10.10.2014 Views

Bibliography - School of Physics - University of Melbourne

Bibliography - School of Physics - University of Melbourne

Bibliography - School of Physics - University of Melbourne

SHOW MORE
SHOW LESS

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

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

These notes were revised in 2006 by David Hoxley, incorporating<br />

suggestions from the 2005 demonstrating team. Thanks, guys.<br />

These notes were imported into L A TEXand substantially revised in 2005 by a<br />

team composed <strong>of</strong>:<br />

Simon Devitt<br />

John-Paul Goldby<br />

Chris Gurrie<br />

David Hoxley<br />

Steven Karataglidis<br />

Tracey Mackin<br />

Alaster Meehan<br />

Matthew Norman<br />

Tim Starling<br />

Melissa Wals<br />

The majority <strong>of</strong> the typesetting and revisions were done by Steven<br />

Karataglidis and Melissa Wals. Preparation <strong>of</strong> the final manuscript was done<br />

mostly by Melissa with red ink supplied by David Hoxley.<br />

The last major revision <strong>of</strong> the notes from which we draw heavily was in<br />

1999 with the assistance and hard work <strong>of</strong> the following people:<br />

Anton Barty, Chris Chantler, Jacinta den Besten, Phillip Fox, Amelia Liu,<br />

Leigh Morpeth, and Justine Tiller.


Contents<br />

1 Introduction 1<br />

1.1 Safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.2 Experimental techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.2.1 Computers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.3 The report . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.3.1 Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.3.2 Writing the report . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.3.3 Assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.4 Sample report . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.5 Hints and tips . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.6 Deadlines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.7 Illness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

I Optics 13<br />

2 2-slit interference with single photons 15<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.2 Background theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.3 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.4 Experiment - Laser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.4.1 Alignment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.4.2 Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.5 Experiment - Single Photon Counting . . . . . . . . . . . . . . . . . . . . 20<br />

i


ii<br />

CONTENTS<br />

3 The Michelson interferometer 23<br />

3.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.2.2 Background theory . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

3.2.3 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.2.4 Experimental work . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.2.5 Conditions <strong>of</strong> interference - polarisation . . . . . . . . . . . . . . . 27<br />

3.2.6 Conditions <strong>of</strong> interference - coherence . . . . . . . . . . . . . . . . 27<br />

3.2.7 Micrometer calibration . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

3.3 Fringe visibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

3.3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

3.3.2 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3.3.3 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3.4 Refractive index <strong>of</strong> air . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.4.2 Background theory: optical path length . . . . . . . . . . . . . . . 30<br />

3.4.3 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.4.4 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

4 Fraunh<strong>of</strong>er diffraction 33<br />

4.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

4.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

4.3 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

4.3.1 Fresnel or Fraunh<strong>of</strong>er? . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

4.3.2 Diffraction by a single slit . . . . . . . . . . . . . . . . . . . . . . 34<br />

4.4 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

4.4.1 Calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

4.4.2 Single slit diffraction . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

4.4.3 Babinet’s principle . . . . . . . . . . . . . . . . . . . . . . . . . . 39


CONTENTS<br />

iii<br />

4.4.4 Multiple slits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

4.5 Further work - circular and other diffracting objects . . . . . . . . . . . . . 42<br />

4.5.1 Overview and background theory . . . . . . . . . . . . . . . . . . 43<br />

4.5.2 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

4.6 Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

4.6.1 CCD camera operating instructions . . . . . . . . . . . . . . . . . 44<br />

4.6.2 How the CCD works . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

5 Holography 47<br />

5.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

5.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

5.3 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

5.4 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

5.4.1 Intensity <strong>of</strong> a wave . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

5.4.2 Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

5.4.3 Holograms and holography . . . . . . . . . . . . . . . . . . . . . . 49<br />

5.4.4 Movement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

5.5 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

5.5.1 Film handling and developing . . . . . . . . . . . . . . . . . . . . 51<br />

5.5.2 White light hologram . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

5.5.3 Transmission or <strong>of</strong>f-axis hologram . . . . . . . . . . . . . . . . . . 54<br />

5.5.4 Time-averaged interferometry <strong>of</strong> a resonating tube . . . . . . . . . 55<br />

5.6 Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

5.6.1 Spherical waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

5.6.2 Interference <strong>of</strong> two waves . . . . . . . . . . . . . . . . . . . . . . 56<br />

5.6.3 Transmission function . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

II General physics 59<br />

6 Magnets And magnetic fields 61


iv<br />

CONTENTS<br />

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

6.2 The Hall effect and magnetic field probes . . . . . . . . . . . . . . . . . . 61<br />

6.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

6.2.2 Background theory . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

6.3 The Hall effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

6.4 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

6.4.1 The semiconductor . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

6.4.2 The electromagnet . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

6.4.3 The Hall effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

6.4.4 The Hall probe . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

6.5 Ferromagnetic materials and AC fields . . . . . . . . . . . . . . . . . . . . 67<br />

6.5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

6.5.2 Background theory . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

6.5.3 The apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

6.5.4 Experimental work . . . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

6.6 Useful data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

7 Fundamental constants 75<br />

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

7.2 Charge to mass ratio <strong>of</strong> the electron . . . . . . . . . . . . . . . . . . . . . 75<br />

7.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

7.2.2 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

7.2.3 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

7.2.4 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

7.2.5 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

7.3 The photoelectric effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

7.3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

7.3.2 The failure <strong>of</strong> classical theory and Einstein’s postulate . . . . . . . 80<br />

7.3.3 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

7.4 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82


CONTENTS<br />

v<br />

7.4.1 Variation <strong>of</strong> stopping potential with frequency . . . . . . . . . . . . 84<br />

7.4.2 Calculation <strong>of</strong> the fundamental constants . . . . . . . . . . . . . . 84<br />

7.5 Useful data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

8 Waves in waveguides 87<br />

8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

8.2 Acoustics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

8.2.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

8.2.2 Stationary wave theory and the reflection <strong>of</strong> plane acoustic waves . 91<br />

8.2.3 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

8.2.4 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

8.3 Microwaves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

8.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

8.3.2 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

8.3.3 Modes in the guide . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

8.3.4 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

8.3.5 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

8.3.6 Polarisation <strong>of</strong> microwaves in the guide . . . . . . . . . . . . . . . 105<br />

8.3.7 Relation between guide and free space wavelengths . . . . . . . . . 106<br />

8.3.8 Impedance measurement . . . . . . . . . . . . . . . . . . . . . . . 106<br />

8.4 Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

8.4.1 The complex reflection coefficient K . . . . . . . . . . . . . . . . 107<br />

8.4.2 The Gunn diode . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109<br />

8.4.3 The standing wavemeter . . . . . . . . . . . . . . . . . . . . . . . 110<br />

8.4.4 Reading the micrometer . . . . . . . . . . . . . . . . . . . . . . . 110<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

9 Electron Spin Resonance 113<br />

9.1 A guide to background reading . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

9.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

9.3 Introduction: Angular Momentum in Quantum Mechanics . . . . . . . . . 114


vi<br />

CONTENTS<br />

9.3.1 Orbital Angular Momentum . . . . . . . . . . . . . . . . . . . . . 114<br />

9.3.2 Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

9.4 Electrons in an external magnetic field . . . . . . . . . . . . . . . . . . . . 115<br />

9.5 Resonance absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

9.6 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

9.6.1 Helmholtz coils . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

9.6.2 RF oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

9.6.3 Investigating the resonance . . . . . . . . . . . . . . . . . . . . . . 119<br />

9.6.4 Electron absorption . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

9.7 Extension: electron diffraction . . . . . . . . . . . . . . . . . . . . . . . . 120<br />

9.7.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120<br />

9.7.2 Basic theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

9.7.3 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

9.7.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124<br />

9.8 Useful data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

9.9 Wave function for the electron - a brief history . . . . . . . . . . . . . . . . 125<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126<br />

III Nuclear physics 127<br />

10 Rutherford scattering 129<br />

10.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129<br />

10.2 Introduction/background reading - importance <strong>of</strong> scattering . . . . . . . . . 130<br />

10.3 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

10.3.1 Cross section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

10.3.2 Solid angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132<br />

10.3.3 Differential cross section . . . . . . . . . . . . . . . . . . . . . . . 133<br />

10.3.4 Theoretical cross section . . . . . . . . . . . . . . . . . . . . . . . 134<br />

10.4 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

10.4.1 Safety precautions . . . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

10.4.2 Relationship between counts and source to detector distance . . . . 135


CONTENTS<br />

vii<br />

10.4.3 Rutherford scattering . . . . . . . . . . . . . . . . . . . . . . . . . 139<br />

10.5 Useful data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

11 β spectroscopy 143<br />

11.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

11.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

11.3 Background concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144<br />

11.4 Experiment - the β decay <strong>of</strong> promethium . . . . . . . . . . . . . . . . . . . 146<br />

11.4.1 Calibrating the magnet . . . . . . . . . . . . . . . . . . . . . . . . 146<br />

11.4.2 Measuring the spectrum . . . . . . . . . . . . . . . . . . . . . . . 148<br />

11.4.3 The spectrometer . . . . . . . . . . . . . . . . . . . . . . . . . . . 149<br />

11.4.4 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150<br />

11.5 Useful data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

12 γ ray detection and spectroscopy 153<br />

12.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153<br />

12.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153<br />

12.3 γ radiation. How? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153<br />

12.4 Calibration <strong>of</strong> the detector . . . . . . . . . . . . . . . . . . . . . . . . . . 154<br />

12.4.1 Detecting photons - the NaI(Tl) crystal . . . . . . . . . . . . . . . 154<br />

12.4.2 The Multi-Channel Analyser . . . . . . . . . . . . . . . . . . . . . 157<br />

12.4.3 The reference sources . . . . . . . . . . . . . . . . . . . . . . . . 157<br />

12.4.4 Gain <strong>of</strong> the detection system and MCA . . . . . . . . . . . . . . . 158<br />

12.4.5 Measuring spectra . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

12.4.6 Calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

12.5 Energy resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

12.6 Unknown sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

12.7 MCA commands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160


viii<br />

CONTENTS<br />

13 Energy loss <strong>of</strong> α particles through matter 163<br />

13.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163<br />

13.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163<br />

13.3 Measuring foil thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . 164<br />

13.3.1 Calibration <strong>of</strong> the system . . . . . . . . . . . . . . . . . . . . . . . 166<br />

13.3.2 Measurement <strong>of</strong> known foils . . . . . . . . . . . . . . . . . . . . . 167<br />

13.4 Making a thin foil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167<br />

13.4.1 Background reading . . . . . . . . . . . . . . . . . . . . . . . . . 167<br />

13.4.2 Foil manufacture . . . . . . . . . . . . . . . . . . . . . . . . . . . 168<br />

13.4.3 Thickness measurement . . . . . . . . . . . . . . . . . . . . . . . 170<br />

13.5 Rules for operating a high vacuum system . . . . . . . . . . . . . . . . . . 170<br />

13.6 MCA commands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171<br />

14 Errors 173<br />

14.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173<br />

14.2 Sources <strong>of</strong> error in experimental work . . . . . . . . . . . . . . . . . . . . 174<br />

14.2.1 Mistakes (!!) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

14.2.2 Random errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

14.2.3 Reading errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

14.2.4 Systematic errors . . . . . . . . . . . . . . . . . . . . . . . . . . . 177<br />

14.3 Graphs and errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179<br />

14.3.1 Error bars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179<br />

14.3.2 Estimates <strong>of</strong> gradients and uncertainties . . . . . . . . . . . . . . . 179<br />

14.4 Excel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181<br />

14.5 Combining errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181<br />

14.6 Significant figures and data/error presentation . . . . . . . . . . . . . . . . 183<br />

14.7 General expressions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184<br />

14.7.1 General expression . . . . . . . . . . . . . . . . . . . . . . . . . . 184<br />

14.7.2 Sums and differences . . . . . . . . . . . . . . . . . . . . . . . . . 184<br />

14.7.3 Products and quotients . . . . . . . . . . . . . . . . . . . . . . . . 184


CONTENTS<br />

ix<br />

14.7.4 Powers and functions . . . . . . . . . . . . . . . . . . . . . . . . . 184<br />

14.7.5 Simple and trigonometric functions . . . . . . . . . . . . . . . . . 184<br />

14.8 Concluding comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185<br />

15 Excel tutorial 187<br />

15.1 Simple example, energy vs. ADC number . . . . . . . . . . . . . . . . . . 187<br />

15.1.1 Entering data into the worksheet . . . . . . . . . . . . . . . . . . . 187<br />

15.1.2 Making the plot . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187<br />

15.1.3 Fitting the data . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188<br />

15.1.4 Absolute referencing . . . . . . . . . . . . . . . . . . . . . . . . . 188<br />

15.1.5 Using formulas to transform the data . . . . . . . . . . . . . . . . . 189<br />

15.2 Nonlinear fitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189<br />

15.2.1 The Solver Add-In . . . . . . . . . . . . . . . . . . . . . . . . . . 189<br />

15.2.2 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190<br />

<strong>Bibliography</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191<br />

16 The Safe Use <strong>of</strong> Lasers 193<br />

16.1 General comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193<br />

16.2 Precautions with lasers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194


Chapter 1<br />

Introduction<br />

The experiments in the second year laboratory cover a diverse range <strong>of</strong> physical phenomena.<br />

There are three main experimental areas: optics, nuclear and general, the last <strong>of</strong> which encompass<br />

waves, quantum mechanics and magnetic fields. Through this diversity, we hope<br />

that you will develop the ability to quickly adapt to and show a critical approach towards the<br />

tasks confronting you.<br />

You will spend four weeks in each area undertaking a new experiment each week. Each<br />

experiment will involve you spending six hours in the lab spread over two consecutive days.<br />

You will be expected to work in a team with your fellow students, calling on your demonstrator<br />

to clarify or discuss various aspects <strong>of</strong> the experimental procedure, results or analysis.<br />

A formal report will also be written for each experiment within the time allotted.<br />

The next few pages will describe what is expected <strong>of</strong> you in the lab in more detail. Please read<br />

these carefully before you enter the class to ensure that you are prepared for the semester.<br />

These notes will be a useful reference for you throughout the semester, so we suggest that<br />

you continuously refer back to them.<br />

Notes and other information are available on our website and the LMS:<br />

http://www.ph.unimelb.edu.au/˜part2<br />

The demonstrators and myself hope you enjoy yourself in the second year lab this semester.<br />

Please feel free to email me at the below address with any queries or requests. I will use<br />

email as the principal means <strong>of</strong> communication with students (via the LMS), so make sure<br />

that you check it every few days.<br />

Stephen Marshall<br />

Part II Coordinator<br />

part2@physics.unimelb.edu.au<br />

1


2 CHAPTER 1. INTRODUCTION<br />

1.1 Safety<br />

The <strong>University</strong> <strong>of</strong> <strong>Melbourne</strong> has adopted the internationally recognised systems SafetyMAP<br />

(ISO 12001) and Environmental Management System (ISO 14001) to ensure a safe and environmentally<br />

friendly workplace for all staff, students, and visitors. As a student <strong>of</strong> the<br />

<strong>University</strong> you are responsible for adopting safe work and study practices, and you are required<br />

to comply with all relevant <strong>University</strong> and Departmental rules and procedures.<br />

Detailed information on <strong>University</strong> policy and procedures is provided in the Environment,<br />

Health and Safety Manual at http://www.unimelb.edu.au/ehsm<br />

The Laboratory Rules and Safe Work Procedures set out below in this practical manual<br />

must be adhered to at all times, and the direction <strong>of</strong> all staff and demonstrators must be<br />

followed. If you have any concerns about the safety or environmental impact <strong>of</strong> any aspect<br />

<strong>of</strong> these practical classes, please raise them with the staff member in charge. Report all<br />

injuries and incidents to the staff member.<br />

Safety rules for the Second Year <strong>Physics</strong> Laboratories:<br />

• No food or drink in the labs at any time.<br />

• No open-toed shoes in the lab.<br />

• Long hair should be tied back.<br />

In the first instance, your demonstrator will alert you to any potential dangers in the lab, and<br />

you should raise any concerns you have with them. You may also consult the Part II Lab<br />

Coordinator (Stephen Marshall) about any safety or environmental concerns you may have.<br />

Some experiments in the labs have special safety issues associated with them (e.g. they<br />

involve the use <strong>of</strong> high voltages or lasers). You will find directions for working with this<br />

equipment in the lab notes for that experiment.<br />

If you have any allergies or medical conditions that you think may be affected<br />

by chemicals, material or procedures to be used in these practical<br />

classes, fill in a “Medical Status - Voluntary Notification for Laboratory<br />

Classes Form” and give it to the staff member in charge so that any risk can<br />

be assessed and the work procedures modified.<br />

A risk assessment has been carried out for these practical classes and any identified risks<br />

have been minimised. The risk assessments are available for examination - contact the Part<br />

II Lab Coordinator should you wish to do so.


1.2. EXPERIMENTAL TECHNIQUES 3<br />

1.2 Experimental techniques<br />

Here are some helpful hints when working in the lab to make your experience in the lab more<br />

efficient, beneficial and enjoyable:<br />

• Share the workload with your partner.<br />

• Make sure your occupations within the experiment are varied, and that the same person<br />

is not always taking the measurements or using the computer.<br />

• Discuss your work with your partner or friends, but remember that copying is prohibited.<br />

• Before you ask your demonstrator a question, think about possible solutions yourself<br />

first.<br />

• Always check a circuit before turning the power on.<br />

• If something is not working, check to see if it is turned on.<br />

• Think about how many measurements to take and over what range. A quick check <strong>of</strong><br />

the phenomena you will measure by quickly varying the conditions and watching the<br />

output will help you determine this.<br />

• Careless experimental techniques can damage or destroy vital and expensive equipment.<br />

Please take care with all equipment and report any damages to your demonstrator<br />

immediately.<br />

Remember that your demonstrator is there purely to teach, to help you come up with an<br />

answer, not to tell you the answer. Make full use <strong>of</strong> their knowledge and discussions.<br />

1.2.1 Computers<br />

Much <strong>of</strong> the analysis <strong>of</strong> your results will be done using Excel. You will <strong>of</strong>ten take your data<br />

and fit it to a known trend using the least squares method. More on how to do this using<br />

Excel is given in chapter 15.<br />

The computers themselves and their setup must not be tampered with. Severe penalties will<br />

be enforced if anyone is found doing so.<br />

1.3 The report<br />

Each experiment has been designed so that it can be finished with a completed report in the<br />

six hours allocated. However, this assumes that you have done adequate preparation<br />

before entering the laboratory.


4 CHAPTER 1. INTRODUCTION<br />

1.3.1 Preparation<br />

Preparing for each experiment involves reading the experimental notes and familiarising<br />

yourself with the theory involved. Begin any report writing by attempting any Pre-lab questions<br />

asked in the notes and carrying out some <strong>of</strong> the calculations. If this is not done you<br />

will be wasting valuable time in the lab and therefore penalising yourself.<br />

Careful! You will find the Pre-lab questions scattered all the way through the notes, so read<br />

them carefully.<br />

1.3.2 Writing the report<br />

The report has two purposes. The first is to act as an experimental log. This includes an<br />

accurate description <strong>of</strong> what was done (not what was written in the manual), as well as an<br />

accurate description <strong>of</strong> all observations. This provides the bulk <strong>of</strong> the Procedure and Results.<br />

The second role is the experimental report itself. This explains the purpose <strong>of</strong> the experiment<br />

in the form <strong>of</strong> the Aim, the meaning <strong>of</strong> the results through the Analysis and Discussion and<br />

what was learnt as a result <strong>of</strong> the experiment in the Conclusion. Each prac consists <strong>of</strong> several<br />

experiments designed to investigate a small area <strong>of</strong> physics. Each experiment is worthy <strong>of</strong><br />

its own report. Below is a little more detail about what is involved in writing a report.<br />

• Clearly state the experimental TITLE, DATE and NAME <strong>of</strong> your lab partner.<br />

• AIM AND THEORY<br />

The aim is a concise statement <strong>of</strong> the purpose <strong>of</strong> the experiment incorporating the<br />

following:<br />

– The theory to be tested<br />

– The quantities to be measured<br />

– Assumptions under which the experiment is to be done<br />

– Expected results including any theoretical values expected to be obtained<br />

– Definition <strong>of</strong> terms<br />

– Relevant formulae<br />

The theory can be easily incorporated into the aim. There is no need to copy sections<br />

<strong>of</strong> the manual as only a summary <strong>of</strong> the theory is required.<br />

We strongly advise you to write your background theory section first, and only then<br />

to write a brief statement <strong>of</strong> the specific aim (one or two sentences only). In this way,<br />

you will have defined all your terms before try to tell your reader what exactly you’re<br />

trying to find out.<br />

• PROCEDURE<br />

The procedure is an accurate record <strong>of</strong> what was done by you in the lab:<br />

– List all apparatus used<br />

– Draw circuit and block diagrams to show the experimental setup


1.3. THE REPORT 5<br />

– Show quantities measured<br />

– Briefly describe how quantities were measured<br />

– Record electronic settings if applicable<br />

This section should contain enough detail to reproduce your results without reference<br />

to the manual. Note: sometimes your procedure will be different to the manual.<br />

• RESULTS<br />

The results section is an accurate record <strong>of</strong> all observations, both visual and numeric:<br />

– Results are to be entered directly into your logbook in pen. Do not record your<br />

results on scraps <strong>of</strong> paper<br />

– Clearly present results - tables, colours, boxes<br />

– All measurements have errors. Explain where your errors come from. (Errors are<br />

explained in detail in chapter 14.)<br />

– Pertinent comments on experimental results or other observations<br />

– Mistakes should be explained, not erased<br />

– Remember to include units on all measurements<br />

• ANALYSIS AND DISCUSSION<br />

In this section the physical meanings <strong>of</strong> your results are discussed through the analysis<br />

<strong>of</strong> your data:<br />

– Graphs - clear and large<br />

– Calculations or sample calculations if many <strong>of</strong> the same analyses need to be made<br />

– Error analysis <strong>of</strong> your results<br />

– Compare your results with the theory. Is this the trend you expected? Is this<br />

value within the error bounds <strong>of</strong> your results? If not, why not?<br />

Through the analysis and discussion you must demonstrate your understanding <strong>of</strong> the<br />

physics involved and the limitations <strong>of</strong> the experiment.<br />

• CONCLUSIONS<br />

The conclusion is a brief summary <strong>of</strong> your findings and the importance they play in<br />

the role <strong>of</strong> physics. There should be no new information in this section;<br />

– Statement <strong>of</strong> final result<br />

– Success <strong>of</strong> theory - do your results concur with the theoretical ones?<br />

– Limitations <strong>of</strong> the experiment<br />

– Unexplained results<br />

– Suggestions for improving your results and technique for further investigation<br />

Ultimately, you must communicate what you learnt in the laboratory.


