For a set of data points xi, the quantity x is the mean of all values obtained for x,
and is defined by
n

x i
x1  x 2  x n
x i 1
 (3)
n n
where n is the number of measurements. For data in which each value of xi generally
occurs more than once, Eq. (1) may provide a more convenient way to calculate the mean
value of x. The "best value" that one finally reports for the quantity x is generally its
mean, as defined above. If a large number of normally distributed data points are taken,
then _x should be close to the value of  for which the curve representing Eq. (2) best fits
the data.

We can characterize the spread of any finite set of n data points by its standard
deviation, symbolized by s.d., which we will define as
n

 (x i  x )2
s. d.  i 1
(4)
n
The s.d. for a set of data thus represents the square root of the average square of the
deviation (i.e., the "rms" deviation) that the data points exhibit with respect to their
average value. For a large number of normally distributed data points, s.d.  , where 
is the value of the parameter of Eq. (2) which produces the best description of the data.

4. REPORTING THE VALUE OF A MEASURED QUANTITY

Because repeated measurements of the same quantity give us, as discussed above,
a range of values reflecting the random fluctuations inherent in the measurement process,
we need to report the result of these measurements in a way which reveals not only the
best value of the quantity, but also the precision to which we are stating it: we report both
the best value and the uncertainty. The distributions discussed above refer to a collection
of single measurements. So if we make a single measurement of a quantity x and had
some knowledge of the single measurement distribution curve, then we could report this
value of x as

x  s. d. (5)
where x is the single measured value and s.d. is the standard deviation of the single
measurement distribution. To determine an appropriate value for s.d. without actually
making several measurements, we would have to know something, a priori, about the

59
Uncertainties Physics 204C Laboratory

single measurement curve; we would either have to be given a value of s.d. or we would
have to guess.

Suppose, as is more common, we take n measurements instead of just one.


Intuitively, it seems reasonable that the mean value should be reported with a smaller
uncertainty than the single measurement discussed above. In fact, if one took many sets
of n measurements (all with the same s.d.) and then calculated the mean for each set, the
distribution of these means would have a smaller standard deviation (denoted S.D.) than
that for a single measurement (s.d.). It is not hard to show (from Eq. (8a) below) that
s.d.
S.D.  (6)
n
What this means is that one could report the result of a set of n measurements as
s. d.
x  S. D.  x  (7)
n
so that the uncertainty associated with the average of 5 measurements, e.g., could be
reduced by about a factor of 2 from the single measurement s.d. NOTE: The expression
above for reporting an average is only meant to serve as a guideline in reporting results; it
assumes that all uncertainties involved are purely random in nature. Systematic errors
must always, at some level, be considered, so that in most cases we cannot create an
arbitrarily small uncertainty simply by taking more and more measurements!

Relative vs. Absolute Uncertainty. One often reports the relative (or percentage)
uncertainty associated with the best value in lieu of the absolute uncertainty x = s.d. or
S.D. No units are associated with these relative uncertainties, so that these may be used
to directly compare the precision of two different measurements. On the other hand, the
absolute uncertainty is more convenient for comparing two measured values.

Example. Suppose we make 10 measurements of the period T of a pendulum, and that


the results are tabulated to the nearest millisecond (ms) as
TRIAL # T (s) TRIAL # T (s)

1 1.411 6 1.468
2 1.452 7 1.437
3 1.403 8 1.446
4 1.414 9 1.425
5 1.459 10 1.434

We need to report the best value for T along with the uncertainty, which we assume is
due to random fluctuations. For the best value, we have the mean,
10

T i
T i 1
 1435
. s.
10
To estimate the uncertainty, we calculate s.d. from Eq. (4):

60
Uncertainties Physics 204C Laboratory

10

 (T  T )
i
2

s. d.  i 1
 0.02 s.
10
But since we made 10 measurements, we can report the uncertainty as
s.d.
 T  S.D.   0.006 s.
10
Our result for T thus appears as

T = 1.435 s ± 0.006 s (with absolute uncertainty)


or
T = 1.435 s ± 0.5 % (with relative uncertainty)

Significant Figures. A decision must always be made, consciously or otherwise, as to


how many significant figures should be used to report a quantity x and its uncertainty x.
Since uncertainties are merely estimates, it is seldom necessary to report x to more than
1 significant figure! The number of significant figures with which you report x is often
taken to imply its uncertainty, therefore this number must be consistent with the explicitly
stated uncertainty x.