6 CHAPTER 1. INTRODUCTION<br />

1.3.3 Assessment<br />

Marks are not given for the quantity <strong>of</strong> material written but for its quality. Comments, which<br />

show that you understand or have thought about what is going on, are valuable. Clarity <strong>of</strong><br />

ideas, thoughts and understanding are essential for increasing your mark - a lack <strong>of</strong> these<br />

will detract from your mark. Your report should be legible but does not have to be a work <strong>of</strong><br />

art. It is your ideas and experimental ability you will be graded on.<br />

Each report will be marked on a scale <strong>of</strong> 0 to 10. As a rough guideline the following marks<br />

are associated with the corresponding areas in the report:<br />

Pre-lab questions 2<br />

Aim/Theory and Conclusion 1.5<br />

Procedure 1.5<br />

Results and Experimental ability 2<br />

Analysis and Discussion 3<br />

1.4 Sample report<br />

See the next three pages, a sample experiment <strong>of</strong> “Measurement <strong>of</strong> resistance”, for an idea<br />

<strong>of</strong> how to set out your report.<br />

1.5 Hints and tips<br />

What follows in this section is a list <strong>of</strong> things that demonstrators tend to find are consistent<br />

problems with student reports. It may not make much sense until you’ve done one or two<br />

experiments, but it should make lots <strong>of</strong> sense from then on! Some are issues <strong>of</strong> presentation,<br />

others <strong>of</strong> content. Some <strong>of</strong> this will seem like simple repetition <strong>of</strong> material that appears<br />

elsewhere in the introduction - that’s because it is. It is a response to the sorts <strong>of</strong> mistakes<br />

students make repeatedly! The further on you go through the lab, the more the demonstrators<br />

will expect <strong>of</strong> you - and the less happy they will be when you don’t jump the basic hurdles!<br />

Scattered amongst the list (which was made up by a demonstrator as she was marking, so<br />

isn’t complete or necessarily in order <strong>of</strong> importance) are some ways <strong>of</strong> thinking about report<br />

writing that, we hope, will help you!<br />

1. All work should be IN PEN please - this includes diagrams.<br />

2. If you write separate aims for different sections <strong>of</strong> the experiment, you must write<br />

separate conclusions. Most <strong>of</strong> the pracs in the general lab (and several labs elsewhere)<br />

fall into at least two clear divisions, each one worthy <strong>of</strong> a separate (consecutive) “minireport”.<br />

3. When presenting a mathematical pro<strong>of</strong>, your job is to justify each step. Basically, you<br />

will always have to include some words, not just a string <strong>of</strong> equations.


1.5. HINTS AND TIPS 7


8 CHAPTER 1. INTRODUCTION


1.5. HINTS AND TIPS 9


10 CHAPTER 1. INTRODUCTION<br />

4. Make your aims specific. For example, not just “to investigate waves” but “to investigate<br />

acoustic standing waves in a tube, and thus to find. . . ” This means you have to<br />

read the prac carefully before you arrive.<br />

5. When answering questions, paraphrase the question in the answer so that your<br />

reader doesn’t have to refer to the manual to find out what you’re talking about - your<br />

report should be self-contained.<br />

6. Make your diagrams big and clear - some <strong>of</strong> the equipment is quite complex, and a<br />

tiny diagram makes it hard to see what’s going on. A good guide is to make every<br />

diagram not less than one-third <strong>of</strong> a page.<br />

7. When describing your observations, make liberal use <strong>of</strong> diagrams (e.g. “the signal<br />

looked like this: ... ”).<br />

8. Do not just present a list <strong>of</strong> numbers - you must always draw physical conclusions<br />

from what you calculate/measure. Sometimes that conclusion will be “this verifies the<br />

theory” (if so, you must have presented the theory so that your reader can make the<br />

comparison), but sometimes it will be “this value is bigger than that one - this does (or<br />

doesn’t) make sense for the following reasons...”<br />

9. You need to imagine a reader that is dumb, lazy and mean. Most readers don’t want<br />

to have to work very hard, but they want pro<strong>of</strong> that you have. Also, they may one day<br />

be a competitor who is actively looking for mistakes - so get into the habit <strong>of</strong> justifying<br />

everything!<br />

10. Make sure your conclusion summarises your results. We are training you, in part,<br />

in how to write a scientific paper. Very few papers are read thoroughly - lots <strong>of</strong> readers<br />

do a quick scan over the method and turn straight to the conclusion to see what,<br />

specifically, you found out, and what if anything limited your results. If you spent<br />

three hours finding three numbers, you need to make sure those numbers are written<br />

down in the conclusion!<br />

11. ERROR ANALYSIS is always required and should be clear and complete, just like<br />

any other mathematical treatment you have to do in the course <strong>of</strong> the experiment. A<br />

number without an error attached is meaningless. There is never a time when you<br />

don’t have to do an error analysis.<br />

12. Don’t forget the importance <strong>of</strong> attaching units to all your numbers - again, without<br />

units your results are essentially meaningless. No reader can possibly judge the accuracy<br />

<strong>of</strong> your results without them.<br />

13. Method must be detailed - by that we mean it should include all the information you<br />

might need if you wanted to go back to this in three years when you’ve forgotten<br />

everything. So, things like the names and locations <strong>of</strong> any files you save are part <strong>of</strong><br />

this, as is the name <strong>of</strong> the s<strong>of</strong>tware you used and so on.<br />

14. It is useful to think <strong>of</strong> the lab report as a story - the story <strong>of</strong> the growth <strong>of</strong> your understanding<br />

throughout the prac. That means that you will write down everything you<br />

know at the beginning (i.e. context/theory) and all the stuff you find out by the end, as<br />

well as how you went about making your discoveries.


1.6. DEADLINES 11<br />

15. Don’t forget to define your terms - mathematical symbols are pointless if you don’t<br />

refer them back to the real world.<br />

1.6 Deadlines<br />

Your report is due at the end <strong>of</strong> the second session. This means you will have the six hours<br />

in the lab. The only extra time will be the night between the first and second sessions. It is<br />

therefore imperative that you write up as you go in order to finish.<br />

The demonstrator will assess your report and it will be available in advance <strong>of</strong> your next session.<br />

The books from the Monday - Tuesday class will be available after 2pm on Friday,<br />

those from the Wednesday - Thursday class will be available after 2pm on the following<br />

Monday and the books from the Thursday - Friday class will be available after 2pm on<br />

the following Wednesday.<br />

1.7 Illness<br />

If a class is missed due to an illness, please bring a medical certificate to your next class and<br />

a time will be arranged for the experiment to be made up. If classes are missed for other<br />

(previously known) reasons the makeup must be organised before the missed class. Please<br />

see your demonstrator or the coordinator to arrange this.


Part I<br />

Optics<br />

13


Chapter 2<br />

2-slit interference with single photons<br />

2.1 Introduction<br />

The double-slit experiment has as its origins an experiment performed in 1665 by Grimaldi,<br />

where he allowed light from the Sun to pass through two pinholes in an opaque screen. He<br />

had hoped to illustrate that when two circles <strong>of</strong> light overlap on a far screen, regions <strong>of</strong><br />

darkness result. That he could not illustrate this was due more to the lack <strong>of</strong> coherence in the<br />

sources given that the primary source <strong>of</strong> light, the Sun, was too spatially extended to provide<br />

such coherence.<br />

Young returned to the experiment in 1805, relying on knowledge <strong>of</strong> diffraction from a single<br />

slit (refer to the Fraunh<strong>of</strong>er) to provide a spatially coherent primary source <strong>of</strong> light: the<br />

central maximum from the diffraction pattern that emerges from a single slit. (The alternative<br />

is to use a laser source, which is spatially coherent by construction, althought that is not<br />

strictly done here.) When that was passed to two identical slits in close proximity, Young<br />

was able to observe interference fringes.<br />

The experiment gives rise to two interesting questions: what are the conditions <strong>of</strong> interference?<br />

And, more importantly, can it be done with a single photon?<br />

The experiment itself consists <strong>of</strong> two parts:<br />

• Investigation using the laser source. This essentially reproduces the classic experiment<br />

as a calibration <strong>of</strong> the equipment;<br />

• Investigation using the low-intensity lamp. After calibration, the equipment is used to<br />

study interference with single photons.<br />

Note! A couple <strong>of</strong> important definitions....<br />

Calibration (Oxford) n 1. The act or proccess <strong>of</strong> calibrating something; 2. each <strong>of</strong> a set <strong>of</strong><br />

graduations on an instrument etc.;<br />

Calibrate (Oxford) v.tr. 2. Correlate the readings <strong>of</strong> (an instrument) with a standard.<br />

15


16 CHAPTER 2. 2-SLIT INTERFERENCE WITH SINGLE PHOTONS<br />

P<br />

S<br />

r 1<br />

S 1<br />

r 2<br />

a<br />

θ<br />

S 2<br />

y<br />

d<br />

D<br />

Figure 2.1: Geometry for Young’s double slit experiment.<br />

2.2 Background theory<br />

Consider the configuration shown in Fig. 2.1. A light source illuminates a single slit S, which<br />

then acts as a source <strong>of</strong> spherical waves. If the distance to the double slit is much greater<br />

than the separation <strong>of</strong> the slits, d >> a, then the path difference for the light arriving at the<br />

double slits S 1 and S 2 will be minimal and the two slits act as sources <strong>of</strong> coherent spherical<br />

waves. Light travelling to point P in the image plane will have a path difference, ∆r, which<br />

for small θ is given by<br />

∆r = aθ = ay<br />

D . (2.1)<br />

The phase difference, δ, induced by this path difference is then<br />

( ) 2π (ay )<br />

δ = k∆r = , (2.2)<br />

λ D<br />

where λ is the wavelength <strong>of</strong> the light. The phase difference will modulate the intensity<br />

<strong>of</strong> the light measured at P due to the superposition <strong>of</strong> the light waves, with the resulting<br />

intensity distribution given by<br />

I = 2I 0 (1 + cosδ) . (2.3)<br />

Prelab Question (A): Show that the intensity distribution given in Eq. (2.3) leads to bright<br />

and dark fringes at y = mλD/a and y = (m + 1/2) λD/a respectively, where m is<br />

an integer.<br />

The spacing between successive bright or dark fringes is then<br />

y = λD a . (2.4)


2.3. APPARATUS 17<br />

Light sources<br />

Double slit and<br />

collimator<br />

Shutter<br />

Power<br />

Single slit and<br />

collimator<br />

Micrometers<br />

Detector box<br />

Figure 2.2: Schematic <strong>of</strong> the apparatus. (Not to scale.)<br />

However, this interpretation does not take into account that the fringe pattern is modulated<br />

due to diffraction effects from the slits themselves. That was the main criticism <strong>of</strong> the experiment<br />

at the time for which Fresnel and Lloyd suggested variations using either a biprism<br />

(Frensel) or a mirror (Lloyd) to create virtual, instead <strong>of</strong> real, sources. In this experiment, as<br />

we use real primary and secondary slits, we must take diffraction into account.<br />

The actual intensity distribution with angle is given by<br />

I(θ) = 4I 0<br />

( sin 2 β<br />

β 2 )<br />

cos 2 α, (2.5)<br />

where α = (ka/2) sinθ and β = (kb/2) sin θ, with a and b being the distance between the<br />

two slits and the slit width, respectively. I 0 is the intensity at θ = 0. This is the reduction for<br />

N = 2 from the multiple slits diffraction irradiance pattern. (See the “multiple slits” section<br />

<strong>of</strong> the Fraunh<strong>of</strong>er experiment.) The cos 2 α factor is the interference which is modulated by<br />

the sinc 2 β factor for single slit diffraction.<br />

Prelab Question (B): Derive Eq. (2.5) from the many-slits distribution as given in the Frauhnh<strong>of</strong>er<br />

experimental notes.<br />

2.3 Apparatus<br />

The apparatus is from TeachSpin and is shown schematically in Fig. 2.2. It should be aligned<br />

with the detector box in front <strong>of</strong> you and to the right. At the left is the power source and light<br />

sources, which consist <strong>of</strong> a red diode laser (5 mW, λ = 670 ± 5 nm), and a low-intensity<br />

white lamp with green detachable filter. The use <strong>of</strong> the filter limits the wavelength to be in<br />

the region 540 to 560 nm. The power is from a DC transformer and has controls which allow<br />

for switching between the laser and the lamp and the variation <strong>of</strong> the intensity <strong>of</strong> the lamp.<br />

You should find the apparatus with the top cover CLOSED. If this is not the case, CONSULT<br />

YOUR DEMONSTRATOR IMMEDIATELY. (For the reasons why, read on.)<br />

The full length is the optic bench. Further along the bench, one finds the first micrometer.<br />

This is the position <strong>of</strong> the double slit. At the right end <strong>of</strong> the apparatus is the detector<br />

box and second micrometer. The second micrometer controls the position <strong>of</strong> an aperture<br />

situation in front <strong>of</strong> the detectors and is calibrated in (true) mm. The detectors themselves


18 CHAPTER 2. 2-SLIT INTERFERENCE WITH SINGLE PHOTONS<br />

are a photodiode, for use with the laser, and a photomultiplier tube (PMT) for single photon<br />

counting. The controls on the box govern the power for the PMT, as well as the outputs for<br />

both.<br />

IMPORTANT! OPERATION OF THE APPARATUS. Also at the detector box is the<br />

shutter. It is the black cylinder with the cable coming out <strong>of</strong> it at the top <strong>of</strong> the box at the end<br />

<strong>of</strong> the optic bench. This switches between the two detectors. When the shutter is FULLY<br />

down, the PMT is blocked and the photodiode detector is exposed to the bench. The reverse<br />

situation occurs when the shutter is up. Along the top <strong>of</strong> the bench is a cover and this should<br />

always be in place when measuring the intensity pattern.<br />

WARNING! NEVER remove the bench cover when the shutter is UP.<br />

The PMT is super-sensitive to photons and even a small amount <strong>of</strong> background light is<br />

enough to destory the very thin layer <strong>of</strong> phosphor at the entrance window to the dynode<br />

chain. (See the Radiation Lab notes for an explanation on how a PMT works.) Be ABSO-<br />

LUTELY SURE that the shutter is FULLY DOWN before removing the cover at any<br />

time.<br />

With the shutter down, and with the PMT bias turned <strong>of</strong>f (toggle switch on power box), remove<br />

the bench cover by loosening the four latches and slowly sliding the cover away and<br />

up from the detector box end. You can now see the internal arrangement with the source slit<br />

and collimator sitting between the light sources and the first micrometer. The slits and collimators,<br />

both sets, are magnetically mounted on their respective posts and may be removed<br />

safely using the tweezers provided. Note that the first micrometer controls the position <strong>of</strong><br />

the second collimator (that associated with the two slits), while the second micrometer sits<br />

just in front <strong>of</strong> the shutter and controls an aperture to allow for sampling <strong>of</strong> the intensity.<br />

2.4 Experiment - Laser<br />

You will note that the white light source/green filter may be raised and lowered in front <strong>of</strong><br />

the laser. Be careful when moving the lamp: it is on a delicate mount and any motion should<br />

be slow and deliberate.<br />

Remove the single source slit and measure the slit width using the microscope provided.<br />

Measure also the widths <strong>of</strong> the double slits and their separation. Be sure to note the errors <strong>of</strong><br />

the masurements.<br />

Question (a): What measurements may you perform in order to have better confidence in<br />

the widths <strong>of</strong> the slits? Discuss with your demonstrator.<br />

2.4.1 Alignment<br />

Once the slits have been replaced (be sure to place them in the appropriate orientation), we<br />

may align the apparatus. Begin by raising the lamp into line with the laser, and remove


2.4. EXPERIMENT - LASER 19<br />

the filter. Level the lamp and turn on the bulb to about 6 on the intensity scale. Adjust the<br />

position <strong>of</strong> the source slit such as the primary source reaches the double slit roughly centered.<br />

Turn the lamp <strong>of</strong>f, replace the filter and move the lamp to the stored position. Turn on the<br />

laser and ensure that the beam reaches the source slit. Use the alignment screws to adjust<br />

the position <strong>of</strong> the beam until it is centered on the slit. What will you observe to guarantee<br />

alignment at the source slit? Make adjustments until the pattern reaches the double slit.<br />

The final bit <strong>of</strong> alignment concerns the collimator beyond the double slit. This should be<br />

positioned such that the vertical edges are parallel with the images <strong>of</strong> the two slits. Rotate<br />

the collimator until either slit instantly appears or disappears as you adjust the position <strong>of</strong> the<br />

collimator using the first micrometer.<br />

2.4.2 Measurement<br />

Using a white card, observe the image <strong>of</strong> the two slits just beyond the collimator. If the<br />

system is correctly aligned, you should see the double-slit interference pattern at the far end<br />

<strong>of</strong> the bench at the shutter’s position. Note the following collimator positions, they will be<br />

<strong>of</strong> great importance later:<br />

• Where both slits are blocked;<br />

• Where light emerges from the farther slit;<br />

• Where light emerges from both slits;<br />

• Where light emerges from the nearer slit;<br />

• Where both slits are blocked from the other side.<br />

Make observations on the images at the shutter for each <strong>of</strong> the cases.<br />

Question (b): What happens when you switch from one to two slits (and vice-versa) to a<br />

maximum or minimum at the shutter?<br />

You will note that at the shutter is another slit, <strong>of</strong> the same width as the source. This acts<br />

as an aperture for the detector, allowing for sampling the intensity pattern as one moves the<br />

micrometer. Do NOT touch this slit or adjust its position in its holder.<br />

Now replace the bench cover. You are now in a position to measure the intensity distribution<br />

<strong>of</strong> the pattern using the photodiode. The photodiode itself is a 1 cm 2 solid-state photodiode<br />

and generates a current upon illumination by a light source. The cable emerging from the top<br />

<strong>of</strong> the shutter carries the current from the photodiode and that should be place into the INPUT<br />

<strong>of</strong> the photodiode on the detector box. The OUTPUT carries the voltage as converted from<br />

the photodiode’s current. You may use the digital multimeter (on the 2V scale) to measure<br />

that voltage, which is proportional to the intensity <strong>of</strong> the light being sampled.<br />

With the cover on and the laser <strong>of</strong>f, measure the zero <strong>of</strong>fset <strong>of</strong> the diode. Turn the laser on,<br />

and position the first micrometer such that it is in the position where both slits are allowed


20 CHAPTER 2. 2-SLIT INTERFERENCE WITH SINGLE PHOTONS<br />

to emerge. Starting at the zero position <strong>of</strong> the second micrometer, measure the intensity at<br />

regular intervals between 0 and 10 mm. Be sure to choose a suitable interval such that any<br />

pattern is adequately resolved. You may enter the data directly into Excel, provided you<br />

printout a copy <strong>of</strong> the spreadsheet as well as the graph after analysis.<br />

At the central maximum, also measure the change in intensity by switching between having<br />

one slit allowed through and two. Does this correspond to your observations earlier?<br />

Using Excel, fit the assumed intensity distribution for double-slit diffraction [Eq. (2.5)].<br />

Question (c): Why is this necessary?<br />

When fitting, choose appropriate parameters with which to vary the function to fit the data.<br />

Compare the fitted values to those expected. Plot the comparison <strong>of</strong> the fitted function to the<br />

data, as well as the single-slit pattern as derived from Eq. (2.5).<br />

2.5 Experiment - Single Photon Counting<br />

At this point, we now move our attention to counting single photons. With the PMT and laser<br />

power OFF and the shutter DOWN, remove the cover and raise the lamp to be level with the<br />

laser. When in operation that would ensure that the light from the lamp would traverse the<br />

same paths as the light from the laser did in the previous run. When the lamp is in position,<br />

switch it on and adjust the bulb intensity until you see it glow. It should not be raised above<br />

6 for any length <strong>of</strong> time, nor should it be required. Ensure that the light from the lamp runs<br />

the length <strong>of</strong> the bench; realigning should not be necessary.<br />

Once the lamp is on and in position, replace the cover. At this point, the room’s lights may be<br />

switched on. With the output <strong>of</strong> the PMT connected to the CRO, and the shutter still closed,<br />

raise the PMT voltage slowly. Each major division corresponds to 100 V; see how the pulses<br />

change as the voltage is raised. Note where the pulses, corresponding to the situation <strong>of</strong> zero<br />

light, emerge. There should be a (dark) count rate <strong>of</strong> around 1-10/sec, once you reach 4-5 on<br />

the dial. At this point, the voltage is set correctly and you may raise the shutter. The count<br />

rate should now be much higher; investigate what happens as you adjust the bulb intensity,<br />

bearing in mind that you should not go above 6. (This isn’t Spinal Tap...)<br />

Connect the output <strong>of</strong> the discriminator to the CRO also and look at those pulses in coincidence<br />

with those from the PMT output. What signal form does the output <strong>of</strong> the discriminator<br />

take? Adjust the discriminator level so that one signal is produced per count above 50 mV<br />

from the PMT. This allows for the correct setting <strong>of</strong> the trigger.<br />

Note that the PMT efficiency is 4%. Using the counter provided, measure the actual count<br />

rate at a maximum in the intensity.<br />

Prelab Question (C): How does the width <strong>of</strong> the (single-slit) maximum change with the<br />

wavelength <strong>of</strong> the incident light?<br />

Adjust the discriminator level such that the count rate is around 10 3 /second.


2.5. EXPERIMENT - SINGLE PHOTON COUNTING 21<br />

Prelab Question (D): What is the error in every measurement <strong>of</strong> counts using the PMT?<br />

(Hint: check the nuclear notes...)<br />

By making sure you have an adequate count rate, do the same measurement as above <strong>of</strong> the<br />

intensity distribution, only this time counting actual photons in 10-second intervals.<br />

Prelab Question (E): Given that the length <strong>of</strong> the bench is around 1 m, calculate the time<br />

it takes for a photon to traverse the length <strong>of</strong> the bench. Given this time-<strong>of</strong>-flight,<br />

calculate the average time between photons for a (true) count rate <strong>of</strong> 2 × 10 4 /second,<br />

as may correspond to the true count rate <strong>of</strong> the experiment at the central maximum.<br />

What does this say <strong>of</strong> the nature <strong>of</strong> the measurement?<br />

Analyse the data as before, making only necessary adjustments to the parameters given the<br />

change in the initial conditions.<br />

Question (d): Discuss the situation as observed. What do you conclude about the nature <strong>of</strong><br />

the interference in this case?


Chapter 3<br />

The Michelson interferometer<br />

3.1 Abstract<br />

These experiments will examine some <strong>of</strong> the principles and applications <strong>of</strong> optical interference<br />

using the Michelson interferometer. The first sections look at the conditions <strong>of</strong> interference<br />

in the interferometer. Later sections will illustrate how the interferometer may be used<br />

to measure the separation <strong>of</strong> spectral lines in sodium and the refractive index <strong>of</strong> air.<br />

3.2 Introduction<br />

3.2.1 Overview<br />

Optical interference historically has provided some <strong>of</strong> the strongest evidence to support the<br />

wave theory <strong>of</strong> light. In addition, the interferometer constructed by Michelson and Morley<br />

in the 19 th century also provided the nail with which the c<strong>of</strong>fin was closed on the concept<br />

<strong>of</strong> the ether 1 . Despite this seeming paradox, the phenomena arising from interference are<br />

most naturally explained if light is considered to be electromagnetic waves. The Michelson<br />

interferometer is an example <strong>of</strong> an amplitude splitting interferometer, where the wavefronts<br />

emitted from the light source remain undisturbed. A beamsplitter is used to produce two<br />

beams, each <strong>of</strong> which travel separate paths before being recombined. This not only allows<br />

individual adjustment <strong>of</strong> the length <strong>of</strong> each path, but also changes in the length <strong>of</strong> each arm<br />

can be measured accurately using the small wavelength <strong>of</strong> visible light.<br />

1 For a discussion <strong>of</strong> the ether, see Hecht [1]. The ether was assumed to be the medium in which electromagnetic<br />

waves propagated, by analogy with water waves propagating through water.<br />

23


24 CHAPTER 3. THE MICHELSON INTERFEROMETER<br />

3.2.2 Background theory<br />

3.2.2.1 Arrangement <strong>of</strong> the Michelson interferometer<br />

The Michelson interferometer uses quite a simple arrangement <strong>of</strong> mirrors and beamsplitters<br />

to produce interference fringes. Illuminating a diffusing ground-glass plate L with light<br />

from a discharge lamp (a mercury or sodium lamp, for example) produces an extended light<br />

source. Figure 3.1 illustrates the configuration. Light waves emitted from the source are<br />

incident on a glass plate, G 1 , the back <strong>of</strong> which is slightly silvered. The metal coating on G 1<br />

is not fully sufficient to reflect the light from L, so approximately half is transmitted through<br />

arm 2 to mirror M 2 , and the rest is reflected along arm 1 to M 1 . The beam traversing arm 2<br />

also passes through the compensator plate G 2 twice before returning to G 1 . G 2 is an exact<br />

copy <strong>of</strong> G 1 but without the silver coating. Part <strong>of</strong> the light from M 1 is transmitted through<br />

G 1 to the observer O, as is some <strong>of</strong> the light from reflected by G 1 from M 2 . In this way the<br />

beams from each arm interfere at O where the resulting pattern may be seen.<br />

M<br />

1<br />

L<br />

arm 1<br />

Source<br />

G<br />

G<br />

1 2<br />

arm 2<br />

M<br />

2<br />

O<br />

Figure 3.1: Schematic arrangement <strong>of</strong> the Michelson interferometer.<br />

Pre-lab Question 3.1 What is the purpose <strong>of</strong> the compensator plate, G 2 ?<br />

Pre-lab Question 3.2 If M 1 is moved a distance d, what is the change in the total path<br />

length for light travelling in arm 1 <strong>of</strong> the interferometer?<br />

The answer to Pre-lab question 3.2 is important with regards to the conditions <strong>of</strong> interference.<br />

3.2.2.2 Conditions <strong>of</strong> interference<br />

The conceptual rearrangement <strong>of</strong> the mirrors shown in figure 3.2 helps to explain how the<br />

fringes are formed. An observer at the detector will see both mirrors M 1 and M 2 simultaneously,<br />

each <strong>of</strong> which forms an image <strong>of</strong> the ground glass plate L, leading to the image planes<br />

L 1 and L 2 .


3.2. INTRODUCTION 25<br />

2d<br />

S<br />

d<br />

S<br />

1<br />

θ<br />

S<br />

2<br />

θ<br />

2dcos θ<br />

Detector<br />

M<br />

1<br />

M<br />

2<br />

L<br />

L<br />

L<br />

1 2<br />

Figure 3.2: Conceptual rearrangement <strong>of</strong> the Michelson interferometer.<br />

Consider a point S on the glass plate L. An image <strong>of</strong> S will be formed due to mirrors M 1<br />

and M 2 at S 1 and S 2 , respectively. If M 1 and M 2 have a relative separation d, then S 1 and S 2<br />

will have a relative separation <strong>of</strong> 2d and the path length D = 2d cos θ. Interference fringes<br />

will be observed at the detector, with destructive interference satisfying<br />

D = nλ . (3.1)<br />

Pre-lab Question 3.3 If the path length difference is an integral number <strong>of</strong> wavelengths,<br />

constructive interference would normally be expected. Why is destructive interference observed<br />

in this case? (You may need to refer to Hecht [1] to answer this one).<br />

Interference between two waves occurs also if two other conditions are met, namely:<br />

Coherence Two waves are said to be coherent if there is a constant phase difference, which<br />

may be zero, between them.<br />

Polarisation The polarisation <strong>of</strong> an electromagnetic wave is the vector indicating the plane<br />

<strong>of</strong> the wave’s electric field.<br />

Only waves which are coherent and have the same polarisation will interfere.<br />

Pre-lab Question 3.4 How does the arrangement <strong>of</strong> the interferometer preserve coherence<br />

and polarisation?<br />

3.2.3 Apparatus<br />

3.2.3.1 Interferometer<br />

The Michelson interferometer is shown schematically in figure 3.1. The mirror M 2 has fine<br />

controls to adjust the horizontal and vertical orientations, while M 1 can be displaced along<br />

arm 1 by adjusting the micrometer. The ground glass plate L has a line marked on it to act<br />

as a pointer for aligning the mirrors.


26 CHAPTER 3. THE MICHELSON INTERFEROMETER<br />

3.2.3.2 Light sources<br />

The light sources provided are a mercury lamp, a sodium lamp and a tungsten lamp. A filter<br />

(a piece <strong>of</strong> green cellophane) is placed over the mercury lamp to absorb all wavelengths<br />

except the one at 0.5461 µm, producing a light source that emits light <strong>of</strong> constant wavelength<br />

(green). Once switched on, the lamp will require a few minutes to achieve full brightness,<br />

and should not be switched <strong>of</strong>f until the session has ended.<br />

• Wavelength <strong>of</strong> Hg lamp: λ = 0.5461 µm<br />

The tungsten lamp is a thermal light source, emitting light with wavelengths covering the<br />

entire visible region.<br />

3.2.3.3 Polaroids<br />

A polaroid is a device that is highly asymmetric in terms <strong>of</strong> the light that it transmits. The<br />

preferred transmission axis in the polaroid is known as the polarisation axis, and only the<br />

components <strong>of</strong> the electric field <strong>of</strong> the incident light that are parallel to the polarisation axis<br />

will be transmitted. The result <strong>of</strong> this is that the light transmitted by the polaroid will have a<br />

well defined polarisation.<br />

Question 3.1 How does the change in the polarisation <strong>of</strong> the transmitted light affect its<br />

intensity?<br />

There are two polaroids that can be mounted in the interferometer. Both are standard camera<br />

lens polaroids and may be rotated around 360 ◦ in their mounts.<br />

3.2.4 Experimental work<br />

WARNING: THE DISCHARGE LAMPS<br />

CAN BECOME QUITE HOT. HANDLE THEM<br />

ONLY BY THEIR STANDS.<br />

Switch on the mercury lamp and allow at least 5 minutes for it to warm up. Position the lamp<br />

so that the emitted light evenly illuminates the ground glass plate. Note that the ground glass<br />

plate has a line marked on it. This acts as a pointer to help in the alignment <strong>of</strong> the images<br />

from mirrors M 1 and M 2 . There should be three images <strong>of</strong> the pointer visible when looking<br />

into the interferometer.<br />

Question 3.2 Explain from where each <strong>of</strong> the three images comes.


3.2. INTRODUCTION 27<br />

3.2.4.1 Initial adjustment<br />

Set the position <strong>of</strong> the micrometer on M 1 to be about the centre <strong>of</strong> the range. (Ask the<br />

demonstrator for help if you have trouble reading the micrometer.) Adjust the mirror M 2 using<br />

its controls until two <strong>of</strong> the three images are aligned. Once properly aligned, interference<br />

fringes should become visible. Careful adjustment <strong>of</strong> M 2 should improve the visibility <strong>of</strong><br />

the fringes and also alter their appearance.<br />

Question 3.3 What effect does adjusting the controls on M 2 have on the appearance <strong>of</strong> the<br />

interference fringes?<br />

Question 3.4 Explain the curvature <strong>of</strong> the fringes (again, you may find Hecht [1] useful).<br />

Discuss with your demonstrator.<br />

3.2.5 Conditions <strong>of</strong> interference - polarisation<br />

Adjust M 2 until at most about 10 slightly curved fringes are in the field <strong>of</strong> view. Place the<br />

polaroids in each arm <strong>of</strong> the interferometer with a slight rotational <strong>of</strong>fset so that reflections<br />

<strong>of</strong>f the surfaces <strong>of</strong> the polaroids will not distort the interference. Observe the effect on the<br />

fringes as you rotate one <strong>of</strong> the polaroids through 360 ◦ . What do you expect to observe?<br />

Does that agree with your observations?<br />

3.2.6 Conditions <strong>of</strong> interference - coherence<br />

Remove the polaroids from the interferometer and adjust M 2 so that there are about 20<br />

slightly curved fringes in the field <strong>of</strong> view and record the position <strong>of</strong> the micrometer on<br />

mirror M 1 . As the micrometer is adjusted the curvature <strong>of</strong> these fringes will change. Why?<br />

Rotate the micrometer so that the fringes become less curved and continue rotating until the<br />

fringes are curved in the opposite direction to that from which you started. Again record<br />

the position <strong>of</strong> the micrometer. Position the micrometer midway between the two positions<br />

and make minor adjustments to the micrometer position until parallel fringes are observed.<br />

Adjust M 2 until only one or two parallel fringes are observed.<br />

Question 3.5 What is the significance <strong>of</strong> adjusting the micrometer to achieve parallel fringes?<br />

What do parallel fringes suggest about the path difference between M 1 and M 2 ?<br />

Remove the mercury lamp (do NOT turn it <strong>of</strong>f) and position the tungsten lamp so that the<br />

ground glass plate is evenly illuminated. Look into the interferometer and SLOWLY rotate<br />

the micrometer until fringes become invisible. They should appear within two full turns <strong>of</strong><br />

the micrometer. If not, return to the original position and rotate in the other direction. Note<br />

the position <strong>of</strong> the micrometer where the fringes are clearest and record your observations.<br />

(Hint!) Note in particular the number <strong>of</strong> black and white fringes and the manner in which<br />

the colours disperse as you move away from the centre <strong>of</strong> the pattern.


28 CHAPTER 3. THE MICHELSON INTERFEROMETER<br />

Question 3.6 Explain why the colours in the pattern disperse as they do and why the fringes<br />

are observed such a small range <strong>of</strong> path difference. What property <strong>of</strong> the light source allows<br />

this pattern to become manifest?<br />

3.2.7 Micrometer calibration<br />

Switch the tungsten lamp <strong>of</strong>f and replace it with the mercury lamp. Adjust M 2 until there<br />

are about 10 straight vertical fringes in the field <strong>of</strong> view and rotate the micrometer so that the<br />

pointer is on a fringe and record the position. Slowly rotate the micrometer until 20 fringes<br />

have passed the pointer and record the new position. Repeat this 10 times, always rotating<br />

in the same direction, so that a total <strong>of</strong> 200 fringes will have moved past the pointer.<br />

Question 3.7 Using equation 3.1, determine the change in path difference necessary to move<br />

the fringe pattern by 200 fringes.<br />

Question 3.8 What is the total change in the position <strong>of</strong> the micrometer that produced the<br />

observed change in the path difference?<br />

Question 3.9 What would be the change in the path difference if the micrometer were rotated<br />

by one unit?<br />

3.3 Fringe visibility<br />

3.3.1 Overview<br />

Light emitted from the discharge lamps stems from atomic transitions in the gas <strong>of</strong> atoms<br />

contained in the lamp. The emitted light has a narrow wavelength band (or bands) corresponding<br />

to the energy (or energies) <strong>of</strong> the transition(s), such that<br />

E = hc<br />

λ<br />

(3.2)<br />

where E is the energy <strong>of</strong> the transition, λ is the wavelength <strong>of</strong> the light, h is Planck’s constant<br />

and c is the speed <strong>of</strong> light. For sodium, there are two transitions <strong>of</strong> similar energy (call them<br />

E 1 and E 2 ) which lead to the emission <strong>of</strong> light, with the separation <strong>of</strong> their wavelengths<br />

being ∆λ = 0.597 nm.<br />

The photons emitted due to two close atomic transitions each form an independent fringe pattern<br />

in the interferometer. The small difference in wavelengths means that the periodicity <strong>of</strong><br />

the interference patterns will be slightly different, and hence at regular intervals the maxima<br />

<strong>of</strong> one pattern will be aligned with the minima <strong>of</strong> the other, causing a secondary interference.<br />

Regions where no fringes can be seen are referred to regions <strong>of</strong> zero fringe visibility. If the<br />

change in path difference between each region is denoted by ∆D, then the number <strong>of</strong> fringes


3.3. FRINGE VISIBILITY 29<br />

counted from λ 1 is m 1 and that counted from λ 2 is m 2 , where m 2 = m 1 + 1. The separation<br />

between the two emitted wavelengths can then be determined by measuring ∆D and using<br />

∆D = m 1 λ 1 = m 2 λ 2 = (m 1 + 1)λ 2 (3.3)<br />

∆λ = λ 1 − λ 2 = λ 1λ 2<br />

∆D ≈ λ2<br />

∆D<br />

(3.4)<br />

Pre-lab Question 3.5 Show that<br />

∆λ = λ 1λ 2<br />

∆D<br />

(3.5)<br />

Pre-lab Question 3.6 Equation 3.2 gives the energy <strong>of</strong> the atomic transition in terms <strong>of</strong> the<br />

wavelength <strong>of</strong> the emitted photon. For photons <strong>of</strong> energies E 1 and E 2 show<br />

∆E = E 2 − E 1 ≈ hc∆λ<br />

λ 2 (3.6)<br />

3.3.2 Apparatus<br />

Sodium Lamp The sodium lamp is similar to the mercury lamp used earlier. Sodium only<br />

emits two very closely spaced wavelengths in the visible region, such that λ 1 ≈ λ 2 ≈<br />

λ.<br />

• Wavelength <strong>of</strong> sodium lamp: λ = 0.5893 µm.<br />

3.3.3 Experiment<br />

Replace the mercury lamp with the sodium lamp and adjust M 2 until there are about 20<br />

fringes in the field <strong>of</strong> view. Starting with the micrometer at 0, rotate until the fringe pattern<br />

starts to fade. Slowly rotate the micrometer until the pattern is no longer visible and record<br />

the position <strong>of</strong> this zero visibility region, along with a reasonable error based on the range<br />

over which the region is observed. Continue rotating the micrometer finding and recording<br />

the position <strong>of</strong> each successive zero visibility region until the end <strong>of</strong> the micrometer scale is<br />

reached.<br />

Determine the change in the micrometer position needed to move from one region to the<br />

next. Any significant discrepancies here may indicate that a region <strong>of</strong> zero fringe visibility<br />

has been missed. Return to these positions on the micrometer to find any missed regions.<br />

Question 3.10 Using the position <strong>of</strong> the first and last region <strong>of</strong> zero visibility, and the number<br />

<strong>of</strong> such regions, calculate the average micrometer change necessary to move between<br />

successive regions. Use the calibration obtained earlier to determine ∆D, the average path<br />

difference change between regions <strong>of</strong> zero fringe visibility.<br />

Question 3.11 Calculate the wavelength separation <strong>of</strong> the sodium lines using equation 3.4<br />

and hence determine the energy difference between the two atomic transitions responsible<br />

for the emissions. Quote the result in eV. Is your answer reasonable? Why?


30 CHAPTER 3. THE MICHELSON INTERFEROMETER<br />

3.4 Refractive index <strong>of</strong> air<br />

3.4.1 Overview<br />

It does have one. . . As the refractive index <strong>of</strong> a medium increases, the speed <strong>of</strong> light in the<br />

medium is reduced, which results in the effective optical path length for the light travelling<br />

through the medium increasing since the time taken to travel any given distance will increase.<br />

In an interferometer, the change in the optical path length generates a relative path difference<br />

if the refractive index in one arm <strong>of</strong> the interferometer is changes. The change in path<br />

difference can be used to measure the change in the refractive index <strong>of</strong> the original medium.<br />

3.4.2 Background theory: optical path length<br />

The optical path length is the distance that would be travelled by light in a vacuum in the<br />

time that it takes light to travel a distance s through a medium <strong>of</strong> refractive index n. If the<br />

refractive index <strong>of</strong> the medium varies along the path travelled, then the optical path length is<br />

given by<br />

∫<br />

OPL = n(s)ds . (3.7)<br />

The integral may be done if the refractive index <strong>of</strong> the medium is homogeneous over a<br />

distance s, whence<br />

OPL = ns . (3.8)<br />

Pre-lab Question 3.7 If a cell <strong>of</strong> length s in arm 1 <strong>of</strong> the interferometer is evacuated, what<br />

is the change in the optical path length as the refractive index changes from that for air to<br />

that for vacuum?<br />

3.4.3 Apparatus<br />

Gas cell The gas cell has a hand-operated pump attached to it that enables air to be removed<br />

from the cell. A release trigger located beneath the gauge allows the cell to be returned<br />

to atmospheric pressure. The length <strong>of</strong> the cell is (4.0 ± 0.1) cm.<br />

Telescope A telescope is available which may be mounted onto the interferometer to magnify<br />

the observed fringes. When mounting, the alignment <strong>of</strong> the telescope may not enable<br />

the fringes to be seen, whence some adjustments to the telescope may be needed.<br />

Mirror M 2 can also be adjusted to move the fringe pattern into the centre <strong>of</strong> the field<br />

<strong>of</strong> view.<br />

3.4.4 Experiment<br />

Using the sodium lamp as a light source, obtain clear circular fringes by refined adjustment<br />

<strong>of</strong> M 2 . Place the gas cell with pump attached into arm 1 <strong>of</strong> the interferometer and adjust the<br />

micrometer to ensure that the fringe pattern is not near a region <strong>of</strong> zero visibility. If the centre


BIBLIOGRAPHY 31<br />

<strong>of</strong> the rings is not clearly visible, mount the telescope onto the interferometer to magnify the<br />

pattern.<br />

Watch the pattern closely. Slowly pump down the gas cell, counting the number <strong>of</strong> dark<br />

fringes that appear from (or collapse into) the centre <strong>of</strong> the pattern. Continue pumping the<br />

cell until the fringe pattern no longer changes. Record the total number <strong>of</strong> fringes counted as<br />

the cell was evacuated. Let the cell back up to air by pressing and holding the release trigger<br />

for about 20 seconds. Repeat the procedure 10 times and determine the average number <strong>of</strong><br />

fringes counted as the cell is evacuated, using the standard deviation in the numbers as the<br />

error.<br />

Question 3.12 Using equation 3.8, determine the change in the optical path difference resulting<br />

from the evacuation <strong>of</strong> the cell.<br />

Question 3.13 Using your answer to Pre-lab question 3.7, calculate the refractive index <strong>of</strong><br />

air. Record the temperature in the room from the thermometer on the wall and determine the<br />

expected value <strong>of</strong> the refractive index at that temperature for the wavelength <strong>of</strong> sodium using<br />

the chart provided. Compare your measured value to that expected. (Hint - remember that<br />

the refractive index <strong>of</strong> the vacuum is n vacuum = 1).<br />

<strong>Bibliography</strong><br />

[1] E Hecht. Optics. Addison-Wesley, 3 rd edition, 1998.


Chapter 4<br />

Fraunh<strong>of</strong>er diffraction<br />

4.1 Abstract<br />

The subject <strong>of</strong> this experiment is the phenomena <strong>of</strong> the diffraction <strong>of</strong> waves. While with<br />

the advent <strong>of</strong> quantum mechanics, and its basis <strong>of</strong> wave-particle duality, we know <strong>of</strong> the<br />

diffraction <strong>of</strong> particles (diffraction <strong>of</strong> electrons and neutrons are the most notable examples),<br />

it was the diffraction <strong>of</strong> light that was a major advance and led to the acceptance <strong>of</strong> the wave<br />

theory <strong>of</strong> light. The main focus <strong>of</strong> this experiment will be the far-field intensity distribution<br />

resulting from the diffraction <strong>of</strong> light from a single slit. Related to that will be the study <strong>of</strong><br />

diffraction from a wire, and from a number <strong>of</strong> slits. Diffraction from other objects, and how<br />

all are related, will round <strong>of</strong>f the experiment.<br />

4.2 Introduction<br />

The work <strong>of</strong> Fresnel and Fraunh<strong>of</strong>er regarding the diffraction <strong>of</strong> light around objects was<br />

instrumental in the acceptance <strong>of</strong> the wave theory <strong>of</strong> light. The principles <strong>of</strong> diffraction<br />

developed by them are applied to any wave/particle, be they photons, electrons, protons,<br />

neutrons, or even atoms and molecules. The language <strong>of</strong> diffraction is even used in subatomic<br />

physics for describing the dominant processes <strong>of</strong> the interaction <strong>of</strong> various systems.<br />

A large number <strong>of</strong> applications <strong>of</strong> diffraction involves waves incident on large periodic structures<br />

such as crystals or molecules. The observed patterns are dependent on the structure <strong>of</strong><br />

the object causing the diffraction <strong>of</strong> the waves. With the knowledge <strong>of</strong> the patterns one may<br />

work backwards to reconstruct the structure <strong>of</strong> the diffracting object 1 .<br />

All <strong>of</strong> these applications share the same basic theory developed by Fresnel and Fraunh<strong>of</strong>er for<br />

the case <strong>of</strong> diffraction <strong>of</strong> light. It is this phenomenon, and Fraunh<strong>of</strong>er diffraction specifically,<br />

that we shall investigate here.<br />

1 Francis Crick and James Watson discovered the double-helix structure <strong>of</strong> the DNA molecule by studying<br />

its X-ray diffraction pattern. They shared the Nobel Prize for Medicine, along with Maurice Wilkins, in 1962<br />

as a result. This was the most celebrated example <strong>of</strong> the importance <strong>of</strong> diffraction as a tool for studying the<br />

structure <strong>of</strong> matter.<br />

33


34 CHAPTER 4. FRAUNHOFER DIFFRACTION<br />

4.3 Theory<br />

Francesco Grimaldi in the 1600s was the first to observe in detail the deviation <strong>of</strong> light<br />

from rectilinear propagation due to the obstruction by a solid, opaque, object. The resulting<br />

shadow is made up <strong>of</strong> bright and dark regions quite unlike anything one might expect from<br />

Newton’s theory <strong>of</strong> light particles. Grimaldi called this phenomenon “diffractio” . The<br />

effect is a general characteristic <strong>of</strong> waves occurring whenever a portion <strong>of</strong> the wave front is<br />

obstructed in some way. This is independent <strong>of</strong> the form <strong>of</strong> the propagating wave. It is,<br />

in fact, a result <strong>of</strong> the interference <strong>of</strong> the segments <strong>of</strong> the wavefronts propagating beyond the<br />

obstacle, when that obstacle alters part <strong>of</strong> the wave front in terms <strong>of</strong> amplitude, phase, or<br />

both.<br />

4.3.1 Fresnel or Fraunh<strong>of</strong>er?<br />

Consider an opaque screen Σ, containing a single, small aperture, which is illuminated by<br />

plane waves from a distant point source, S. A pattern is observed after Σ in a plane σ parallel<br />

and very close to Σ. At this point, a clear image <strong>of</strong> the aperture is visible. When one starts<br />

to move σ away from Σ,<br />

Fresnel diffraction results. Fringes begin to appear about the structure <strong>of</strong> the aperture in<br />

question. This is near-field diffraction. If one moves still further, to a distance far<br />

from Σ,<br />

Fraunh<strong>of</strong>er diffraction results. A continuous change in the fringes is effected to a point<br />

where the projected pattern is spread out considerably, and the pattern bears no discernible<br />

structure <strong>of</strong> the aperture. Beyond this, the change is only to the size and not<br />

the structure <strong>of</strong> the pattern as σ is moved further away. This is also known as far-field<br />

diffraction.<br />

The Fraunh<strong>of</strong>er condition is R > a 2 /λ, where R is the smaller <strong>of</strong> the two distances from S<br />

to Σ and from Σ to P (a point on σ), see figure 4.1.<br />

4.3.2 Diffraction by a single slit<br />

Consider plane waves incident on a slit <strong>of</strong> width d, as shown in figure 4.1. The contribution<br />

to the electric field dE at a point P on the screen plane σ, due to a segment <strong>of</strong> the slit dx, is<br />

given by<br />

dE = ε L<br />

sin (ωt − kr)dx , (4.1)<br />

R<br />

where ε L /R is the amplitude <strong>of</strong> the incident wave. The phase <strong>of</strong> the wave is dependent on<br />

r(x) which can, in the Fraunh<strong>of</strong>er approximation, be written as a linear function <strong>of</strong> x, such<br />

that<br />

r = R − x sin θ. (4.2)<br />

Integrating over x leads to the expression <strong>of</strong> the electric field at P , viz.<br />

E = ε L<br />

R<br />

∫ d/2<br />

−d/2<br />

sin [ωt − k (R − x sin θ)] dx , (4.3)


4.3. THEORY 35<br />

X<br />

Σ<br />

σ<br />

P<br />

d/2<br />

dx<br />

R<br />

−d/2<br />

Z<br />

L<br />

Figure 4.1: Plane waves incident on a slit <strong>of</strong> width d. The irradiance at point P is given by<br />

equation 4.6.<br />

from which,<br />

where<br />

E = ε Ld sin β<br />

sin (ωt − kR) , (4.4)<br />

R β<br />

β = kd<br />

2<br />

sin θ =<br />

πd<br />

λ<br />

sin θ . (4.5)<br />

The irradiance or time-averaged intensity, I(θ), is the observed quantity and it is given by<br />

I(θ) = ε 0 c 〈 E 2〉<br />

= ε ( ) 2 ( ) 2<br />

0c εL d sin β<br />

2 R β<br />

( ) 2 sin β<br />

= I(0) , (4.6)<br />

β<br />

where 〈 sin 2 (ωt − kR) 〉 = 1/2, and I(0) is the intensity at β = 0. This defines the intensity<br />

at any point on the detection plane in terms <strong>of</strong> measurable quantities. The distribution is<br />

shown in figure 4.2.<br />

Pre-lab Question 4.1 What is the relative intensity between the central maximum and the<br />

first subsidiary maxima?<br />

Pre-lab Question 4.2 Assuming that the separation, L, <strong>of</strong> the slit to the screen is known,<br />

derive a formula relating the irradiance to the distance along the diffraction pattern. Identify<br />

any quantities that are not known and which need to be determined from fitting the data.<br />

Pre-lab Question 4.3 Why is the slit in figure 4.1 machined to have bevelled (kinked) edges?