Rule of thumb for significant figures: The measurement and the uncertainty
should have their last digits in the same location relative to the decimal point. Example:
(1.06 ± 0.01)  103 m.

Thus, a length stated as 1.0 m ± 0.002 m would present a problem because the
stated uncertainty would imply that the observer is withholding information by not
reporting more decimal places in the length. A result stated as 1.06 m  0.2 m reports a
length with too much accuracy, given the uncertainty. If x and x imply different
uncertainties, one of them should be adjusted so that they are both consistent with the
largest of the two uncertainties.

Guesstimating Uncertainties. The uncertainty to assign a given single measurement of


x depends on the technique employed in the measurement, the quality of the equipment
used, and the care with which the measurement is made. The number you come up with
as x for a particular measurement is often quite subjective, because only you know what
happened during your measurement. There are a couple of options for estimating x:

1. Take several measurements and calculate s.d., as in the example above for T;

2. Estimate x from the nature of the measurement. For example, when reading a
scale, the uncertainty is sometimes, but not always, half of the smallest scale
division, but you will encounter situations when the scale divisions have nothing
to do with x. In other words, you will always have to use your judgment.

You will find yourself using option #2 quite often; it's a real timesaver.

61
Uncertainties Physics 204C Laboratory

5. PROPAGATION OF UNCERTAINTIES

Quite often the quantity of interest is measured indirectly, i.e., derived from one
or more quantities which are each measured directly and reported with uncertainties, as
discussed above. In these cases, the estimate of the uncertainty in the indirectly measured
quantity must reflect the uncertainties associated with all of the relevant directly
measured ones.

For example, suppose we wish to infer the acceleration of gravity (g) from direct
measurement of both Tf (the time taken for an object to fall) and of L (the height from
which it is dropped). The equations of kinematics tell us that g can be calculated from Tf
and L:
2L
g = 2
.
Tf
How, then, do the uncertainties in L and Tf contribute to g, the uncertainty in g?

You will need to know the answer to this question, and to others just like it. We
first give the general expression for propagating the uncertainties in directly measured
quantities to find the uncertainty in the derived result. We then treat two special cases
which occur quite often.

General Case: Let z be a quantity to be derived from the direct measurement of two
independent quantities x and y by using the relationship z = f(x,y). Suppose we were to
make small definite (known) errors x and y in the measurements of x and y respec-
tively. These would show up as an error z in z that would be given by
f f
z  x  y .
x y
But because x and y are actually random uncertainties, a more realistic estimate of the
uncertainty z is
FG f xIJ  FG f yIJ
2 2

z 
H x K H y K , (8)

as can be demonstrated from considerations of probability theory. Given uncertainties in


the two quantities x and y, Eq. (8) can be used to estimate the propagated uncertainty
z. The generalization to more than two directly measured quantities is straightforward.

Origin of Eq. (8) (Optional): Consider two independent random variables P and Q, each with an average of
zero and with s.d.'s of P and Q, respectively. The s.d. of the sum P + Q (denoted (P+Q)) is, according
to the definition of Eq. (4),

(P+ Q) = (P+ Q )2

If we compute the square of this quantity in terms of P and Q, we get

62
Uncertainties Physics 204C Laboratory

( (P+ Q) )2 = (P+ Q )2 = P2 + 2 PQ+ P2


= P2 + Q2 = (  P )2 + (  Q )2 ,

from which the s.d. for P + Q is


(P+ Q) = (  P )2 + (  Q )2 .
Now considering the function z = f(x,y): the uncertainty in z (z), will be due to the combined effects of
the uncertainty in x (x) and the uncertainty in y (y). If P and Q represent the effect on z of changes in x
and y, respectively, then P and Q, the s.d.'s for P and Q, will be given by
f f
P =  x; Q =  y,
x y
where x and y are the s.d.'s, or uncertainties, in x and y. The spread, or uncertainty z in z will then be
(P+Q), given above. Substitution for P and Q results directly in Eq. (8) for z. It should be noted that
this same argument can also be applied to derive Eq. (6) above for the uncertainty (S.D.) in the average.