36 CHAPTER 4. FRAUNHOFER DIFFRACTION<br />

1<br />

0.8<br />

I(θ)/I(0)<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-10 -8 -6 -4 -2 0 2 4 6 8 10<br />

β<br />

Figure 4.2: Intensity distribution for the diffraction <strong>of</strong> light by a single slit.<br />

4.4 Experiment<br />

BEFORE BEGINNING, READ<br />

CAREFULLY CHAPTER 16 ON LASER<br />

SAFETY.<br />

You will be operating a laser during the course <strong>of</strong> this experiment and you must take great<br />

care in preventing stray reflections away from the apparatus. Stray reflections into people’s<br />

eyes may cause damage. Use the black card provided to block the beam when moving things<br />

on and <strong>of</strong>f the optical bench. The beam can reflect <strong>of</strong>f the posts and holders into the eyes <strong>of</strong><br />

someone standing behind you.<br />

4.4.1 Calibration<br />

Turn on the laser and ensure that the beam runs down the length <strong>of</strong> the optic bench. Place the<br />

microscope objective onto the bench in front <strong>of</strong> the laser, making sure to prevent any stray<br />

reflections. Position the objective such that a smooth diverging (spherical) beam is produced<br />

and that it illuminates evenly the screen in front <strong>of</strong> the camera. Vertical and horizontal<br />

controls on the optic mount are available for small adjustments to the alignment. Place the<br />

aluminium plate with holes separated by 5 mm into the slot where the screen is mounted.<br />

Turn on the camera and computer, following the instructions given (see appendix, section<br />

4.6.1), and run the imaging s<strong>of</strong>tware. Record the image <strong>of</strong> the dots from the plate.


4.4. EXPERIMENT 37<br />

You may need to adjust the plate to ensure that the line <strong>of</strong> dots is horizontal. Once a suitable<br />

line is produced, capture the image and take an intensity pr<strong>of</strong>ile <strong>of</strong> the dots. The peaks in the<br />

spectrum should correspond to each <strong>of</strong> the holes in the plate, and there should be about 10<br />

well-defined peaks. Playing with exposure times may help in getting a good pr<strong>of</strong>ile.<br />

Identify the pixels corresponding to each <strong>of</strong> the peaks in the pr<strong>of</strong>ile, and determine the average<br />

pixel separation between them. From that determine the pixel calibration <strong>of</strong> the camera<br />

in mm/pixel.<br />

4.4.1.1 The CCD camera<br />

The CCD camera provides a simple and efficient way in which to capture intensity information<br />

and convert it into voltage information. The linear way in which light intensity is<br />

converted to voltage makes the camera ideal for quantitative applications 2 . The voltage information<br />

can be digitised, making storage and analysis <strong>of</strong> CCD images very easy. The<br />

applications <strong>of</strong> CCD are too numerous to list here; bear in mind that all digital cameras are<br />

CCDs.<br />

The appendix (section 4.6.2) describes the operation <strong>of</strong> the CCD. You should read the material<br />

and answer the following question<br />

Pre-lab Question 4.4 The potential wells in the substrate can only hold a limited number<br />

<strong>of</strong> electrons due to the Pauli exclusion principle. When the potential well is full, the pixel is<br />

said to be saturated. What will happen to electrons emitted from the SiO 2 substrate if the<br />

closest pixel is full? (This is known as “blooming”.)<br />

The camera you will be using is equipped with “anti-blooming” technology, which greatly<br />

improves the quality <strong>of</strong> the images. Your demonstrator will explain the principles involved.<br />

4.4.2 Single slit diffraction<br />

4.4.2.1 Initial adjustments and observations<br />

Remove the plate and return the microscope objective to the apparatus rack. We now need<br />

to make sure that the laser runs parallel to the optic bench. Using a mounted piece <strong>of</strong> tracing<br />

paper in one <strong>of</strong> the carriages, check that the position <strong>of</strong> the laser beam spot on the paper is<br />

unchanged as the carriage is moved along the length <strong>of</strong> the bench. Use the horizontal and<br />

vertical controls on the laser mounts to correct any misalignment. (There are two mounts for<br />

the laser, and so the adjustment may be a little tricky.)<br />

Once aligned, replace the tracing paper with the single slit mount. A diagram <strong>of</strong> the slit<br />

is shown in figure 4.3. It is mounted on an aluminium plate which is appropriate for an<br />

optic mount. The slit should be placed as close as possible to the laser so that the diffraction<br />

pattern can propagate over as long a distance as possible. Allow room for the variable neutral<br />

density filter to be inserted between the laser and the slit.<br />

2 It is noteworthy that Einstein got his first Nobel prize for the photoelectric effect, the basis for all CCD<br />

technology.


38 CHAPTER 4. FRAUNHOFER DIFFRACTION<br />

Diffracted beam<br />

Laser<br />

Slit width adjustment<br />

Figure 4.3: The single slit.<br />

4.4.2.2 Neutral density filter<br />

WARNING! THE NEUTRAL DENSITY FILTER MAY PRODUCE STRONG<br />

REFLECTIONS. TAKE CARE TO ENSURE THAT REFLECTIONS ARE<br />

DIRECTED AWAY FROM ANY PATH THAT MIGHT BRING THEM IN<br />

CONTACT WITH PEOPLE IN THE LAB.<br />

The neutral density filter is an optically dense non-polarising material used to reduce the<br />

intensity <strong>of</strong> incident light. It is mounted on a stand separate to the optic bench to allow for<br />

easy placement between the laser and the diffracting element. It is circular, and rotating it<br />

changes the density <strong>of</strong> material thus changing the attenuation 3 <strong>of</strong> the light.<br />

Align the slit with the beam by adjusting the horizontal and vertical controls on the mount to<br />

obtain a clear and symmetric diffraction pattern at the screen.<br />

Question 4.1 What effect does changing the width <strong>of</strong> the slit have on the pattern produced?<br />

Question 4.2 What effect does rotating the slit about the beam axis have on the pattern<br />

produced?<br />

4.4.2.3 Recording an image<br />

Adjust the slit so that the 0.3 mm wire barely slides through. Closing the slit and opening<br />

slowly until the wire passes through will set the appropriate width without crushing the wire<br />

in the jaws <strong>of</strong> the slit. Use the micrometer to measure the width <strong>of</strong> the wire, then close the<br />

micrometer with no object in place to record any <strong>of</strong>fset.<br />

Scan the image <strong>of</strong> the diffraction pattern.<br />

3 “attenuation” means “reducing the intensity <strong>of</strong>”.


4.4. EXPERIMENT 39<br />

Question 4.3 Does the pattern produced on the monitor have as much detail as that seen<br />

on the screen? Consider your answer in the context <strong>of</strong> your answer to Pre-lab question 4.2.<br />

What is happening in the camera to reduce the resolution <strong>of</strong> the system? Discuss with your<br />

demonstrator.<br />

Rotate the slit to make sure that the entire diffraction pattern is horizontal. Adjust the ambient<br />

light in the lab to reduce background if necessary.<br />

The image may now be stored to produce a pr<strong>of</strong>ile <strong>of</strong> the diffraction pattern. Check the<br />

following before saving the data:<br />

• Is there any saturation in the pattern? If so, can the saturation be reduced so that the<br />

whole pattern may be recorded?<br />

• Is the pattern symmetric, i.e. are the heights <strong>of</strong> the peaks to each side <strong>of</strong> the centre the<br />

same?<br />

The high intensity <strong>of</strong> the central maximum compared to the rest <strong>of</strong> the pattern precludes any<br />

real possibility <strong>of</strong> analysing the whole pattern.<br />

Question 4.4 Based on your answer to Pre-lab question 4.2, use Excel to fit the formula to<br />

the observed line pr<strong>of</strong>ile.<br />

Run the program and extract the relevant parameters from your data. If you removed the<br />

central maximum from the fitting, use the parameters to recalculate the whole pattern and<br />

plot the whole pattern to the pr<strong>of</strong>ile from the recorded image without the central maximum.<br />

Question 4.5 How well does the fitted pr<strong>of</strong>ile match the data?<br />

4.4.3 Babinet’s principle<br />

The diffraction patterns produced by two complementary screens are the same, in the Fraunh<strong>of</strong>er<br />

limit, save for a small region in the centre <strong>of</strong> the pattern. This is Babinet’s principle.<br />

Its usefulness lies in the ability to study the structures <strong>of</strong> diffracting objects that are difficult<br />

to mount, by studying their complementary partners 4 .<br />

4.4.3.1 Background theory<br />

Two screens, Σ 1 and Σ 2 , are said to be complementary if the opaque regions in one match<br />

exactly the apertures in the other, such that if added together they form a completely opaque<br />

screen. If the electric field at the image plane due to screen Σ 1 is E 1 , and the field due to Σ 2<br />

is E 2 , then<br />

E 1 + E 2 = E 0 (4.7)<br />

where E 0 is the unobstructed field.<br />

4 see Hecht, [1], page 500.


40 CHAPTER 4. FRAUNHOFER DIFFRACTION<br />

Pre-lab Question 4.5 When E 0 = 0 it is clear that the diffraction patterns from the two<br />

screens are the same. Show what would happen for E 0 ≠ 0, and how the diffraction patterns<br />

are related.<br />

4.4.3.2 Experiment<br />

A piece <strong>of</strong> wire ≈ 0.3 mm thickness (the same as that used to set the slit width), attached to<br />

a circular mount is used as the complementary screen to the slit. Measure the thickness <strong>of</strong><br />

the wire with the micrometer, being extra careful not to crush the wire. The mount may be<br />

rotated on its optic mount to allow for a horizontal pattern to be observed.<br />

Mount the wire and observe the pattern. Removing the neutral density filter will produce a<br />

clearer pattern.<br />

Question 4.6 How does the pattern produced by the wire differ from that produced by the<br />

slit? Is Babinet’s theorem satisfied?<br />

Put back the neutral density filter and check the alignment <strong>of</strong> the wire with respect to the<br />

laser to make sure that the diffraction pattern produced is horizontal. Scan the image, noting<br />

again the distance between the object and screen. Once completed, turn <strong>of</strong>f the CCD camera<br />

(you won’t be needing it again).<br />

Question 4.7 With the central region removed, how does the diffraction pattern compare to<br />

the single slit pattern with the central maximum removed? What does this suggest?<br />

Question 4.8 Fit the expected intensity distribution and determine the width <strong>of</strong> the wire.<br />

How does it compare to the slit width you determined above? How do the values for I(0)<br />

compare?<br />

4.4.4 Multiple slits<br />

Diffraction from many slits combines both the ideas <strong>of</strong> diffraction and interference. While<br />

there is no actual physical distinction between the two, interference is generally reserved for<br />

instances where only a few wavefronts are involved whereas diffraction is the case corresponding<br />

to many wavefronts, as per the Huygens-Fresnel principle.<br />

4.4.4.1 Background theory<br />

For a system with N slits <strong>of</strong> width b and separated by a distance a, the integral in equation 4.3<br />

may be written as<br />

E = C<br />

∫ b/2<br />

−b/2<br />

F(x)dx + C<br />

∫ a+b/2<br />

a−b/2<br />

F(x)dx + · · · + C<br />

∫ (N−1)a+b/2<br />

(N−1)a−b/2<br />

F(x)dx, (4.8)


4.4. EXPERIMENT 41<br />

where F(x) = sin [ωt − k(r − x sin θ)] as before and C is a constant related to the amplitude<br />

<strong>of</strong> the wave. This leads to the following expression for the irradiance distribution at the image<br />

plane in the Fraunh<strong>of</strong>er limit,<br />

I(θ) = I 0<br />

( sin β<br />

β<br />

) 2 ( ) 2 sin Nα<br />

, (4.9)<br />

sin α<br />

where β = (πb/λ) sin θ, α = (πa/λ) sinθ, and I 0 = I(0)/N 2 . The distribution is shown in<br />

figure 4.4.<br />

1<br />

a = 3b; N = 8<br />

(sin(β)/β) 2 (sin(Nα)/sin(α)) 2<br />

(sin(β)/β) 2<br />

0.8<br />

I(θ)/I(0)<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-6 -4 -2 0 2 4 6<br />

β<br />

Figure 4.4: The irradiance distribution for 8 slits with the geometry given.<br />

The (sin (Nα)/ sin α) 2 factor can be considered to contain all the interference effects, with<br />

the overall diffraction pattern being produced with this interference being modulated by the<br />

irradiance distribution <strong>of</strong> a single slit. This leads to the following observations:<br />

1. Principal maxima will occur when (sin (Nα)/ sin α) = N, or when<br />

2. Minima will occur for sin(Nα) = 0;<br />

α = 0, ±π, ±2π,...; (4.10)


42 CHAPTER 4. FRAUNHOFER DIFFRACTION<br />

3. Local maxima in sin(Nα) will lead to subsidiary maxima.<br />

Pre-lab Question 4.6 Determine the values <strong>of</strong> α corresponding to the minima and subsidiary<br />

maxima in the diffraction pattern. For an aperture <strong>of</strong> N slits, how many subsidiary<br />

maxima will occur between principle maxima? How many minima will occur in the same<br />

region?<br />

A diffraction grating is effectively a large number <strong>of</strong> slits <strong>of</strong> equal spacing and equal width.<br />

The most useful aspects <strong>of</strong> gratings are derived from equation 4.10, with the substitution <strong>of</strong><br />

α leading to<br />

a sin θ = mλ, (4.11)<br />

where m is an integer. This is known as the grating equation, and defines the following<br />

properties:<br />

1. The angular dependence <strong>of</strong> the principle maxima does not depend on the number <strong>of</strong><br />

slits, provided N > 1. The positions <strong>of</strong> the principle maxima in the imaging plane<br />

should then be fixed.<br />

2. The angular position <strong>of</strong> the maxima depends on λ, which makes the diffraction grating<br />

a wavelength dependent device. This has numerous applications, particularly where<br />

a light source composed <strong>of</strong> many wavelengths needs to be split into its constituent<br />

colours. (Another is the subject <strong>of</strong> a prac in this laboratory. . . 5 )<br />

4.4.4.2 Experiment<br />

There are two microscope slides. The first has groups <strong>of</strong> 1, 2, and 4 slits, each <strong>of</strong> equal width<br />

and spacing; the second has groups <strong>of</strong> 8 and 16 slits.<br />

Place the first <strong>of</strong> the microscope slides into the mount so that the pattern for the single slit is<br />

observed. Take note <strong>of</strong> the distribution and move the slide to observe the diffraction pattern<br />

<strong>of</strong> 2 slits. Continue with the other groups, taking note <strong>of</strong> the number <strong>of</strong> subsidiary maxima<br />

and minima.<br />

Question 4.9 Are the numbers <strong>of</strong> subsidiary maxima and minima consistent with your answers<br />

to Pre-lab question 4.6?<br />

Question 4.10 If a green laser were incident on the multiple slits concurrently with the red<br />

laser, how would the angular separation <strong>of</strong> the principle maxima change?<br />

4.5 Further work - circular and other diffracting objects<br />

Time and mojo permitting, you may wish to continue with the following sections.<br />

5 i.e. the Michelson experiment, chapter 3.


4.5. FURTHER WORK - CIRCULAR AND OTHER DIFFRACTING OBJECTS 43<br />

4.5.1 Overview and background theory<br />

Remember that the diffraction pattern produced by an object is directly related to the structure<br />

<strong>of</strong> that object. Hence, the importance <strong>of</strong> diffraction is obvious when one is dealing with<br />

a complex object whose structure is not readily apparent but produces a diffraction pattern<br />

that is far more easily studied.<br />

The theory <strong>of</strong> diffraction patterns produced by a circular object is far more complicated than<br />

the example <strong>of</strong> the single slit. Herein, we will only touch on the basics: you may wish to<br />

read section 10.2.5 <strong>of</strong> Hecht [1] if you want to know more.<br />

Plane waves incident on a circular aperture <strong>of</strong> radius a will form a circularly symmetric<br />

diffraction pattern on the screen σ at a distance L. That circular symmetry is a consequence<br />

<strong>of</strong> the electric field at σ being described by a class <strong>of</strong> functions known as Bessel functions,<br />

which themselves arise quite naturally when one is dealing with situations involving circular<br />

or spherical symmetry. The geometry is shown in figure 4.5.<br />

y<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

000000000<br />

111111111 O<br />

000000000<br />

111111111<br />

000000000<br />

111111111 a<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

x<br />

R<br />

L<br />

θ<br />

q 1<br />

y<br />

Figure 4.5: Geometry for the diffraction from a circular aperture.<br />

The resulting pattern has a large, high intensity, central region with secondary maxima forming<br />

rings around the centre. The area covered by the central maximum is called the Airy<br />

disc 6 . The secondary maxima are known collectively as Airy rings. The Airy disc extends to<br />

the centre <strong>of</strong> the first dark ring, with the radius given by<br />

q 1 = 1.22 Rλ<br />

2a , (4.12)<br />

where the variables are defined by the geometry (figure 4.5). Most <strong>of</strong> the intensity (84%) is<br />

concentrated in the centre, with the intensity falling away sharply.<br />

x<br />

4.5.2 Experiment<br />

Remove the mounted wire and replace it with the circular apertures. The three apertures,<br />

<strong>of</strong> varying radii, are mounted on a mounting similar to that for the wire. Align the smallest<br />

6 after Sir George Biddell Airy (1801–1892), who first derived the mathematical description <strong>of</strong> the intensity<br />

pr<strong>of</strong>ile.


44 CHAPTER 4. FRAUNHOFER DIFFRACTION<br />

aperture to obtain a clear diffraction pattern at the camera screen. With the calipers provided,<br />

measure the diameter <strong>of</strong> the radius <strong>of</strong> the first dark ring. Measure also the distance between<br />

the aperture and the image.<br />

Question 4.11 Is the diffraction pattern as expected? Does it resemble that for the single<br />

slit?<br />

Question 4.12 What happens to the diffraction pattern as one varies the radius <strong>of</strong> the aperture?<br />

Question 4.13 What is the radius <strong>of</strong> the smallest aperture?<br />

Replace the circular apertures with the mount <strong>of</strong> the circular discs. Those are <strong>of</strong> the same<br />

radii as the apertures, so this too is a test <strong>of</strong> Babinet’s theorem. Align the laser with the<br />

smallest disc to obtain a clear diffraction pattern.<br />

Question 4.14 How does the observed pattern differ with that from the aperture?<br />

Question 4.15 Allowing for those differences, measure the radius <strong>of</strong> the first dark ring and<br />

determine the radius <strong>of</strong> the smallest disc. Is this consistent with the measured radius <strong>of</strong> the<br />

aperture?<br />

4.6 Appendices<br />

4.6.1 CCD camera operating instructions<br />

• Ensure that the camera cables are connected before switching on the CCD camera,<br />

which is powered through the CCD computer (black box). This is connected to the<br />

PC.<br />

• Run the Fraunh<strong>of</strong>er image collection s<strong>of</strong>tware in the CCD s<strong>of</strong>tware folder on the PC.<br />

This is called “MaximDL”.<br />

• There are three main windows in MaximDL that you’ll use, which may be found in<br />

the view menu: the Image Window, the CCD Control Window, and the Line Pr<strong>of</strong>ile<br />

(LP) Window.<br />

• Under the CCD Control window, proceed to the Setup tab. Connect the camera (under<br />

s<strong>of</strong>tware) and cool the CCD. It should reach 0 ◦ C.<br />

• Once the desired temperature is reached and is stable, take an image. This is done in<br />

the CCD/Expose tab. Feel free to play with the settings to achieve best results for the<br />

image.<br />

• You may also crop the image to the desired size. All future images are captured to this<br />

size.


4.6. APPENDICES 45<br />

• An intensity pr<strong>of</strong>ile is obtained in the LP window. When that window is open, clicking<br />

on the captured image will place a horizontal line on the image. The pr<strong>of</strong>ile along that<br />

line will be plotted. That pr<strong>of</strong>ile may then be exported for analysis.<br />

• Save images as 12-bit FITS uncompressed (*.fit). Do not use compression as this is<br />

liable to lose information.<br />

• Powering down: from the CCD/Setup tab, turn the cooler <strong>of</strong>f and disconnect the camera<br />

BEFORE closing the program and turning <strong>of</strong>f the PC.<br />

4.6.2 How the CCD works<br />

The CCD camera consists <strong>of</strong> a two dimensional array <strong>of</strong> photodiodes coupled to the CCD<br />

array. A photodiode is a device that will absorb a photon and release an electron through<br />

the photoelectric effect. The photon energy required depends on the material used for the<br />

photodiode. For a material such as silicon dioxide (SiO 2 ) the energy required is 1.09 eV,<br />

corresponding to a photon with a wavelength <strong>of</strong> 1140 nm, which makes it ideal for detecting<br />

visible and near-infrared light.<br />

Figure 4.6: Schematic representation <strong>of</strong> CCD camera in a) initial condition with electrons<br />

trapped in potential well, b) intermediate state, and c) electrons displaced after one clock<br />

pulse.


46 CHAPTER 4. FRAUNHOFER DIFFRACTION<br />

To understand what the CCD array is and what it does consider the case <strong>of</strong> the one-dimensional<br />

CCD shown in figure 4.6. A silicon dioxide layer is deposited onto a p-type semiconductor<br />

silicon substrate. An array <strong>of</strong> metal electrodes is placed over the top <strong>of</strong> the SiO 2 layer. When<br />

high and low voltages are applied alternately to electrodes, a local potential well is formed in<br />

the silicon substrate in the vicinity <strong>of</strong> the high potential electrode. Electrons emitted from the<br />

silicon dioxide are confined to the potential wells. Reversing the potentials on the electrodes<br />

shifts the potential well, which takes the electrons with it. In this way the electrons can be<br />

shunted along the array to the last potential well where an output voltage is recorded, with<br />

the number <strong>of</strong> pulses counted since the initial shift indicating the original pixel for the data.<br />

A longer description is provided in the ST-237 CCD Camera manual.<br />

<strong>Bibliography</strong><br />

[1] E Hecht. Optics. Addison-Wesley, 3 rd edition, 1998.