Many times, it is not necessary to use Eq. (8) directly because the form of the
expression for z will fit one of two special cases.

Special Case I: The quantity of interest (z) is given in terms of two directly measured
quantities x and y, each with their respective uncertainties x and y, as z = Ax + By. In
this case, the absolute uncertainty in z is given by

z  (A x)2  (B y)2   (8a)

where the "   " indicates that this expression is easily generalized to the case where z is
given as the sum of three or more quantities.

Example: We wish to determine the mass of water, mw, contained in a beaker from
independent measurements of the mass of the beaker alone, mb, and the combined mass
of the beaker + water, mt. Then mw = mt - mb and the uncertainty in the derived quantity
mw is
m w  ( m t )2  ( m b )2 ,
where mt and mb must either be estimated or determined from repeated measurements
of these quantities.

Special Case II: The derived quantity (z) is the product of powers of measured quantities
x and y: z = K xa yb, where K is a constant. In this case we can give the fractional
(relative) uncertainty in z as

z
=
LMa  x OP + LMb  y OP
2 2

(8b)
z N xQ N yQ
where we can again generalize to any number of quantities, as indicated for Eq. (8a).

63
Uncertainties Physics 204C Laboratory

Example: If the acceleration of gravity g is given, as discussed above, in terms of L and


Tf as g = 2L/Tf2, what is g in terms of the uncertainties L and Tf? If g is rewritten as
g = 2(L1)(Tf-2), then we have something that fits the form of special case II with x = L,
a = 1 and y = Tf, b = -2. So for the relative uncertainty g/g we get
g
=
LM  L OP + 4 LM  T OP
2
f
2

.
g NLQ N TQ
Note that because g depends on the second power of Tf but on only the first power of L,
the uncertainty in g is much more "sensitive" to the uncertainty in Tf than to the
uncertainty in L.

If neither of the special cases applies to the quantity being investigated, it is some-
times possible to derive the uncertainty by applying Eq. (8a) and (8b) in combination;
otherwise, use the general expression of Eq. (8) directly.

Graphs and Uncertainties. Physical laws often predict that two quantities (x and y) that
can be measured (directly or indirectly) are proportional to each other. For example,
Newton's second law in one dimension (F = Ma) predicts that the acceleration (a) of an
object is proportional to the force F (the mass M being the constant of proportionality);
Ohm's law (V  IR) predicts, for certain materials, that the voltage V and current I are
proportional. It is often the constant of proportionality between these measured
quantities (M or R in the above examples) that is to be determined by the experiment,
rather than just individual pairs of values for x and y. The linear relationship between x
and y can be expressed as

slope 

y  mx  b

y-intercept

If y is plotted vs. x, as in graph (a) below, then the result should be a line with a slope m,
calculated as y/x. (Note that here y and x refer to the rise and run, respectively, not
to the uncertainties in y and x!).
y
m (best fit) = x
y y m+
_
m
__
y (x,y)
x

Of course,
exper- x
iments
being what
x x
they are, the data
points (a) (b) (as indi-

64
Uncertainties Physics 204C Laboratory

cated in graph (a)) will not all be on a single line; it is up to you to draw the "best" line
that "fits" the data so that the slope m (or y-intercept b) may be determined. There is a
well-known method for determining the best line for a given set of points (x,y) -- the least
squares fit (see below). For the purposes of this lab, however, it will often be sufficient
to use a straightedge and your eyes to draw the line which represents the best
compromise, i.e., "comes as close to as many points as possible."

NOTE: If the slope or the intercept is to be measured directly from the graph, it is
most often advantageous to use as much of the page as possible to draw the graph!