Chapter 5<br />

Holography<br />

5.1 Abstract<br />

Holography is an application <strong>of</strong> the interference property <strong>of</strong> light. From its limited beginnings,<br />

it is now enjoying widespread application from becoming a medium in the arts to the<br />

recording <strong>of</strong> minute deformations in materials from external forces. In the course <strong>of</strong> this experiment,<br />

you will make several holograms thereby learning the basics <strong>of</strong> holography. And<br />

this is the only experiment in the entire undergraduate physics program where you keep the<br />

results!<br />

5.2 Introduction<br />

Holography is an application <strong>of</strong> the interference property <strong>of</strong> light. By superimposing two<br />

sets <strong>of</strong> monochromatic and coherent wave fronts on a photographic plate, a microscopic interference<br />

pattern is recorded. The developed plate/film is known as a hologram and the photographed<br />

interference pattern stores both amplitude and phase information. When the hologram<br />

is placed in a beam <strong>of</strong> the same coherent, monochromatic light, the beam is diffracted<br />

through the fringes to produce wavefronts identical to those originally reflected from the<br />

object. Consequently, when viewed, the resultant wavefronts give a remarkably realistic,<br />

three-dimensional picture <strong>of</strong> the object.<br />

In this experiment, you will produce two or three holograms. Note that this experiment does<br />

not involve taking quantitative measurements or doing calculations, and as a result many<br />

students have difficulty writing the report. Remember that although you are not finding<br />

“a number”, you will be making observations on the holograms produced and so you are<br />

required to report an accurate qualitative description <strong>of</strong> the observations made and <strong>of</strong><br />

the relevant physics ascribed to each. You must also take great care to keep a thorough<br />

informative record <strong>of</strong> what was done during the sessions, and explain clearly all experimental<br />

aims and procedures. As you have a lot <strong>of</strong> freedom in what you may wish to do, the record<br />

is <strong>of</strong> what you did.<br />

47


48 CHAPTER 5. HOLOGRAPHY<br />

5.3 Overview<br />

The principles <strong>of</strong> holography were first developed by Dennis Gabor in 1947 1 . Until 1960<br />

and the invention <strong>of</strong> the laser, application <strong>of</strong> this theory was limited (though not impossible).<br />

However, between then and the early 1970’s many useful and attractive applications for this<br />

process were discovered, making this one <strong>of</strong> the major new techniques, and physics success<br />

stories, <strong>of</strong> recent history.<br />

The production <strong>of</strong> holograms is quite simple, and the basic theory is relatively straightforward.<br />

Most <strong>of</strong> the background is given in the appendices, although a synopsis is given below.<br />

And you get to keep the holograms you produce to show <strong>of</strong>f to your friends later!<br />

SAFETY NOTE! THE LASER USED IN THIS EXPERIMENT IS A 30 MW<br />

HE-NE LASER AND QUITE POWERFUL. TAKE GREAT CARE IN USING IT<br />

AS IT MAY CAUSE PERMANENT DAMAGE IF REFLECTED INTO THE EYE.<br />

5.4 Background<br />

The concept behind the hologram is relatively simple. Normal photography only records<br />

the intensity <strong>of</strong> an image, and all colours therein, and thus only a two-dimensional image is<br />

recorded. In order to produce the exact wavefront <strong>of</strong> the reflected light <strong>of</strong>f any object, the<br />

relative phases <strong>of</strong> the light reflecting <strong>of</strong>f the object, which vary according to its structure,<br />

must also be recorded. The observer then perceives the reconstructed image, when coherent<br />

light is passed through the diffraction grating (hologram), as three-dimensional. Gabor<br />

discovered the means as to how this relative phase may be recorded.<br />

5.4.1 Intensity <strong>of</strong> a wave<br />

The electromagnetic wave is constructed from transverse electric and magnetic fields; both<br />

are transverse to the direction <strong>of</strong> propagation and perpendicular to each other. The intensity 2<br />

<strong>of</strong> the wave is defined as the square <strong>of</strong> the amplitude <strong>of</strong> the electric field. For a plane wave<br />

<strong>of</strong> the form<br />

E = E 0 cos (k · r − ωt) (5.1)<br />

the intensity is given by<br />

I inst = ε 0 c ∣ ∣E 2∣ ∣ = ε 0 cE 2 0 cos 2 (k · r − ωt), (5.2)<br />

where E 0 is the amplitude <strong>of</strong> the electric field, k is the wave number and ω is the frequency.<br />

This intensity is known as the instantaneous intensity, being that given at time t. Normal<br />

1 for which he won the Nobel Prize in <strong>Physics</strong> in 1971.<br />

2 Often (as in Hecht, [1] section 3.3.2, page 50) the word irradiance is used instead <strong>of</strong> intensity. Both refer<br />

to the average energy per unit area per unit time.


5.4. BACKGROUND 49<br />

photography is an example where the time-averaged intensity, is recorded. That is given by<br />

I = ε 0 c 〈 |E| 2〉<br />

= ε 0 c 〈 E0 2 cos 2 (k · r − ωt) 〉<br />

= ε 0cE 2 0<br />

2<br />

(5.3)<br />

Note that in these experiments, spherical waves will be used to construct the holograms. The<br />

generalisation <strong>of</strong> these ideas to spherical waves can be found in appendix 5.6.1.<br />

5.4.2 Interference<br />

A problem arises in the recording <strong>of</strong> phase. Absolute phase is a little difficult to record: it<br />

doesn’t exist. Phase may only be discussed as between two waves, i.e., the relative phase is<br />

the relative angle between the two waves and interference is determined by the size <strong>of</strong> this<br />

relative phase. There are also relative phases due to reflections from different parts <strong>of</strong> the<br />

object and we need a way to be able to measure these.<br />

If one is to record a relative phase, then one needs two waves with which to record an<br />

interference pattern. The phase may then be recovered from this pattern. One <strong>of</strong> the waves<br />

is obvious: it is that reflected from the object. The other must be in phase (i.e. a relative<br />

phase <strong>of</strong> zero) with the wave used to reflect <strong>of</strong>f the object if the changes in the phase after<br />

reflection are due only to the structure <strong>of</strong> the object itself. This reference beam is best<br />

constructed from the same light source to preserve the initial phase. How this is done is<br />

partly determined by which hologram we wish to produce.<br />

Appendix 5.6.2 develops the ideas <strong>of</strong> interference for plane waves.<br />

5.4.3 Holograms and holography<br />

Gabor coined the term holography from the (ancient) Greek word holos, meaning whole 3 .<br />

Holography, from Gabor’s earliest experiments through to modern applications, differs from<br />

standard photography in that it does not use lenses. Instead, the object is illuminated with a<br />

coherent monochromatic source <strong>of</strong> light, such as a laser, and the film is positioned to receive<br />

the light reflected from the object. As there is no lens, each point on the object reflects light<br />

to every point on the film. On its own, such an exposure would be rather boring: it would be<br />

a uniform (grey) and structureless picture.<br />

The same source also produces the reference beam and it too is reflected by a series <strong>of</strong> mirrors<br />

to illuminate the film with the same type <strong>of</strong> wave that illuminates the object. The interference<br />

<strong>of</strong> the two waves is formed in the plane <strong>of</strong> the film and it is that which is recorded. This interference<br />

pattern stores both amplitude and phase information. When light <strong>of</strong> the same form<br />

as the initial wave passes through it, the pattern acts as a diffraction grating, reconstructing<br />

the wavefronts as those reflected <strong>of</strong>f the object.<br />

The transmission functions may be found in appendix 5.6.3.<br />

3 In modern Greek, the word is entirely different: olos!


50 CHAPTER 5. HOLOGRAPHY<br />

Pre-lab Question 5.1 Why are we able to discern three dimensions with our sight? Why do<br />

we perceive the reconstructed image from the hologram to be three-dimensional?<br />

There are two main types <strong>of</strong> hologram: the transmission and the reflection. They are distinguished<br />

by the method used for recreating the wavefront.<br />

5.4.3.1 Transmission holograms<br />

The transmission hologram is one in which the initial wavefronts illuminate the grating (i.e.<br />

the film) and one perceives the (virtual) image as the transmission through the grating. This<br />

form <strong>of</strong> hologram requires that a coherent monochromatic source be used, although the wavelength<br />

<strong>of</strong> the light used to recreate the image need not be that <strong>of</strong> the light used to make the<br />

hologram. In that case, however, given that diffraction depends on wavelength, the image<br />

will change size if a different wavelength is used.<br />

5.4.3.2 Reflection holograms<br />

Reflection holograms, as the name suggests, require that the light used to recreate the image<br />

be on the same side as that <strong>of</strong> the observer, and the light is reflected back to the observer<br />

through the grating, forming the image. These holograms are necessarily thick, and are<br />

sometimes known as volume holograms. Such holograms are easily viewed with white light;<br />

those wavelengths which do not correspond to that which produced the hologram are absorbed<br />

by the plate. The reflection is then only <strong>of</strong> light with the same wavelength and the<br />

image is readily formed.<br />

5.4.3.3 In-line or <strong>of</strong>f-axis?<br />

The first holograms produced were in-line. That refers to the reference beam being in the<br />

same line as that reflected <strong>of</strong>f the object. The interference is seen as between the reference<br />

beam passing around the sample and the component <strong>of</strong> the beam which is distorted by the<br />

object. Coherence is not a problem as the same beam is used for both the object and reference.<br />

While these holograms may also be viewed using white light, it places constraints on the<br />

object. It must be small relative to the film, so that the reference beam may pass around it to<br />

reach the film, or the object must be transparent, so that the reference beam may pass straight<br />

through. Large, opaque objects will not work. Also, the beam used to reconstruct the image<br />

must be perpendicular to the hologram, making it difficult to see, as the observer tends to be<br />

in the way.<br />

Leith and Upatnieks each independently invented the <strong>of</strong>f-axis hologram in 1962. Those<br />

require high coherence, since two beams are required. Those are normally from the same<br />

coherent source. Any type <strong>of</strong> object may be used, but the same coherent, monochromatic<br />

source is required to reconstruct the image.


5.5. EXPERIMENT 51<br />

5.4.4 Movement<br />

Note! The wavelength <strong>of</strong> the light used determines the scale <strong>of</strong> the interference. This is<br />

usually <strong>of</strong> the order <strong>of</strong> micrometres or less, hence any vibration in the apparatus during the<br />

course <strong>of</strong> the exposing <strong>of</strong> the hologram will destroy the interference pattern and no hologram<br />

will be recorded. The optic bench on which the apparatus sits is greatly damped to prevent<br />

any vibrations from external sources ruining the exposure.<br />

Pre-lab Question 5.2 What would be seen in the image if part <strong>of</strong> the object was changed<br />

minutely during the course <strong>of</strong> taking the hologram? Discuss with your demonstrator.<br />

5.5 Experiment<br />

During the course <strong>of</strong> the experiment, you will be making up to 3 holograms. The first is the<br />

easiest, and will teach you aspects <strong>of</strong> optical alignment necessary in setting up the subsequent<br />

holograms. It is also the most fun.<br />

5.5.1 Film handling and developing<br />

Before any hologram may be produced, some notes on the film used and the technique to<br />

develop it.<br />

5.5.1.1 Holographic film<br />

Both Geola PGF-01 (red sensitive) holographic film and emulsion-coated glass plates are<br />

used in this laboratory. The film is used for the <strong>of</strong>f-axis holograms, while the plates are used<br />

for the white light. Note that care must be used in handling the glass plates: any undue force<br />

exerted will cause them to break. Also, take care to handle all film and plates along the edges<br />

to minimise any smudging <strong>of</strong> the emulsion which may ruin the hologram.<br />

The emulsion side <strong>of</strong> the film is facing away from you when you feel the clipped corner is on<br />

the top-right. Place the film with the emulsion side away from you in the brass film holder<br />

as shown in the figure 5.1.<br />

That film holder is then placed in the stand at the time <strong>of</strong> exposure. Take care when handling<br />

the brass plates to avoid moving the film away from its fixed position.<br />

The glass plates use a different method and holder. The holder is on its own stand and the<br />

glass plates slide straight into it. Great care must be taken to avoid breaking the glass as<br />

the experiment requires that this be a snug fit to prevent movement <strong>of</strong> the plate while the<br />

exposure is being taken. The emulsion side <strong>of</strong> the glass plate is slightly sticky to the touch<br />

when felt by a slightly damp finger-tip. (Do this at the edge; a finger-print at the centre <strong>of</strong><br />

the plate does not do the hologram any good. . . )


52 CHAPTER 5. HOLOGRAPHY<br />

Film<br />

Cut Corner<br />

Film Holder<br />

Figure 5.1: Orientation <strong>of</strong> the film in the film holder.<br />

5.5.1.2 Film exposure<br />

To produce high-quality holograms, the intensity <strong>of</strong> the reference beam should be at least<br />

three times and not more than ten times the intensity <strong>of</strong> the light reflected <strong>of</strong>f the object<br />

beam, as measured at the position <strong>of</strong> the film 4 . Glaring reflections from the object on the film<br />

should therefore be avoided, but they may provide some nice contrast in certain situations.<br />

All beams should be in the same plane for optimum hologram production. Avoid touching<br />

ALL mirror surfaces, especially the spherical mirrors.<br />

The resultant image will have the same appearance as the object as seen through a window<br />

having the same size and position as the photographic film relative to the object in the original<br />

configuration.<br />

5.5.1.3 Developing<br />

ALL FILM HANDLING AND PROCESSING<br />

MUST BE DONE IN COMPLETE DARKNESS.<br />

In a darkroom, develop the film in CWC2 Developer for 4 minutes, gently agitating the<br />

film. Rinse the film in the running water bath for 2 minutes. Place the film in the bleach<br />

for 3 minutes, again providing gentle agitation. Rinse again in the running water bath for 2<br />

minutes. ONLY AFTER bleaching may the light be switched back on. Allow the film to dry<br />

in the drying cabinet, typically for about 5 minutes, before viewing.<br />

If you wish to make a phase hologram, you must bleach the film for about 50% more time<br />

4 The maximum fringe contrast would occur for equal intensity but this would result in distortion.


5.5. EXPERIMENT 53<br />

than it takes to clear. Rinse for 3 minutes and then dry. Drying dehydrates the film which<br />

shrinks the emulsion. You should allow a few minutes after drying for the film to absorb<br />

moisture and expand back to normal before viewing.<br />

The developer is mixed in equal parts <strong>of</strong> Part A (20 g catechol, 10 g ascorbic acid, 10 g<br />

sodium sulphite, 50 g urea, and water to make 1 litre) and Part B (60 g sodium bicarbonate<br />

in 1 litre <strong>of</strong> water). It should last the whole <strong>of</strong> the experiment, after which it should be<br />

discarded safely in the waste drum provided.<br />

The bleach consists <strong>of</strong> 35 g copper sulphate, 10 ml glacial acetic acid, 110 g potassium<br />

bromide, and made up to 1 litre <strong>of</strong> water. It looks green when good and cyan (deep blue)<br />

when bad, but it is highly unlikely that it will turn during the course <strong>of</strong> the experiment. It<br />

is provided ready mixed for you at the beginning and will also last the entire experiment. It<br />

should be discarded safely at the same time as the developer in the waste drum provided for<br />

it.<br />

5.5.2 White light hologram<br />

The first hologram to be produced is the white-light hologram. It utilises the laser, a single<br />

plane mirror, and a spherical mirror. The arrangement <strong>of</strong> the system is shown in figure 5.2.<br />

Glass plate<br />

Object<br />

LASER<br />

Reference beam<br />

Figure 5.2: Experimental set-up for the white light hologram.<br />

Object beam<br />

This is your opportunity to be a little creative: bring in your own object(s). Note that the<br />

glass plates are 63 × 63 mm 2 , and the plate must be quite close (almost touching) the object<br />

in situ. Care must be taken in handling both the object and film in darkness. Bear in mind<br />

that with the red laser, suitable objects must also be chosen.<br />

Pre-lab Question 5.3 What characteristics make an object suitable for recording a hologram<br />

using a laser <strong>of</strong> a given colour? Choose your object(s) accordingly and remember to<br />

bring them at the beginning <strong>of</strong> the first session.<br />

If you are not so inclined, a selection <strong>of</strong> objects which may be used are provided. You may<br />

wish to use one <strong>of</strong> the lenses provided in front <strong>of</strong> whichever object you photograph (highly<br />

recommended).


54 CHAPTER 5. HOLOGRAPHY<br />

After setting up the experiment, take note <strong>of</strong> the intensity <strong>of</strong> the reference beam using the<br />

light meter provided. Take as many holograms <strong>of</strong> the object as number <strong>of</strong> partners involved,<br />

using an exposure time <strong>of</strong> 2 seconds. Develop using the method outlined and allow the<br />

holograms to dry completely before handling again.<br />

5.5.2.1 Viewing holograms<br />

As these are white-light holograms you may be able to see them in white light. The best<br />

source is the sun, so if the weather permits view your holograms outside. If that is not the<br />

case, a lamp is provided as a suitable point source.<br />

Question 5.1 Does the image appear three dimensional? What characteristics do you observe<br />

about the image? (The effects will be magnified if you used a lens as part <strong>of</strong> the<br />

object 5 .)<br />

Question 5.2 How many images may be seen? To what do these images correspond? Which<br />

is the clearer?<br />

5.5.3 Transmission or <strong>of</strong>f-axis hologram<br />

The set up for the transmission hologram is shown in figure 5.3.<br />

Plane mirror<br />

Spherical mirror<br />

Plane mirror<br />

Object<br />

LASER<br />

Variable beamsplitter<br />

Spherical mirror<br />

Film<br />

Figure 5.3: The experimental setup for the transmission hologram.<br />

You may use a similar object or objects as with the white light. In this case, you may wish<br />

to use the dragon as supplied.<br />

5 And, yes, the pun IS intended.


5.6. APPENDICES 55<br />

Once you have set up figure 5.3, and ensured correct optical alignment, take measurements <strong>of</strong><br />

the reference and object beam intensities at the film plane. Adjust the variable beam-splitter<br />

until a ratio <strong>of</strong> intensities between 3 and 10 is achieved.<br />

Question 5.3 What is the optimum intensity ratio for this setup? Discuss with your demonstrator.<br />

Using your exposure time from the first experiment as a guide, choose a new exposure time<br />

based on the intensity <strong>of</strong> the reference beam. Take 3 holograms, making minor adjustments<br />

in the chosen exposure time to allow for a variation in the exposure. Develop as before.<br />

5.5.3.1 Viewing holograms<br />

In this case, you must use the same monochromatic coherent source as that used for the<br />

reference beam. The reference beam itself in your setup is sufficient. NB! Care must be<br />

taken not to look directly into the reference beam as this is coming straight from the<br />

laser. Your demonstrator will show you how to observe the holograms. Adjust the angle<br />

<strong>of</strong> the film until you see the image <strong>of</strong> the object. The film should be roughly at the same<br />

angle relative to the reference beam when the hologram was taken, in which case the image<br />

should appear roughly the same distance from the film as the original object.<br />

You should be able to make a number <strong>of</strong> observations at this point. Can you see both images<br />

as with the white light? Does the image have the same 3-d properties as before?<br />

Using the best <strong>of</strong> the three holograms taken, cut one into halves. What do you observe?<br />

5.5.4 Time-averaged interferometry <strong>of</strong> a resonating tube<br />

If time permits, you may wish to try a second transmission hologram, that <strong>of</strong> a resonating<br />

tube.<br />

Replace the object from the previous experiment with the brass tube which has one end<br />

covered with thin paper and over which the other end has a small speaker, attached to a<br />

signal generator. Using the exposure time <strong>of</strong> the best hologram <strong>of</strong> the previous object, take<br />

three holograms at frequencies <strong>of</strong> 1500, 2100, and 3900 Hz. Note that while the stability <strong>of</strong><br />

the signal may be an issue, you should still be able to see something in at least two <strong>of</strong> the<br />

holograms.<br />

Question 5.4 What do you expect to see in the holograms? Describe each.<br />

5.6 Appendices<br />

5.6.1 Spherical waves<br />

A plane wave is given by the electric field defined in equation 5.1. From that, the intensity,<br />

irradiance, interference properties, etc., may be derived. Similarly, a spherical wave may be


56 CHAPTER 5. HOLOGRAPHY<br />

defined as<br />

E = E 0<br />

cos (kr − ωt). (5.4)<br />

r<br />

The instantaneous intensity <strong>of</strong> such a wave is then<br />

I inst = ε 0 c ∣ ∣ E<br />

2 ∣ ∣ = ε0 c E2 0<br />

r 2 cos2 (kr − ωt), (5.5)<br />

and the time averaged intensity becomes<br />

I = ε 0 c E2 0<br />

2r 2 . (5.6)<br />

A spherical wave thus decreases in intensity as the square <strong>of</strong> the radius to the source.<br />

5.6.2 Interference <strong>of</strong> two waves<br />

For any two waves, E 1 and E 2 , at any point in space or time they will add vectorially. The<br />

resulting intensity pattern will depend on how these vectors add. For E = E 1 + E 2 , the<br />

intensity is given by<br />

I inst = ε 0 c ∣ ∣E 2∣ ∣ = ε 0 c ∣ ∣(E 1 + E 2 ) 2∣ ∣<br />

= ε 0 c ( E 2 1 + E 2 2 + 2E 1 · E 2<br />

)<br />

= ε 0 c ( E 2 1 cos 2 (k 1 x − ω 1 t) + E 2 2 cos 2 (k 2 x − ω 2 t)<br />

+ 2E 1 E 2 cos 2 (k 1 x − ω 1 t) cos 2 (k 2 x − ω 2 t) cos θ 12<br />

)<br />

, (5.7)<br />

for the case <strong>of</strong> two planes. The angle θ 12 is the relative phase between the two waves.<br />

5.6.3 Transmission function<br />

After developing, the transmission function <strong>of</strong> the hologram is something like<br />

t(x,y) = t 0 + βI = t 0 + βε 0 c ∣ ∣E 2∣ ∣<br />

= t 0 + βε 0 c ∣ ∣E 2 1 + E 2 2 + 2E 1 · E 2<br />

∣ ∣ , (5.8)<br />

where t 0 and β depend on the transmission (or opaqueness) <strong>of</strong> the unexposed film, the sensitivity<br />

<strong>of</strong> the film to the light, and the developing process.<br />

If the reference wave is a plane wave pointing in the x direction,<br />

E 1 = E 01 cos (kx − ωt), (5.9)<br />

and the same wave is used to recreate the image, then the image wavefront is constructed as<br />

the transmission through the grating as defined by the transmission function. The hologram<br />

function is thus<br />

E holo = t(x,y)E 1 = (t 0 + βI)E 1<br />

= { t 0 + βε 0 c ( E 2 1 + E 2 2 + 2E 1 · E 2<br />

)}<br />

E1<br />

= t 0 E 1 + βε 0 c { E 2 1 cos 2 (k 1 x + ωt) + E 2 2 cos 2 (k 2 x + ωt)<br />

+ 2E 1 E 2 cos (k 1 x + ωt) cos (k 2 x + ωt) cos θ 12 } E 1 cos (k 1 x + ωt) . (5.10)<br />

This is a rather complicated beast, but it illustrates how the relative phase is recovered.


BIBLIOGRAPHY 57<br />

<strong>Bibliography</strong><br />

[1] E Hecht. Optics. Addison-Wesley, 4 th edition, 2002.