If the slope m is the quantity to be reported as the result of the experiment, then
the uncertainty m must be reported along with it: we need m ± m. The uncertainty m
can be thought of as arising from the uncertainties in the position of each plotted point on
the graph, since each y value may be characterized by an uncertainty y. One way of
representing these uncertainties is by means of the "error bars" drawn for each point on
graph (b) above. As with the slope m, there is a well known (and somewhat complex)
expression that can be used to determine m, but for our purposes it is sufficient to "eye
ball" the range of slopes. As indicated in graph (b), this can be done by using the error
bars to estimate m+, the highest possible slope for a line that passes through or near most
of the error bars, and m-, the lowest slope. The uncertainty m is then very roughly
given by
m  m
m  .
2

To estimate the slope (m) and its uncertainty (m) by eye:

1. Plot the data points (x,y) on graph paper. Use as much of the page as possible.

2. Place error bars through each point to indicate roughly the extent to which the
location of that point on the graph is uncertain. The size of these bars can be
estimated from the discussion under the "uncertainty" section in each experiment.
An alternative to bars is to enlarge your dot to indicate its "fuzziness." If the
uncertainty is too small to show up on your graph, state this in the report.

3. Plot the point ( x, y) on your graph (where x and y are the means of the x and y
values, respectively). It so happens that this point always lies on the "best line"
defined below by the least squares fit procedure.

4. Using a straight edge, draw the line passing through ( x, y) that "comes closest" to
rise
the data points. Measure the slope m as , using as much of the line as
run
possible and paying attention to the units on the axis scales.

65
Uncertainties Physics 204C Laboratory

5. Estimate the uncertainty m by drawing 2 dotted lines through ( x, y) and passing


within most of the error bars--one with the minimum slope, m, and one with the
maximum slope, m. Estimate m from the expression in the paragraph above.

Least Squares Fit (Linear Regression): Though somewhat more involved and less
intuitive than the graphical method described above for finding the slope and intercept of
a best fit line, the analytic method of linear regression finds the best fit by minimizing the
sum of squares of deviations of y values from the fitted straight line. For this reason, it is
called the method of least squares. Given N data points x1, y1; x2, y2; ...; xN, yN, the slope
m and y-intercept b of the best fit line (given by y = mx + b) are calculated as:
N

(x  x)(y  y)
m= i=1
; b = y  mx, (9)
N x 2

where x and y are the means of the x and y values respectively, given by Eq. (3), and x
is the s.d. of the x values as given by Eq. (4).

As an alternative to the graphical method described above for estimating the


uncertainty m for the slope, you may use the least squares methodology to estimate this
quantity along with the uncertainty in the intercept, b. The expressions for these are
N

1 d 2

m  i 1
; b  m  x 2  x 2 , (10)
x N( N  2)

in which di  yi  (mxi + b) is the y deviation of the point (xi,yi) from the best fit straight
line y = mx + b. If your calculator does linear regression calculations which result in the
least squares fit slope m and correlation coefficient r, then it may be simpler to calculate
the uncertainty in the slope (m) with the equation*

tan  arccos  r   m 1  r2
m  m  . (11)
N2 r N2

References and further reading:

1. Lichten, William, Data and Error Analysis in the Introductory Physics Laboratory
(Allyn and Bacon, Inc.)

*
J. Higbie, American Journal of Physics, 59(2) 1991

66
Physics 204C Laboratory

Appendix B: Meter and Oscilloscope Uncertainties

Fluke 179: The rightmost column of the table below indicates the uncertainties to be
associated with the readings of the Fluke 179 multimeter. The term [Counts] in the
expression for the “Accuracy” at the top refers to the value of the least significant figure
displayed, which is also the entry in the “Resolution” column. Example: For a reading of
V  1.500 V (on the 6.000 V range), V  (0.09% of 1.500 V)  2(0.001 V)  0.003 V.
So record V  1.500  0.003 V.