Part II<br />

General physics<br />

59


Chapter 6<br />

Magnets And magnetic fields<br />

6.1 Introduction<br />

The following two experiments examine the behavior <strong>of</strong> materials in two types <strong>of</strong> magnetic<br />

fields. The first day deals with the use <strong>of</strong> DC to observe the Hall Effect and constructing<br />

magnetic field probes. The second experiment deals with the study <strong>of</strong> AC fields by looking<br />

at the hysteresis <strong>of</strong> a material. This will allow us to calculate the permeability <strong>of</strong> a substance.<br />

6.2 The Hall effect and magnetic field probes<br />

6.2.1 Overview<br />

The Hall effect was discovered in 1879 by Dr. E. H. Hall as part <strong>of</strong> an investigation into the<br />

nature <strong>of</strong> the force acting on a current carrying conductor in a magnetic field. It wasn’t until<br />

the introduction <strong>of</strong> the semiconductor in the 1950’s that the full usefulness <strong>of</strong> this effect was<br />

recognised. Not only does the Hall effect allow for characterisation <strong>of</strong> semiconductors useful<br />

for their implementation in computer chips and electronic equipment, but the Hall effect in<br />

semiconductors also allows us to build devices that can measure an unknown DC magnetic<br />

field. This latter property is what you will be attempting to use today.<br />

6.2.2 Background theory<br />

6.2.2.1 Lorentz force equation<br />

The magnetic force, F , on a charge q, moving in a region <strong>of</strong> magnetic field, B, with a<br />

velocity, ⃗v, is given by,<br />

⃗F = q⃗v × ⃗ B. (6.1)<br />

In general, due to the fact that we are dealing with electric charges, not only are magnetic<br />

fields present, but electric fields also. Recall that the electric force on a charge q, is given by,<br />

⃗F = q ⃗ E. (6.2)<br />

61


62 CHAPTER 6. MAGNETS AND MAGNETIC FIELDS<br />

Hence the net force on a charged particle in both electric and magnetic fields is given by,<br />

⃗F = q( E ⃗ + ⃗v × B). ⃗ (6.3)<br />

This is the Lorentz force equation.<br />

Question 6.1 Explain how the Lorentz force equation can be used to implement a velocity<br />

discriminator for a collection <strong>of</strong> charged particles with a random distribution <strong>of</strong> velocities<br />

(e.g. e − from a hot filament).<br />

6.3 The Hall effect<br />

We can now consider the implications <strong>of</strong> the Lorentz force when we examine a semiconducting<br />

material. Figure 6.1 represents a thin strip <strong>of</strong> semiconducting material placed in<br />

an external magnetic field B Z . This material has a current I Y flowing through it at right<br />

angles to B Z . In order to analyse this situation, we must consider two types <strong>of</strong> semiconductors.<br />

One type is called n-type where the mobile charge carriers have negative charge, the<br />

other is where the mobile charge carriers are positive (p-type conductors). In both types <strong>of</strong><br />

z<br />

Bz<br />

x<br />

y<br />

z<br />

y<br />

x<br />

Iy<br />

Ex<br />

Figure 6.1: Current carrying specimen in a magnetic field.<br />

semiconductors, any moving charge can be characterised by a drift velocity v y in the same<br />

dimension as the current I Y . Since a charge has a drift velocity, it will experience a force<br />

in an electric field according to equation 6.1. In this case the right hand rule on the cross<br />

product shows the resultant force for the n-type semiconductor is down, while the force for<br />

the p-type semiconductor is up. Since the charges within the semiconductor are not free, but<br />

restricted in space by the boundary <strong>of</strong> the material, the charges will accumulate on the edges<br />

<strong>of</strong> the semiconductor. As the charges build up, an electric field E X will begin to appear.<br />

After a little thought it should be clear that for p and n-type semiconductors, the electric field<br />

produced will lead to a force on charge carriers opposite in direction to the force produced<br />

by the magnetic field. As the charge build up continues, E X will increase until the net force<br />

on the carriers is zero. Once this has been reached, we can re-write the force equation 6.3 as<br />

q(E X + v y × B Z ) = 0 (6.4)<br />

and hence<br />

E X = −v y × B Z . (6.5)


6.3. THE HALL EFFECT 63<br />

We now assign dimensions (X,Y,Z) to the specimen being considered. Assuming that the<br />

specimen has a conductivity σ, and n charge carriers with charge q per unit volume, we can<br />

define various properties <strong>of</strong> the conductor with respect to the electric and magnetic fields.<br />

We define the two dimensional current density j y as<br />

j y = nqu y = σE Y . (6.6)<br />

Where u y is the MEAN drift velocity u y = 〈v y 〉 <strong>of</strong> the carriers (in this case, for electrons,<br />

q = q e = 1.6 × 10 −19 ). The current is rightly defined as the sum <strong>of</strong> the current density over<br />

the cross sectional area, hence is an integral over the area <strong>of</strong> our semiconductor,<br />

∫<br />

I Y = j y dA = j y XZ (6.7)<br />

AREA<br />

If the magnetic field is applied in the z-direction, then the force (in the x-direction) is F X =<br />

B Z u y q, and using equation 6.5, the electric field at equilibrium is given by E X = −B Z u y .<br />

Recalling that the electric potential <strong>of</strong> a system is given by the path integral over the electric<br />

field, we have,<br />

∫<br />

V X = E X dX = XE X = −XB Z u y . (6.8)<br />

A similar expression holds for the relationship between V Y and E Y , with the negative sign<br />

implying that the potential across X is opposite to the direction <strong>of</strong> the applied field B Z . We<br />

now define the Hall coefficient as<br />

R H = E X<br />

j y B Z<br />

. (6.9)<br />

As will been seen later, the Hall coefficient provides an excellent indication <strong>of</strong> the charge<br />

concentration in the specimen.<br />

We also define the mobility <strong>of</strong> the charge carriers, given by,<br />

µ = u y<br />

E y<br />

. (6.10)<br />

The mobility <strong>of</strong> charge carriers gives an indication <strong>of</strong> how easy it is for the charge carriers to<br />

move through the specimen.<br />

Question 6.2 Derive the following expressions in terms <strong>of</strong> the measurable quantities;<br />

(a) σ from X, Y , Z, I Y and V Y .<br />

(b) R H from I Y , B Z , Z, and V X .<br />

(c) n from R H and q.<br />

(d) u y from V X , X and B Z .<br />

(e) µ from u y , V Y and Y .<br />

Question 6.3 Derive error expressions for all the terms in the previous question. Show a<br />

FULL derivation from Newton’s partial derivative expressions for errors for σ, in order to<br />

show that the standard product rule for errors can be applied to the other quantities.


64 CHAPTER 6. MAGNETS AND MAGNETIC FIELDS<br />

This section <strong>of</strong> the experiment will act to characterise the specific semiconductor used in this<br />

lab. Once characterisation has been performed, and values <strong>of</strong> σ, R H , n, u y and µ have been<br />

found, we can use measurements <strong>of</strong> V X , I Y etc. to determine an unknown field B Z .<br />

6.4 Experiment<br />

The experiment we will be performing will be to measure V X , V Y , and I Y for various values<br />

<strong>of</strong> B Z . This will allow for determination <strong>of</strong> all characteristic quantities discussed earlier.<br />

6.4.1 The semiconductor<br />

Figure 6.2 shows a schematic diagram <strong>of</strong> the circuit that will be used. The variable resistor<br />

at the base <strong>of</strong> the circuit can be used to control the amount <strong>of</strong> current passing through the<br />

specimen.<br />

Figure 6.2: Electrical circuit for the semiconductor.<br />

The dimensions <strong>of</strong> the semiconductor (placed between the poles <strong>of</strong> the large soliton magnet)<br />

are:<br />

• X = 4.0 mm<br />

• Y = 16.0 mm<br />

• Z = 0.47 mm<br />

The resistance <strong>of</strong> the semiconductor is 385 Ω.


6.4. EXPERIMENT 65<br />

Question 6.4 What are the errors associated with the dimensions and the resistance as<br />

stated? Are they negligible compared with the other quantities measured?<br />

6.4.2 The electromagnet<br />

The specimen itself is enclosed in an epoxy package and is fixed between the magnet: Do<br />

not attempt to remove it. The magnet has been calibrated as B = 0.97 T @ 100 mA.<br />

SAFETY NOTE: SOME OF THE<br />

VOLTAGES USED IN THIS EXPERIMENT<br />

ARE HIGH!!<br />

The current I m , for the Hall effect magnet is provided by the soliton power supply, the maximum<br />

current for which is 120 mA. The direction <strong>of</strong> the induced field is down when the<br />

conventional current (red) enters the top terminal.<br />

The maximum safe control current I y for the specimen is 15 mA.<br />

Question 6.5 What is the corresponding maximum value for V Y ? Use the given resistance<br />

from the semiconductor.<br />

6.4.3 The Hall effect<br />

Measure the Hall potential difference, V X , when the magnet current I m is 100 mA for at<br />

least 10 values <strong>of</strong> the specimen current, I Y , by varying the resistor. Also take note <strong>of</strong> the<br />

applied specimen voltage, V Y . Then reverse the magnetic field and repeat the measurements<br />

at the same I Y values. Make sure the magnet current is returned to zero before changing<br />

terminals.<br />

Question 6.6 Determine whether the contacts on the specimen are placed symmetrically.<br />

Note: a yes/no answer is not appropriate. Explain physically how you determine this and<br />

physically what is going on (draw a diagram). Remember that the contacts used to measure<br />

V X and supply I Y are finite and are placed roughly in the centre <strong>of</strong> the Y Z and XZ planes<br />

respectively.<br />

Question 6.7 From the relationships derived in the Pre-lab questions, calculate the values<br />

for the following at each setting <strong>of</strong> I y and determine the averages and errors.<br />

• conductivity σ,<br />

• Hall coefficient R H ,<br />

• charge density n,<br />

• mean drift velocity u y ,<br />

• mobility µ.


66 CHAPTER 6. MAGNETS AND MAGNETIC FIELDS<br />

Question 6.8 Using the background theory on the Hall effect and the sign <strong>of</strong> V X for a particular<br />

orientation <strong>of</strong> B Z , determine:<br />

• the sign <strong>of</strong> the charge carrier;<br />

• if the material is a p-type or n-type semiconductor.<br />

One interesting phenomena when looking at the magnetic fields is the way a current carrying<br />

wire in a magnetic field experiences a force.<br />

Question 6.9 Now that you know how the Hall effect works, predict the direction the wire in<br />

figure 6.3 will move.<br />

Figure 6.3: Current carrying wire in a magnetic field.<br />

Question 6.10 Using table 6.1 determine:<br />

• the dopant used (e.g. Ga,As).<br />

• the concentration <strong>of</strong> charge carriers.<br />

Impurity Concentration Type σ<br />

atom/m 3<br />

Ω −1 m −1<br />

nil zero intrinsic 2<br />

As 8 × 10 19 n 5<br />

As 1.5 × 10 21 n 90<br />

As 5 × 10 22 n 2000<br />

Ga 9 × 10 19 p 3<br />

Ga 8 × 10 20 p 30<br />

Ga 1 × 10 22 p 300<br />

Table 6.1: Conductivity <strong>of</strong> doped Germanium at 300 K<br />

6.4.4 The Hall probe<br />

Assuming that the Hall potential difference, V X , is a linear function <strong>of</strong> magnetic field, B Z ,<br />

the specimen that we have characterised can be used as a Hall probe to measure an unknown


6.5. FERROMAGNETIC MATERIALS AND AC FIELDS 67<br />

magnetic field. The linearity can be determined from the magnet calibration. Plot a graph <strong>of</strong><br />

B Z versus soliton current, I m . Proceed in the following way:<br />

• Start with 120 mA, decrease to zero;<br />

• reverse the supply polarity;<br />

• decrease to -120 mA then;<br />

• increase to zero.<br />

Discuss the features <strong>of</strong> the graph. Specifically, what equation(s) determined earlier show<br />

that the plot should be linear, and what possible effects could be producing non-linearity for<br />

large values <strong>of</strong> ±I m ?<br />

6.5 Ferromagnetic materials and AC fields<br />

6.5.1 Overview<br />

The process <strong>of</strong> putting a ferromagnetic material through a cyclic change in magnetic field<br />

was first studied by Sir J.A. Ewing. By changing the magnetic field in this way, one is able<br />

to determine the work done in putting the material through this process, the permeability and<br />

magnetic polarisation <strong>of</strong> the magnetic material.<br />

6.5.2 Background theory<br />

6.5.2.1 Magnetic polarisation<br />

If we consider the electron as a current, I, flowing around the nucleus about an area, A, the<br />

atom and electron spin can be modelled by a magnetic moment. We then define the magnetic<br />

moment<br />

m = iA, (6.11)<br />

A<br />

e<br />

Nucleus<br />

Figure 6.4: Magnetic model <strong>of</strong> an atom.<br />

where A is in the direction <strong>of</strong> the normal as shown in figure 6.4. If there are N atoms per<br />

unit volume, then the magnetisation, M is defined as<br />

i


68 CHAPTER 6. MAGNETS AND MAGNETIC FIELDS<br />

M = Nm = NiA. (6.12)<br />

The magnetisation is an indication <strong>of</strong> how many <strong>of</strong> the atoms are aligned in a particular<br />

orientation.<br />

6.5.2.2 Magnetic field intensity H and magnetic flux density B<br />

The magnetic flux density, B, is defined through Ampere’s circuital law<br />

∮<br />

B.dl = µ 0 i (6.13)<br />

The magnetic field intensity, H, is defined as<br />

H = B µ 0<br />

− M (6.14)<br />

The assumption may be made that M ∝ H, and therefore<br />

M = χ m H (6.15)<br />

where χ m is the magnetic susceptibility. Thus, the relationship between the magnetic field<br />

intensity and the magnetic flux density is<br />

B = µ 0 (1 + χ m )H = µH = µ 0 µ r H, (6.16)<br />

where µ is the absolute permeability <strong>of</strong> the material. In free space, M (and thus χ m ) are<br />

zero, giving the absolute permeability <strong>of</strong> free space as µ 0 = 4π × 10 −7 Wb A −1 m −1 . µ r<br />

is the relative permeability.<br />

A general way <strong>of</strong> writing equation 6.14 is<br />

B = µ 0 (H + M) (6.17)<br />

6.5.2.3 Effects <strong>of</strong> materials in a magnetic field<br />

The motion <strong>of</strong> electrons in an atom is made up <strong>of</strong> two parts: orbital motion (around the<br />

nucleus) and spin. It is these components, both orientation and magnitude, which determine<br />

the type <strong>of</strong> magnetism the material will exhibit in a magnetic field (for further detail see<br />

Ohanian [4], chapter 33, page 806).<br />

• Diamagnetism - when in the absence <strong>of</strong> an external magnetic field, the electron spins<br />

are randomly orientated and produce a zero atomic dipole strength. Once inside a


6.5. FERROMAGNETIC MATERIALS AND AC FIELDS 69<br />

magnetic field, magnetic moments are induced in the material, reducing the magnetic<br />

field in the material. This effect disappears with the removal <strong>of</strong> the field.<br />

• Paramagnetism - when in the absence <strong>of</strong> a magnetic field some <strong>of</strong> the atoms are<br />

aligned, but overall the magnetisation within the material is zero. Once the field has<br />

been introduced these atoms partially align producing a weak magnetic field. This<br />

effect disappears with the removal <strong>of</strong> the field AND thermal agitation.<br />

• Ferromagnetism - in this case the electrons are more strongly aligned and form domains<br />

(see figure 6.5) even without the introduction <strong>of</strong> a magnetic field. These domains<br />

are randomly aligned until the introduction <strong>of</strong> a magnetic field. As the external<br />

field is increased the domains tend to align with the external field until the material is<br />

essentially one large domain producing a very strong field.<br />

Figure 6.5: Magnetic domains in a ferromagnetic material (see Ohanian [4], chapter 33 page<br />

812.)<br />

So, when a ferromagnet is in an external field, which is strong enough to align all the domains,<br />

the material is said to be saturated and any further increase in the external field will<br />

not see a contribution from the material.<br />

Now let us investigate what happens to the material as we turn <strong>of</strong>f the external field, reapply<br />

it in the opposite sense until saturation is obtained again and then turning <strong>of</strong>f the field and<br />

reapplying it in the original sense. This is the process <strong>of</strong> recording the history <strong>of</strong> the material<br />

or obtaining the hysteresis loop (figure 6.6).<br />

• The first stage is from zero magnetisation in the material to the point <strong>of</strong> saturation,<br />

point P .<br />

• Next we decrease the external field, H, to zero, but there is a residual flux density,<br />

B, within the magnet, i.e. some domains are still in alignment. This is called the<br />

remanence. The remanence can be destroyed by thermal agitation. The temperature<br />

this occurs at is the Curie Temperature.<br />

• If the external field is reversed, then the aligned domains will begin to realign with the<br />

new field orientation. Firstly, the domains are randomly aligned and cancel, i.e. zero


70 CHAPTER 6. MAGNETS AND MAGNETIC FIELDS<br />

Figure 6.6: Typical hysteresis loop for a ferromagnetic material.<br />

flux density. Then they will reach saturation at point P ′ . At the point <strong>of</strong> zero flux<br />

density, the strength <strong>of</strong> the applied field required to destroy the remnant magnetisation<br />

<strong>of</strong> the material is called the coercive force.<br />

• If the field is then decreased and then reversed once again, a closed loop can be obtained.<br />

To do this continuously, the magnetic field is supplied by an AC field.<br />

Pre-lab Question 6.1 In light <strong>of</strong> this information, and knowing the hysteresis <strong>of</strong> the material,<br />

what properties would a good permanent magnet require? Subsequently, what would be<br />

the shape <strong>of</strong> the hysteresis loop for this material?<br />

Examining equation 6.17 one can see that for a diamagnetic material M is small and in the<br />

opposite sense <strong>of</strong> H, while for a paramagnetic material M is in the same sense, but small.<br />

For a ferromagnetic material the relationship is not so simple: this was seen in figure 6.6, the<br />

hysteresis loop. In this case, the relationship between H and B is no longer linear and will<br />

depend on the previous history <strong>of</strong> the material or the direction the magnet has been cycled.<br />

One may still determine the relative permeability through the normal magnetisation curve.<br />

This curve can be obtained by cycling the magnet in the same direction but reducing the<br />

peak magnetic field intensity and plotting the positions <strong>of</strong> P and P ′ such as in figure 6.7.<br />

The gradient <strong>of</strong> the curve at zero will provide a good estimate for µ.<br />

The work done in each cycle in changing the magnetisation <strong>of</strong> a specimen <strong>of</strong> volume τ is<br />

given by<br />

∮<br />

W = τ<br />

H.dB (6.18)<br />

where the integral is the area <strong>of</strong> the hysteresis loop.<br />

6.5.2.4 Magnetic circuits<br />

For the case <strong>of</strong> a transformer or a multi-turn device, Ampere’s circuital law can be written<br />

as


6.5. FERROMAGNETIC MATERIALS AND AC FIELDS 71<br />

Figure 6.7: Normal magnetisation curve.<br />

∮<br />

H.dl = Ni (6.19)<br />

where N is the number <strong>of</strong> turns on the coil. Hence for a transformer, if the cross-section<br />

A is large enough to make flux leakage negligible, then the magnetic field intensity will be<br />

constant around the loop, length L. Hence equation 6.19 reduces to HL = Ni.<br />

Figure 6.8: Dimensions <strong>of</strong> a transformer.<br />

The magnetic flux is defined by<br />

∫<br />

Φ = B.dA (6.20)<br />

For a transformer<br />

A<br />

Φ = µANi<br />

L<br />

(6.21)<br />

6.5.2.5 Magnetic induction<br />

If the magnetic flux density, B, is varied with time, such as with an AC field, an electric field,<br />

strength ǫ, encircling the magnetic field is induced satisfying


72 CHAPTER 6. MAGNETS AND MAGNETIC FIELDS<br />

∮<br />

ǫ.ds = − dΦ<br />

dt<br />

(6.22)<br />

where s is the path enclosing the surface through which magnetic flux <strong>of</strong> density B is passing.<br />

This is Faraday’s law and is a statement <strong>of</strong> the conservation <strong>of</strong> energy. For a transformer,<br />

this expression becomes<br />

V = −N dΦ<br />

dt<br />

(6.23)<br />

provided the same flux is passing through each <strong>of</strong> the N coils.<br />

6.5.3 The apparatus<br />

Figure 6.9 shows the AC circuit, which will be used to produce a hysteresis loop and determine<br />

the magnetic properties <strong>of</strong> the core material in the transformer.<br />

The transformer used in the AC fields experiments has a demountable core to facilitate measurement<br />

<strong>of</strong> the cross section and path length. Care must be taken to ensure that the surfaces<br />

mate properly and that no air gap is introduced during reassembly. The transformer has 540<br />

turns in the primary coil and 50 turns in the secondary coil.<br />

Figure 6.9: Circuit diagram for an AC magnetic field in the magnetic core <strong>of</strong> a transformer.<br />

Consider the circuit for the secondary coil. The arrangement <strong>of</strong> the resistor and capacitor<br />

is a low pass filter, but is similar in principle to a voltage divider where the ‘resistance’ <strong>of</strong><br />

the capacitor is now the impedance, Z c = 1/jωC, ω is the frequency and j 2 = −1 1 (see<br />

Ohanian [4], chapter 34, page 823 for a review <strong>of</strong> the theory).<br />

1 It is unfortunate that there are so few letters available to use as variable names - therefore throughout<br />

science, both i and j are used to mean the unit complex number. You will need to learn the appropriate<br />

meaning by context. In this experiment, i is the current.


6.5. FERROMAGNETIC MATERIALS AND AC FIELDS 73<br />

Pre-lab Question 6.2 Derive an expression for the voltage across the capacitor in terms <strong>of</strong><br />

the voltage across the secondary coil by considering the circuit as a voltage divider such as<br />

in figure 6.10.<br />

Figure 6.10: Voltage divider.<br />

An alternating current produces a time dependent voltage <strong>of</strong> the form V y = |V y |e −jωt .<br />

Pre-lab Question 6.3 Using the result from Pre-lab question 6.2 and differentiating equation<br />

6.20 and combining with 6.23 derive an expression for the magnetic flux density, B.<br />

R p is a variable resistor and can be used to qualitatively look at changing the shape <strong>of</strong> the<br />

loop.<br />

Pre-lab Question 6.4 Derive an expression for the magnetic field intensity with respect to<br />

the voltage across R p given by equation 6.19.<br />

Pre-lab Question 6.5 Explain why the display on the CRO is representative <strong>of</strong> a hysteresis<br />

loop.<br />

6.5.4 Experimental work<br />

SAFETY WARNING: LARGE VOLTAGES<br />

ARE USED AND MUST BE HANDLED<br />

CAREFULLY AND SLOWLY.<br />

• Wire up the circuit in figure 6.9 and obtain a hysteresis loop on the CRO. Comment on<br />

the shape <strong>of</strong> the loop and the effect <strong>of</strong> varying the resistor, R p .<br />

• Determine the values <strong>of</strong> the parameters used in the expressions for Pre-lab questions<br />

6.3 and 6.4 and use these to calibrate the horizontal and vertical axes on the CRO in<br />

suitable units.