67
Scope/Meter Uncertainties Physics 204C Laboratory

SENCORE LC103:
Uncertainties for capacitance measurements:
For C  1990 F: (1%  1 pF  1 digit41)
For 2000 F  C  19.99 F: (5%  0.1% of range full scale42)

Uncertainties for inductance measurements: (2%  1 digit  0.1 H)

Resolutions and Ranges (right column) for Capacitance Measurements with ReZolver

41
The term “1 digit” refers to the change in the value when the least significant figure changes by 1. This is the same
as the “resolution” shown above with the full scale readings.
42
See above for the full scale values corresponding to your readings.

68
Physics 204C Laboratory

Appendix C: Reading a Vernier Scale

The VERNIER Scale


A Vernier scale is a small, moveable scale placed next to the main scale of a measuring
instrument. It is named after its inventor, Pierre Vernier (1580 - 1637). It allows us to make
measurements to a precision of a small fraction of the smallest division on the main scale of the
instrument. (In the first example below the "small fraction" is one tenth.) Vernier scales are
found on many instruments, for example, spectroscopes, supports for astronomical telescopes
etc. One specific example, the Vernier caliper, is considered below.
Using a Vernier Scale
Figure 1 shows a Vernier scale reading zero. Notice that 10 divisions of the Vernier scale have
the same length as 9 divisions of the main scale.

figure 1

In the following examples we will assume that the smallest division on the main scale is 1mm so
the divisions on the Vernier scale are 0·9mm each. The position of the zero of the Vernier scale
tells us the number of cm and mm in our measurement. For example, in figure 2, the reading is a
little over 1·2cm.

figure 2

To find a more precise reading, consider figure 3 (which is a magnified view of part of figure 2).

figure 3

69
App C: Vernier Scale Physics 204C Laboratory

We are, in effect, trying to find the distance, x.


To find x, find the mark on the Vernier scale which most nearly coincides with a mark on the
main scale. In figure 3 it is obviously the third mark.
Now, it is clear that ............x = d - d’
Remembering that each division on the main scale is 1mm and that each division on the Vernier
scale is 0·9mm, we have:
x = 3mm - 3(0·9)mm = 3(0·1)mm
Therefore, the reading in the example is: 1·23cm
Similarly, if it had been, for example, the seventh mark on the Vernier scale which had been
exactly opposite a mark on the main scale, the reading would be: 1·27cm
Hence, the level of precision of an instrument which has a Vernier scale depends on the
difference between the size of the smallest division on the main scale and the size of the smallest
division on the Vernier scale.
In the example above, this difference is 0·1mm so measurements made using this instrument
should be stated as: reading ±0·1mm.
Another instrument might have a scale like the one shown in figure 4.

figure 4

Therefore, the precision is: 1mm - (49/50)mm = (1/50)mm = 0·02mm.


Results of measurements made using this instrument should therefore be stated as: reading
±0·02mm.
This principle is used in the Vernier caliper shown below.

70
App C: Vernier Scale Physics 204C Laboratory

The diagrams below illustrate how to use a Vernier caliper to measure:


A. the internal diameter of a hollow cylinder
B. the external dimensions of an object
C. the depth of a hole in a piece of metal.

A. B.

C.

71
App D: Hardware Spec Sheets Physics 204C Laboratory

Appendix D: Vernier Hardware Spec Sheets

72
App D: Hardware Spec Sheets Physics 204C Laboratory

73
App D: Hardware Spec Sheets Physics 204C Laboratory

74
App D: Hardware Spec Sheets Physics 204C Laboratory

75
App D: Hardware Spec Sheets Physics 204C Laboratory

76
App D: Hardware Spec Sheets Physics 204C Laboratory

77
App D: Hardware Spec Sheets Physics 204C Laboratory

78
App D: Hardware Spec Sheets Physics 204C Laboratory

79
App D: Hardware Spec Sheets Physics 204C Laboratory

80
App D: Hardware Spec Sheets Physics 204C Laboratory

81
App D: Hardware Spec Sheets Physics 204C Laboratory

82
App D: Hardware Spec Sheets Physics 204C Laboratory

83
App D: Hardware Spec Sheets Physics 204C Laboratory

84
App D: Hardware Spec Sheets Physics 204C Laboratory

85
App D: Hardware Spec Sheets Physics 204C Laboratory

86

Das könnte Ihnen auch gefallen