74 CHAPTER 6. MAGNETS AND MAGNETIC FIELDS<br />

• Draw a graph <strong>of</strong> the observed loop and indicate the values <strong>of</strong> the points <strong>of</strong> saturation,<br />

remanence and coercive force.<br />

• Think <strong>of</strong> a way to obtain a normal magnetisation curve. Plot this curve.<br />

• From the magnetisation curve calculate the relative permeability <strong>of</strong> the core material.<br />

• Calculate the work done in the transformer for one magnetic cycle.<br />

• Calculate the magnetic polarisation, M, and compare it with the magnetic field strength.<br />

What does this say about the type <strong>of</strong> material used as the core <strong>of</strong> the transformer?<br />

6.6 Useful data<br />

Quantity<br />

Value<br />

Electron charge<br />

−1.6 × 10 −19 C<br />

Absolute permeability <strong>of</strong> free space 4π × 10 −7 Wb A −1 m −1<br />

Units <strong>of</strong> magnetic flux density Wb m −2 (Tesla)<br />

Units <strong>of</strong> magnetic field intensity A m −1<br />

Units <strong>of</strong> magnetic polarisation A m −1<br />

Units <strong>of</strong> magnetic flux<br />

Wb<br />

Diamagnetism χ m negative, µ r < 1<br />

Paramagnetism χ m positive, µ r > 1<br />

Ferromagnetism χ m large, µ r ≫ 1<br />

<strong>Bibliography</strong><br />

[1] P Lorrain, D P Corson, and F Lorrain. Electromagnetic Fields and Waves. W.H. Freeman<br />

and Co., 3 rd edition, 1988.<br />

[2] S G Starling. Electricity and Magnetism. Longmans, Green and Co. Ltd., 8 th edition,<br />

1953.<br />

[3] D Halliday, R Resnick, and J Walker. Fundamentals <strong>of</strong> <strong>Physics</strong>. Wiley, 6 th edition, 2003.<br />

[4] H C Ohanian. <strong>Physics</strong>. 2 nd edition, 1989.


Chapter 7<br />

Fundamental constants<br />

7.1 Introduction<br />

Over the next two days you will investigate various properties <strong>of</strong> electrons. The first day will<br />

be investigating the charge to mass ratio <strong>of</strong> the electron and the second day will look at the<br />

photoelectric effect. As a result <strong>of</strong> these two experiments you will then be in a position to<br />

present your own experimental values for some important fundamental constants.<br />

7.2 Charge to mass ratio <strong>of</strong> the electron<br />

7.2.1 Overview<br />

Faraday in 1833 demonstrated the quantisation <strong>of</strong> charge through a careful series <strong>of</strong> electrolysis<br />

experiments. Thomson ended a heated debate among physicists when he was able to<br />

identify the cathode rays seen in low pressure gas discharges as caused by negatively charged<br />

particles. These particles, electrons, were accepted as being fundamental constituents <strong>of</strong> matter,<br />

firmly banishing the notion that the atom was indivisible and fundamental. Thomson was<br />

able to measure the charge to mass ratio <strong>of</strong> this new particle with an experiment that made<br />

use <strong>of</strong> the same principles as the one you are about to perform. The value <strong>of</strong> e/m that Thomson<br />

determined was 1.0×10 11 C/kg. This is to be compared with the currently accepted value<br />

<strong>of</strong> (1.7588028 ± 0.0000054) × 10 11 C/kg. Since the properties <strong>of</strong> the electron depend almost<br />

solely on these two quantities we seek an accurate determination <strong>of</strong> their value. It would be<br />

<strong>of</strong> interest to compare the value <strong>of</strong> e/m that you calculate with Thomson’s first effort!<br />

7.2.2 Theory<br />

If a beam <strong>of</strong> electrons is accelerated through a known potential then their kinetic energy is<br />

given by the expression<br />

E k = 1 2 mv2 = eV (7.1)<br />

75


76 CHAPTER 7. FUNDAMENTAL CONSTANTS<br />

where m is the mass <strong>of</strong> the electron, e is the charge <strong>of</strong> the electron and V is the potential.<br />

If a magnetic field is applied perpendicular to the velocity <strong>of</strong> the electrons, they will then<br />

experience a Lorentz force mutually perpendicular to both the field, B, and the electron<br />

direction <strong>of</strong> motion (see the magnets experiment, section 6.3). The magnitude <strong>of</strong> this force<br />

is<br />

F = evB. (7.2)<br />

Since the electrons always experience a force perpendicular to their motion they move in a<br />

circle and thus experience a centripetal force with a magnitude given by<br />

F = mv2<br />

r<br />

(7.3)<br />

where r is the radius <strong>of</strong> the circle described by the electrons’ path.<br />

Pre-lab Question 7.1 What path would the electrons follow if their velocity were not totally<br />

perpendicular to the magnetic field (say at 45 ◦ )?<br />

Pre-lab Question 7.2 Using the equations above determine a formula for the ratio e/m<br />

in terms <strong>of</strong> the applied field strength, B, the radius <strong>of</strong> the circle, r, and the accelerating<br />

potential, V .<br />

7.2.3 Apparatus<br />

The apparatus are displayed schematically in figure 7.1.<br />

7.2.3.1 Helmholtz coils<br />

The magnetic field for this experiment is generated by a pair <strong>of</strong> Helmholtz coils with the<br />

magnitude <strong>of</strong> the field generated at a volume in their centre given by<br />

( ) 3/2 4 Nµ 0 I<br />

B =<br />

5 a<br />

(7.4)<br />

where N is the number <strong>of</strong> turns in each coil, µ 0 = 4π × 10 −7 T m A −1 is the permeability <strong>of</strong><br />

free space, I is the current through the coils and a is the radius <strong>of</strong> the coils. In this case the<br />

radius <strong>of</strong> the coils is 150 mm and each coil has 130 turns.<br />

7.2.3.2 The cathode ray tube<br />

The cathode ray tube is a 130 mm diameter helium filled glass bulb. This contains at its<br />

base the electron gun consisting <strong>of</strong> a heated cathode to produce the electrons that are then


7.2. CHARGE TO MASS RATIO OF THE ELECTRON 77<br />

Figure 7.1: The apparatus used to determine e/m<br />

accelerated towards the anode. The anode is made <strong>of</strong> a wire grill so that the electrons may<br />

pass through it and enter the body <strong>of</strong> the tube. The helium gas is retained at a pressure<br />

<strong>of</strong> 10 −2 Torr and is ionised by the electrons as they pass through it. As the helium ions<br />

recombine with their valence electrons to form atoms, they emit light, allowing the radius <strong>of</strong><br />

the electron beam to be determined.<br />

Pre-lab Question 7.3 What happens to the kinetic energy <strong>of</strong> the electrons as they ionise the<br />

helium? How will this affect the quality <strong>of</strong> your results?<br />

7.2.3.3 Mirrored scale<br />

A mirrored scale is attached to the rear coil and will be illuminated when power is provided<br />

to the apparatus. This may be used to align the beam and its image and so eliminate errors<br />

due to parallax.<br />

7.2.3.4 Cathode ray tube power supply<br />

A power supply is provided for the cathode ray tube. This provides 6 V AC to heat the<br />

cathode and a variable DC voltage to accelerate the electrons.


78 CHAPTER 7. FUNDAMENTAL CONSTANTS<br />

7.2.3.5 Helmholtz coil supply<br />

Current to the Helmholtz coils is supplied by a current regulating power supply capable <strong>of</strong><br />

providing 2 A at small voltages.<br />

Pre-lab Question 7.4 Why must both <strong>of</strong> these power supplies be kept as far from the cathode<br />

ray tube as possible?<br />

7.2.4 Experiment<br />

• Ensure that both the power supplies are connected as shown in figure 7.1. Before<br />

switching on the supplies check that the high voltage control and the current controls<br />

(supply and apparatus) are set to be minimum.<br />

• Switch on the supplies and allow the heater to warm up. Increase the accelerating<br />

potential until the electron beam can be seen. Turn the current control on the apparatus<br />

to maximum and adjust the current in the Helmholtz coils until the beam traverses a<br />

circle. If the beam travels in a spiral consult your demonstrator. Please note that the<br />

bulb is designed to rotate, so don’t be afraid to adjust the alignment, but carefully!<br />

• Adjust the accelerating potential to 300 V and adjust the Helmholtz coil current until<br />

the radius <strong>of</strong> the beam is 4.0 cm. Note the current, voltage and left and right radii.<br />

NOTE: Why are the two radii measurements different? How will this affect the way you<br />

measure radius?<br />

• Repeat this procedure at 5 V intervals until the beam no longer traverses a circle. Try<br />

as much as is practicable to adjust the current to keep the radius constant at 4.0<br />

cm. Detail the methods you employ to increase the accuracy <strong>of</strong> your results.<br />

7.2.5 Analysis<br />

Enter your data into the computer and use Excel to calculate the value <strong>of</strong> e/m for each<br />

voltage. Plot your results against potential.<br />

Pre-lab Question 7.5 How can you explain the shape <strong>of</strong> this graph?<br />

A better method for determining the value <strong>of</strong> e/m is to calculate the field variable. The field<br />

variable, χ, is a measure <strong>of</strong> the energy loss <strong>of</strong> the electrons due to their interactions with the<br />

gas. We may fit the function<br />

V =<br />

( e<br />

m)<br />

χ + δ (7.5)<br />

where e/m and δ are parameters and V and χ = [(Br) 2 /2] × 10 11 are variables. This allows<br />

a unique value <strong>of</strong> e/m to be determined while allowing for the effective potential loss due to


7.2. CHARGE TO MASS RATIO OF THE ELECTRON 79<br />

the presence <strong>of</strong> the gas with the parameter δ. Plot V against χ on another graph and overlay<br />

the fit. Hence determine a value for e/m and compare it to the accepted value.<br />

Pre-lab Question 7.6 Given the ionisation energy <strong>of</strong> He, how many electron-He collisions<br />

occur during the electron path through the tube on average?


80 CHAPTER 7. FUNDAMENTAL CONSTANTS<br />

7.3 The photoelectric effect<br />

7.3.1 Overview<br />

At the middle <strong>of</strong> the nineteenth century the wave theory <strong>of</strong> light was yet unchallenged. However,<br />

in 1887, Hertz noticed that a spark induced in his circuit was stronger when he deliberately<br />

illuminated the detector with UV light. Although his study concentrated on the<br />

properties <strong>of</strong> electromagnetic radiation, his was the first observation <strong>of</strong> the liberation <strong>of</strong> electrons<br />

from the clean surfaces <strong>of</strong> metals under the action <strong>of</strong> radiant energy. This is known as<br />

the photoelectric effect. His colleague, Lenard, later confirmed that the emitted carriers were<br />

negatively charged. However, it wasn’t until Einstein’s revolutionary postulate in 1905, that<br />

all the observed features <strong>of</strong> this effect were reconciled with mainstream theory.<br />

7.3.2 The failure <strong>of</strong> classical theory and Einstein’s postulate<br />

Experiment showed that electrons ejected from the surface had small but finite speeds ranging<br />

from zero to some maximum value. By making the collecting plate negatively charged<br />

with respect to the illuminated plate, they could measure the force required to stop the most<br />

energetic <strong>of</strong> electrons. This is how the stopping potential, V 0 , is defined.<br />

Pre-lab Question 7.7 Cite instances where classical theory failed to account for the observed<br />

features <strong>of</strong> this phenomenon. You may wish to consult the suggested references described<br />

at the end <strong>of</strong> these notes.<br />

Einstein extended the quantum theory <strong>of</strong> Planck to the radiation field itself, hypothesising<br />

that the light existed in quanta <strong>of</strong> energy with magnitude hν.<br />

Thus the mechanism <strong>of</strong> the photoelectric effect became as follows: an electron absorbs a<br />

photon <strong>of</strong> energy hν and attains enough energy to escape the surface. Energy not used to<br />

overcome the binding <strong>of</strong> the electron to the atom becomes kinetic energy. Thus<br />

E k = eV 0 = hν − W 0 (7.6)<br />

where E k is the kinetic energy <strong>of</strong> the photoelectrons and W 0 is the work function particular<br />

to the material that is illuminated. The work function is the energy required for the electron<br />

to escape the surface.<br />

7.3.3 Apparatus<br />

7.3.3.1 The mercury vapour lamp<br />

The spectral lamp you will use is a high intensity (100 W) mercury vapour lamp. Atoms are<br />

ionised by passing a current through the mercury vapour. The recombination <strong>of</strong> electrons


7.3. THE PHOTOELECTRIC EFFECT 81<br />

Table 7.1: Frequencies for the light spectrum issued from the mercury lamp.<br />

Colour Frequency (10 −1 PHz 1 )<br />

yellow 5.19<br />

green 5.49<br />

blue 6.88<br />

violet 7.41<br />

deep violet 8.22<br />

and mercury ions produces a light spectrum composed <strong>of</strong> discrete wavelengths. The emitted<br />

frequencies are displayed in table 7.1.<br />

7.3.3.2 The diffraction grating<br />

The light from the lamp will be dispersed using a transmission diffraction grating 2 and focussed<br />

with a lens. The grating equation is a sin θ m = mλ, where a is the distance between<br />

the ruled grooves, m is the order <strong>of</strong> interference and θ m is the angle through which the order<br />

is deflected. A diagram illustrates this below. Note that each order is resolved in space, as is<br />

each spectral line.<br />

Figure 7.2: The transmission diffraction grating with the orders <strong>of</strong> interference shown.<br />

7.3.3.3 The photoelectric cell<br />

You are supplied with a photoelectric cell for measuring the stopping potential. The schematics<br />

inside the photoelectric cell look roughly like figure 7.3. Light from the mercury lamp<br />

strikes the surface <strong>of</strong> the metal, and those photons with sufficient energy will liberate electrons<br />

from the metal surface, which will then be collected by the detector. Each electron<br />

collected will increase the overall negative charge at the detector, and hence make it harder<br />

for electrons to jump the gap. Eventually, the number <strong>of</strong> electrons will reach a maximum.<br />

1 1 PHz = 10 15 Hz<br />

2 see Hecht [3], section 10.2.7, page 465.


82 CHAPTER 7. FUNDAMENTAL CONSTANTS<br />

detector<br />

light<br />

electrons<br />

metal<br />

Figure 7.3: Light is incident on the metal surface, which liberates electrons, which are then<br />

collected by the detector.<br />

This will take a finite amount <strong>of</strong> time (called the stopping time, t s ) and will have an associated<br />

stopping voltage V 0 (as in equation 7.6). In this experiment, both V 0 and t s will be<br />

measured. The red discharge button on the photoelectric cell grounds the built-up electrons,<br />

ready for a new measurement.<br />

7.3.3.4 Filters<br />

Three filters are required for this experiment:<br />

• A filter with 5 strips <strong>of</strong> various transmission strengths (from 20% to 100%, in steps <strong>of</strong><br />

20%)<br />

• A yellow filter to prevent ambient light sources (UV from overhead fluorescent lights<br />

and also the violet line from the 3 rd order spectrum) from interfering with measurements<br />

<strong>of</strong> the stopping potential for the yellow line<br />

• A green filter for use with the green line for reasons cited above.<br />

All <strong>of</strong> these filters have magnetic strips for easy attachment to the photocell.<br />

7.4 Experiment<br />

A representation <strong>of</strong> the apparatus is given in figure 7.4 with the features you will need to be<br />

familiar with identified.<br />

• Allow the lamp to warm up for five minutes. This should give you the opportunity<br />

to identify the orders <strong>of</strong> interference produced by the transmission grating. Focus a<br />

spectral line <strong>of</strong> the 1 st order onto the white reflective mask <strong>of</strong> the photocell by moving<br />

the lens and grating back and forth. Fix the lens in place when you have obtained a<br />

sharp image.


7.4. EXPERIMENT 83<br />

Figure 7.4: The apparatus for the observation <strong>of</strong> the photoelectric effect.<br />

• Roll the light shield <strong>of</strong> the photocell (black tube) back and rotate the cell until the<br />

line is centred on the aperture <strong>of</strong> the photodiode. Tighten the locking screws on the<br />

photocell and grating.<br />

• Place the transmission filter on the front <strong>of</strong> the photocell to allow 100% <strong>of</strong> the light<br />

to enter the photodiode. If you have chosen the yellow or the green spectral lines<br />

remember to use the appropriate filters. Connect the voltmeter to the output <strong>of</strong> the<br />

photocell and switch the photocell on. Zero the photocell by pushing the red discharge<br />

button.<br />

Question 7.1 Record the stopping potential after a minute or so. Is it necessary to measure<br />

the stopping potential for all transmission filters? If not, why not?<br />

There are two effects that make the measurement <strong>of</strong> stopping time t s difficult. The first is the<br />

swiftness <strong>of</strong> the electron buildup - it is simply not viable to accurately measure time-spans<br />

<strong>of</strong> less than about a second by hand with a stopwatch. The second is the inaccuracy <strong>of</strong> the<br />

end-point. Note how the stopping potential is not constant, but instead fluctuates constantly.<br />

Certainly, in the act <strong>of</strong> measuring the stopping potential, it is difficult to tell when the final<br />

point is reached - and your choice <strong>of</strong> end-point will markedly affect the results. Instead, if<br />

we assume that the build-up is linear, we can instead choose some arbitrary factor κ 3 with<br />

which to multiply the stopping voltage, and then measure the time taken to reach κV 0 , which<br />

will be κt s . The factor κ needs to be small enough that the end-point is now well defined, yet<br />

3 0 < κ ≤ 1.


84 CHAPTER 7. FUNDAMENTAL CONSTANTS<br />

big enough that the length <strong>of</strong> time measured is at least a few seconds for 100% transmission.<br />

We suggest κ = 0.99. The quantity t s is restored by simply dividing κt s by κ.<br />

Zero the photocell again and record κt s . Repeat this procedure for all intensity filters. Readjust<br />

the apparatus for another spectral line and repeat the above procedure.<br />

Question 7.2 Explain what you observe in terms <strong>of</strong> operation <strong>of</strong> the photocell.<br />

Question 7.3 Make a plot <strong>of</strong> percent transmission versus stopping time t s for both the green<br />

and yellow lines. Is it linear? If not, why not?<br />

7.4.1 Variation <strong>of</strong> stopping potential with frequency<br />

Remove the transmission filter. Adjust the apparatus for all <strong>of</strong> the spectral lines you can<br />

observe (both 1 st and 2 nd order lines on both sides) and record the stopping potential for<br />

each line. Remember to use the appropriate filters. Consider also the effects <strong>of</strong> other light<br />

sources, for instance, overhead lighting.<br />

Question 7.4 What can you do to minimise these effects?<br />

Question 7.5 Average your results in Excel. Fit your data with Einstein’s equation. Determine<br />

h/e and W 0 from the parameters <strong>of</strong> this fit. Compare your result with the expected<br />

value <strong>of</strong> h/e = (4.135708 ± 0.000014) × 10 −15 J s C −1 . 4<br />

Question 7.6 Do your results agree better with quantum mechanics or classical mechanics?<br />

7.4.2 Calculation <strong>of</strong> the fundamental constants<br />

The third fundamental constant is introduced in the electron spin resonance experiment and<br />

is the Bohr magneton (µ b = e/2m e ). The value <strong>of</strong> this constant has been measured as<br />

(9.274096 ± 0.000065) × 10 −24 J T −1 .<br />

Now, use your results from the photoelectric effect determination <strong>of</strong> h/e, the previous practical<br />

exercise determination <strong>of</strong> e/m, and the result for the Bohr magneton cited, to calculate<br />

the values <strong>of</strong> h, m e and e. You should determine errors for each <strong>of</strong> these quantities so that<br />

you can effect a meaningful comparison with the accepted values <strong>of</strong> these constants.<br />

4 I know I know, we keep saying that errors should only be quoted to one significant figure. However, when<br />

a number has been measured using a huge number <strong>of</strong> readings, we sometimes give a second significant figure,<br />

especially if the first is a 1.


7.5. USEFUL DATA 85<br />

7.5 Useful data<br />

Quantity<br />

Value<br />

h<br />

6.6261 ×10 −34 J s<br />

e<br />

1.6022 ×10 −19 C<br />

m e<br />

9.1096 × 10 −34 kg<br />

e/m<br />

(1.7588028 ± 0.0000054) × 10 11 C kg −1<br />

h/e<br />

(4.135708 ± 0.000014) × 10 −15 J s C −1<br />

µ 0 4π × 10 −7 T m A −1<br />

µ b (9.274096 ± 0.000065) × 10 −24 J T −1<br />

ionisation energy <strong>of</strong> He I He 24.58 e V<br />

<strong>Bibliography</strong><br />

[1] R A Serway, C J Moses, and C A Moyer. Modern <strong>Physics</strong>. Saunders, 1989.<br />

[2] D Halliday, R Resnick, and J Walker. Fundamentals <strong>of</strong> <strong>Physics</strong>. Wiley, 6 th edition, 2003.<br />

[3] E Hecht. Optics. Addison-Wesley, 3 rd edition, 1998.


Chapter 8<br />

Waves in waveguides<br />

8.1 Introduction<br />

In the following two experiments you will be studying the transmission <strong>of</strong> waves along<br />

waveguides. Even though sound waves and microwaves are physically very different, the<br />

fact that they are wave phenomena means that their transmission in waveguides may be<br />

treated with similar mathematical formalisms.<br />

The background theory for this experiment covers a lot <strong>of</strong> ground, and is quite mathematical.<br />

Don’t be intimidated by this - if you get confused, remind yourself <strong>of</strong> your basic knowledge<br />

<strong>of</strong> how standing waves behave in tubes 1 , and try to match the mathematics that is being<br />

presented with your prior understanding.<br />

8.2 Acoustics<br />

8.2.1 Theory<br />

Acoustics is defined as the study <strong>of</strong> vibration and sound. Sound is the propagation <strong>of</strong> elastic<br />

disturbances in a continuous medium; vibration refers to such disturbances in more simple<br />

systems, such as springs. You will be familiar with the situation <strong>of</strong> particles being displaced<br />

from their equilibrium positions, developing potential energy, and then being restored to<br />

their equilibrium positions, releasing energy. We may consider a particle that is displaced<br />

repeatedly such that it has the same displacement and velocity some number <strong>of</strong> times per<br />

second. If the displacements being considered are infinitesimal we may then consider differential<br />

elements <strong>of</strong> displacement. Such a treatment yields the wave equation for particle<br />

displacement.<br />

1<br />

c 2 ∂ 2 ξ<br />

∂t 2 = ∇2 ξ (8.1)<br />

Here ξ is the particle displacement and c is the wave velocity, defined as the speed at which<br />

1 see Ohanian [7], chapter 17 page 438.<br />

87


88 CHAPTER 8. WAVES IN WAVEGUIDES<br />

Quantity Symbol Quantity Symbol<br />

Particle displacement ξ Absolute pressure P<br />

Angular frequency ω = 2πf Equilibrium density ρ 0<br />

Sound pressure p Equilibrium pressure P 0<br />

Particle velocity v Bulk modulus B<br />

Wave velocity c Characteristic impedance Z 0 = ρ 0 c<br />

Wavelength λ<br />

Table 8.1: Quantities useful in characterising sound<br />

Real part<br />

ξ = A cos(ωt) + B sin(ωt)<br />

v = ∂ξ<br />

∂t<br />

= ω[−A cos(ωt) + B sin(ωt)] v<br />

a = ∂2 ξ<br />

∂t 2 = −ω 2 [A cos(ωt) + B sin(ωt)]<br />

Complex notation<br />

ξ = Ce iωt ,C = A − iB<br />

= ∂ξ<br />

∂t = iωCeiωt = iωξ<br />

a = ∂2 ξ<br />

∂t 2 = −ω 2 Ce iωt = −ω 2 ξ<br />

Table 8.2: Complex notation for the solution to the wave equation<br />

the wavefront moves through the medium. This equation has a solution <strong>of</strong> the form:<br />

ξ = A cos(ωt) + B sin(ωt) (8.2)<br />

describing simple harmonic motion, or pure tone. A and B are constants and ω is the<br />

angular frequency. Because any sound can be represented as a sum <strong>of</strong> simple harmonic<br />

terms the following discussion will be limited to single frequencies 2 .<br />

The quantities <strong>of</strong> displacement and wave velocity have already been encountered. Other<br />

useful terms used to characterise sound are given in table 8.1.<br />

If the motion is simple harmonic it is better to introduce complex notation 3 to represent the<br />

quantity ξ. In this case, as may be seen from table 8.2, ξ is a complex number, which is much<br />

easier to manipulate than the real or imaginary parts alone.<br />

8.2.1.1 Wave velocity, c<br />

The wave velocity c has already been encountered in the wave equation <strong>of</strong> motion. For media<br />

such as air or water, c is independent <strong>of</strong> frequency. For an ideal non-viscous fluid c may be<br />

related to the average or equilibrium density, ρ 0 via the relation<br />

2 We can use Fourier’s theorem to crack more complicated cases.<br />

3 It would be nice if we could choose standard meanings for i and j, e.g. set i 2 = −1, and use j for current.<br />

Sadly, the meanings <strong>of</strong> these individuate letters needs to be obtained from the context. j will sometimes be<br />

imaginary, a current density or something else. This is true <strong>of</strong> many letters in physics, so beware!


8.2. ACOUSTICS 89<br />

c =<br />

√<br />

B<br />

ρ 0<br />

(8.3)<br />

where B is the adiabatic bulk modulus. The bulk modulus measures how the pressure <strong>of</strong><br />

the gas changes, when the volume <strong>of</strong> the container which the gas occupies is changed. Bulk<br />

modulus can also be thought <strong>of</strong> as the inverse <strong>of</strong> the compressibility <strong>of</strong> a gas. For ideal gases,<br />

the bulk modulus is proportional to the average pressure P 0 .<br />

So the Laplace expression is obtained<br />

B = γP 0 (8.4)<br />

c =<br />

√<br />

γP 0<br />

ρ 0<br />

(8.5)<br />

which is a standard relation for two gases, where γ is a ratio <strong>of</strong> heat capacities <strong>of</strong> the two<br />

gases. This particular invocation <strong>of</strong> Laplace’s equation is for air at standard temperature and<br />

pressure (1 atm and 0 ◦ C), c = 341 ms −1 .<br />

Given an ideal gas we may invoke the gas equation in the form<br />

P = NkT (where N = n V ) (8.6)<br />

where N is the number density <strong>of</strong> particles, k is Boltzmann’s constant and P and T are the<br />

equilibrium pressure and temperature respectively. If we write the density, ρ as Nµ where<br />

µ is the average mass <strong>of</strong> a single molecule it becomes apparent that c may be expressed as<br />

c =<br />

√<br />

γkT<br />

µ<br />

(8.7)<br />

Pre-lab Question 8.1 Derive an expression for c 0 , the wave velocity at 0 ◦ C, given that you<br />

know the wave velocity at a temperature T . (Note: the expression should be in terms <strong>of</strong> T ,<br />

T 0 and c T only).<br />

Pre-lab Question 8.2 Calculate the wave velocity <strong>of</strong> sound in hydrogen given that<br />

µ air = 29.0 a.m.u.<br />

µ hydrogen = 2.016 a.m.u.<br />

at STP. Remember that a.m.u. stands for atomic mass unit. What is one a.m.u. when expressed<br />

in kg?<br />

The expression for the root-mean-square (RMS) particle velocity <strong>of</strong> molecules in random<br />

thermal motion is


90 CHAPTER 8. WAVES IN WAVEGUIDES<br />

v RMS =<br />

√<br />

3kT<br />

µ<br />

(8.8)<br />

Pre-lab Question 8.3 Calculate v RMS for hydrogen at standard temperature. Can you remark<br />

on any similarity between the wave velocity and the RMS velocity?<br />

The speed <strong>of</strong> sound may also be calculated if the frequency and wavelength have been<br />

determined via<br />

c = λf. (8.9)<br />

However, a correction must be made for standing waves in a tube as compared to plane<br />

waves in free space. This is both due to the mechanical friction and heat conduction. The<br />

expression<br />

c T =<br />

c<br />

1 − 0.37<br />

d √ ω<br />

(8.10)<br />

has been theoretically determined and empirically confirmed for air in a brass tube. Here c T<br />

is the speed <strong>of</strong> sound in free space at temperature T , d is the diameter <strong>of</strong> the tube in cm,<br />

and c is the speed <strong>of</strong> sound in the tube at that temperature.<br />

For plane acoustic waves the vector particle velocity is parallel to the direction <strong>of</strong> propagation<br />

<strong>of</strong> the wave. Such waves are called longitudinal (compressional or irrotational).<br />

8.2.1.2 Pressure<br />

Most acoustics experiments measure the sound or acoustic pressure, p, defined as<br />

p = P − P 0 (8.11)<br />

When there are no sound waves travelling in the tube, the pressure inside the tube is the<br />

average or equilibrium pressure, equal at every point inside the tube. However, when a<br />

standing wave exists inside the tube, there will be regions <strong>of</strong> over-density and under-density.<br />

P gives the instantaneous pressure at a point, and so p can be either positive or negative,<br />

depending on the position in the tube.<br />

For a plane wave travelling in the positive x-direction we solve a wave equation for pressure<br />

(analogous to the equation for displacement, equation 8.1) to obtain<br />

p + = a + e i(ωt−kx) (8.12)<br />

Here a + is a constant and k is the wavenumber, equal to 2π/λ. We may define the characteristic<br />

acoustic impedance, Z 0 <strong>of</strong> the material as the constant <strong>of</strong> proportionality linking<br />

pressure to particle velocity


8.2. ACOUSTICS 91<br />

Acoustic<br />

Electric<br />

Pressure P Voltage V<br />

Particle velocity V Current I<br />

Intensity I = PV Power P = IV<br />

Acoustic impedance Z = P/V Electrical impedance Z = V/I<br />

Table 8.3: This table makes an analogy between acoustic and electric parameters.<br />

p + = Z 0 v + (8.13)<br />

Z 0 is a real number that is a function only <strong>of</strong> the physical properties <strong>of</strong> the medium in which<br />

the wave travels. Similarly, for a plane wave travelling in the negative x-direction<br />

p − = a − e i(ωt+kx) = −Z 0 v − (8.14)<br />

Ideas <strong>of</strong> impedance are here introduced for reasons analogous to their use in electrical circuit<br />

theory. A list <strong>of</strong> such analogues is given in table 8.3.<br />

Let us conduct two thought experiments.<br />

The first is with electrical transmission lines. If we have two different wires joined together<br />

end-to-end, then if the two wires have similar impedances, a large amount <strong>of</strong> power will be<br />

transferred when electricity travels across the interface between the two wires. If however<br />

the impedances <strong>of</strong> the two wires are very different, only a little power will be transferred.<br />

Similarly, let us think about sound waves in a tube. The first medium is the air, and the<br />

second medium is the end piece <strong>of</strong> the tube (for example, lead or foam). The foam will have<br />

a very similar impedance to the air (much more so than the lead), and hence a large amount<br />

<strong>of</strong> power will be transferred when the sound wave hits the foam. Therefore, only a little <strong>of</strong><br />

the sound wave will be reflected. With lead however, the impedance will be very different<br />

than air, and so very little power is transferred into the lead, and most <strong>of</strong> it is reflected.<br />

8.2.2 Stationary wave theory and the reflection <strong>of</strong> plane acoustic waves<br />

In practice a stationary wave system is due to the partial reflection <strong>of</strong> a travelling wave at a<br />

discontinuity in the medium. The study <strong>of</strong> such systems is most conveniently carried out in<br />

a tube. The advantage <strong>of</strong> studying wave phenomena in a tube is that the only type <strong>of</strong> wave<br />

propagated is a plane wave, despite the size <strong>of</strong> the source. This allows us to ignore any<br />

curvature.<br />

Consider the situation shown in figure 8.1.<br />

A plane wave travelling in the positive x-direction is partially reflected and partially transmitted<br />

at a plane <strong>of</strong> discontinuity in the medium. Since the frequency is preserved upon<br />

reflectance we may write the displacement as


92 CHAPTER 8. WAVES IN WAVEGUIDES<br />

Figure 8.1: Partial reflection at normal incidence<br />

ξ + = a + e i(ωt−kx) (8.15)<br />

ξ − = a − e i(ωt+kx) (8.16)<br />

The real part <strong>of</strong> the resultant displacement is<br />

ξ = Re(ξ + + ξ − )<br />

= a + (1 + r) cos(ωt) cos(kx) + a + (1 − r) sin(ωt) sin(kx) (8.17)<br />

where<br />

r = a −<br />

a +<br />

= A max − A min<br />

A max + A min<br />

= Z 2 − Z 1<br />

Z 2 + Z 1<br />

(8.18)<br />

A max and A min are the displacement maxima and minima respectively, and Z 1 and Z 2 are the<br />

absolute values <strong>of</strong> the impedances - where Z 1 is the impedance <strong>of</strong> the original medium the<br />

wave is travelling in, and Z 2 is the medium which is being ‘reflected <strong>of</strong>f’. The envelope, or<br />

standing wave pr<strong>of</strong>ile for r = 1/3 is shown in figure 8.2. Note that the resultant envelope is<br />

the sum <strong>of</strong> two stationary waves with amplitudes a + (1 + r) and a + (1 − r) respectively, each<br />

with a period <strong>of</strong> 2π/ω, and have a relative displacement <strong>of</strong> maxima (and minima) <strong>of</strong> λ/4.<br />

You will be more familiar with the situations arising from an open or closed tube. In these<br />

cases, A min ≈ 0, giving an r <strong>of</strong> about 1. However, in the case <strong>of</strong> a closed tube, a stationary<br />

wave system is observed with displacement nodes located at λ/2, λ, 3λ/2 etc. from the<br />

end. In the situation with the open end, the displacement nodes are located at λ/4, 3λ/4,<br />

5λ/4 etc. from the end. These observations suggest that in order to completely characterise a<br />

stationary wave system not only a magnitude factor, but a phase factor needs to be supplied.


8.2. ACOUSTICS 93<br />

Figure 8.2: The resultant standing wave with r=1/3<br />

ξ − = a − e i(ωt+kx−φ) (8.19)<br />

Now the real part <strong>of</strong> the resultant displacement is<br />

ξ = Re(ξ + + ξ − )<br />

= a +<br />

[<br />

cos(ωt) cos(kx) + a+ sin(ωt) sin(kx) (8.20)<br />

+a − cos(ωt) cos(kx) cos(φ) − r cos(ωt) sin(kx) sin(φ)<br />

− r sin(ωt) sin(kx) cos(φ) − r sin(ωt) sin(kx) sin(φ) ]<br />

A stationary wave produced by partial reflectance/transmittance at a discontinuity in the<br />

medium is completely characterised by the complex reflection coefficient, K. K embodies<br />

both ideas about magnitude; it has magnitude r, and phase φ (see the appendix, section 8.4.1,<br />

for the derivation):<br />

K = re iφ = K ′ + iK ′′ , K ′ = r cos φ, K ′′ = r sin φ (8.21)<br />

So<br />

ξ = a + (1 + K ′ ) cos(ωt) cos(kx) + a + (1 − K ′ ) sin(ωt) sin(kx) (8.22)<br />

−a + K ′′[ sin(ωt) cos(kx) + cos(ωt) sin(kx) ]<br />

Pre-lab Question 8.4 There are two ways to express K, as re iφ , and K ′ + iK ′′ . Explain<br />

which way you think is most relevant for understanding the physics <strong>of</strong> the situation.


94 CHAPTER 8. WAVES IN WAVEGUIDES<br />

We may calculate the RMS <strong>of</strong> the real part <strong>of</strong> the resultant displacement pr<strong>of</strong>ile upon reflection<br />

at a termination to calculate the theoretical standing wave pr<strong>of</strong>ile:<br />

√<br />

1<br />

τ<br />

∫ 0<br />

Using this average, this quantity is easily shown to be<br />

where<br />

ξ RMS =<br />

τ<br />

[Re(ξ)] 2 dt (8.23)<br />

√<br />

(A RMS<br />

MAX )2 cos 2 [k(x − x 0 )] + (A RMS<br />

MIN )2 sin 2 [k(x − x 0 )] (8.24)<br />

A RMS = A √<br />

2<br />

, φ = 2kx 0 (8.25)<br />

which we shall later use to plot against a measured wave pr<strong>of</strong>ile.<br />

Figure 8.3: The resultant standing wave with phase for equation 8.24<br />

Equation 8.24 is completely defined by 4 parameters, where we can choose 2 <strong>of</strong> the 3 out<br />

<strong>of</strong> φ, k and x 0 (these are linked by equation 8.25) and 2 <strong>of</strong> the 3 out <strong>of</strong> r, A MAX and A MIN<br />

(these are linked by equation 8.18). Let us choose A RMS<br />

MAX , r, x 0 and k, three possible plots are<br />

displayed in figure 8.3, with choices <strong>of</strong> (0.5, 0.1, 0, 1) for the dashed line, (1,0.5,1,1) for the<br />

solid line and (0.5,0.5,2,1) for the dash-dotted line. By inspection <strong>of</strong> these it can be seen that<br />

x 0 is the distance between the discontinuity and the first minimum pressure (maximum<br />

displacement) node.<br />

K is real when x 0 = 0,λ/4,λ/2 etc. K is imaginary when x 0 = λ/8, 3λ/8, 5λ/8 etc.<br />

Pre-lab Question 8.5 Calculate K for the open and closed tubes. Remember that x 0 is<br />

defined as the distance from the discontinuity to the first pressure node. Recall also that<br />

pressure and displacement are π/2 out <strong>of</strong> phase, that is, a displacement node is a pressure<br />

antinode.


8.2. ACOUSTICS 95<br />

Figure 8.4: Experimental setup.<br />

The analogous expression to equation 8.24 for pressure is<br />

p RMS =<br />

√ [<br />

(A RMS<br />

MAX )2 cos 2 k(x − x 0 ) − π ]<br />

2<br />

[<br />

+ (A RMS<br />

MIN )2 sin 2 k(x − x 0 ) − π ]<br />

2<br />

(8.26)<br />

You will be measuring the pressure standing wave pr<strong>of</strong>ile with a microphone, as in figure<br />

8.4.<br />

8.2.3 Apparatus<br />

1 sound tube, with cart attached, on a ruler base<br />

1 signal generator<br />

1 CRO<br />

3 BNC cables<br />

1 BNC cable adaptor<br />

various ends for the sound tube<br />

8.2.4 Experiment<br />

8.2.4.1 Setup<br />

• Set up your equipment as in figure 8.5, using the BNC cables. Note:<br />

– On the signal generator, plug into “Output 50 Ω”, not “Output pulse”<br />

– For the first part <strong>of</strong> the experiment, keep the amplitude on the signal generator<br />

just below what is audible, to avoid a headache.<br />

– Check that the triggers on the CRO are both set to “AC”.<br />

• Check first that the signal coming from the signal generator is correctly read by the<br />

CRO. Ensure that you can read the scale in the vertical direction correctly (the “volts/div”<br />

knob controls this, it gives the volts per big division on the screen).


96 CHAPTER 8. WAVES IN WAVEGUIDES<br />

Figure 8.5: Connection diagram<br />

• Switch to channel B, and get the signal clearly on the screen. Roll the cart slowly up<br />

and down the rails, and notice how you get a varying function <strong>of</strong> voltage with respect<br />

to position <strong>of</strong> the cart. Notice also how the amplitude is not zero at the nodes, but only<br />

very close to (there is no real “zero” <strong>of</strong> a continuous quantity in experimental physics).<br />

8.2.4.2 Measurement <strong>of</strong> wave velocity<br />

Question 8.1 Set the signal generator to 500 Hz. Determine the positions <strong>of</strong> the minima in<br />

pressure along the sound tube, by reading <strong>of</strong>f the ruler. The difference between two minima<br />

is half <strong>of</strong> one wavelength - how can you maximise the accuracy <strong>of</strong> this measurement?<br />

(a) Determine the wavelength <strong>of</strong> the standing wave in the tube.<br />

(b) Calculate the speed <strong>of</strong> sound in the tube, using the above value.<br />

(c) Calculate the theoretical speed <strong>of</strong> sound at 298 K, given that the CRC Handbook gives<br />

c 0 = 331.45 m s −1 . (Hint: use your answer to the first Pre-lab question).<br />

(d) Compare the speed <strong>of</strong> sound in the tube with the speed <strong>of</strong> sound in free space.<br />

Repeat parts 1 - 4 for two other frequencies, both in the range 1000-3000 Hz.<br />

8.2.4.3 Standing wave pr<strong>of</strong>iles<br />

In this section, we will measure the entire pr<strong>of</strong>ile <strong>of</strong> the wave - not just the positions <strong>of</strong> the<br />

minima. The response <strong>of</strong> the microphone and speaker is dependent on the amplitude <strong>of</strong> the<br />

signal generator output. It is important that all your measurements are made in a region<br />

where the response is linear. As a result, you need to adjust the speaker level on the signal<br />

generator, so that the CRO reads 1.0 V peak-to-peak from channel A, and keep it at this level<br />

for the rest <strong>of</strong> the experiment.<br />

Obtain the standing wave pr<strong>of</strong>ile at 1000 Hz for a CLOSED tube. Measurements should be<br />

taken from the END (cart pushed in all the way) <strong>of</strong> the tube in 2cm steps, with an extra point<br />

taken at the precise position <strong>of</strong> each minima


8.2. ACOUSTICS 97<br />

Question 8.2 Why do you think this is necessary for minima only and not maxima? This<br />

will become clearer once the pr<strong>of</strong>ile has been plotted.<br />

Make a plot <strong>of</strong> the data you have taken. Now, using the measured values <strong>of</strong> A max , A min ,<br />

λ and x 0 , overlay the theoretical plot (equation 8.26), and see how closely they match (DO<br />

NOT try to ‘fit’ for any parameters, simply compare the two!).<br />

You have been supplied with a range <strong>of</strong> end pieces for the sound tube, attach a few more and<br />

repeat steps 1 and 2.<br />

For each absorber, calculate the reflection coefficient K, from measurements <strong>of</strong> A max , A min ,<br />

λ, and x 0 .<br />

Question 8.3 Make a summary table, and comment on the physics.<br />

Question 8.4 (Advanced) Find out the impedance <strong>of</strong> air, and using r = Z 2−Z 1<br />

Z 2 +Z 1<br />

and your<br />

knowledge <strong>of</strong> r, determine Z 2 for each end piece.


98 CHAPTER 8. WAVES IN WAVEGUIDES<br />

8.3 Microwaves<br />

8.3.1 Introduction<br />

In the laboratory we generally use only low frequency signals that are conveniently routed<br />

with wires or a coaxial cable. However the use <strong>of</strong> wires becomes intractable in the high<br />

frequency regime where field radiation in the space around the wires results in intolerable<br />

energy losses. In the high frequency realm guided microwaves are more appropriate for<br />

conveying electromagnetic signals. At 10 kHz the attenuation <strong>of</strong> a signal in a waveguide<br />

is 1/4 that in an equivalent coaxial cable. At 1 MHz and above, a coaxial cable has no<br />

application.<br />

8.3.2 Theory<br />

It should be no surprise that electromagnetic radiation may be propagated in a hollow metal<br />

tube. As Feynman facetiously remarks, if the tube is straight, we can see through it! 4 To<br />

understand pictorially how microwaves are propagated in a guide, refer to figure 8.6 below.<br />

Figure 8.6: Generation <strong>of</strong> electromagnetic disturbance in a guide<br />

Imagine that an oscillator exists and is connected to the end <strong>of</strong> the tube causing a deficiency<br />

4 Feynman lectures [4], Volume II section 24-2.


8.3. MICROWAVES 99<br />

<strong>of</strong> electrons on the bottom wall and a surplus on the top. An electric field exists between<br />

the top and the bottom walls <strong>of</strong> the conductor causing current to flow as shown in figure<br />

8.6(a). In this figure, magnetic fields trying to encircle these currents are prevented because<br />

<strong>of</strong> the presence <strong>of</strong> the solid metal wall, so they join up instead, as shown in figure 8.6(b). As<br />

the oscillator reverses its polarity, the magnetic fields generate other currents by induction<br />

further along the guide and so the disturbance is propagated. One cycle <strong>of</strong> this disturbance<br />

is shown in figure 8.6(c). A more mathematical treatment is required to extract all the interesting<br />

observable parameters from this phenomenon, specifically, what kind <strong>of</strong> waves can<br />

exist in a rectangular pipe? For electromagnetic radiation we can expect that both electric<br />

and magnetic fields will be perpendicular to the direction <strong>of</strong> propagation and that any mathematical<br />

form expressing these fields should satisfy Maxwell’s equations. We may construct<br />

the coordinate axis as shown in figure 8.7.<br />

Figure 8.7: The axes for the metal tube, including the dimensions <strong>of</strong> the rectangular tube, a<br />

and b.<br />

Since we expect no electric field in the z direction we may expect some variation in the<br />

electric field E across x or y. This function must satisfy the condition that the electric field<br />

goes to zero at the sides <strong>of</strong> the conductor since the currents and charges in a conductor<br />

conspire so that there is no tangential component <strong>of</strong> E at the surface. So we may have E y<br />

varying with x as shown in figure 8.8.<br />

Figure 8.8: The lowest mode for the y component <strong>of</strong> the electric field in the x direction.<br />

For a rectangular geometry, the waveform is generally simple harmonic so we may guess<br />

that the wave would have some form<br />

E y = E 0 sin(k x x)e i(ωt−kzz) (8.27)


100 CHAPTER 8. WAVES IN WAVEGUIDES<br />

The field is perpendicular to the top and bottom conductor surfaces and zero at the side walls<br />

if we choose<br />

k x a = nπ (8.28)<br />

where n is any integer. The z dependence indicates a wave travelling in the z direction with<br />

frequency ω and wave velocity v, where<br />

v = ω k z<br />

. (8.29)<br />

The divergence <strong>of</strong> E must be zero within the free space inside the conductor (by Maxwells<br />

equation). In this case E has only a y component, and it doesn’t change with y, so<br />

⃗∇. ⃗ E = 0 (8.30)<br />

is satisfied. The other Maxwell equations must also be satisfied, specifically the wave equation<br />

⃗∇ 2 E ⃗<br />

∂ 2 E y =<br />

∂x + ∂2 E y<br />

2 ∂y + ∂2 E y<br />

2 ∂z 2<br />

Unless E y is zero everywhere this equation specifies that<br />

= 1 c 2 ∂ 2 E y<br />

∂t 2 (8.31)<br />

(−k 2 xE y ) + 0 + (−k 2 zE y ) = − ω2<br />

c 2 E y (8.32)<br />

k 2 x + k 2 z − ω2<br />

c 2 = 0 (8.33)<br />

k z =<br />

√ (ω ) ( )<br />

2 n2 π<br />

−<br />

2<br />

c 2 a 2<br />

(8.34)<br />

k x is already fixed, so<br />

k z = 2π<br />

λ g<br />

(8.35)<br />

where λ g is the guide wavelength.<br />

We may define the free space wavelength as<br />

λ 0 = 2πc<br />

ω<br />

(8.36)<br />

and so


8.3. MICROWAVES 101<br />

4π 2<br />

λ 2 g<br />

= 4π2<br />

λ 2 0<br />

− n2 π 2<br />

a 2 (8.37)<br />

1<br />

λ 2 g<br />

= 1 λ 2 0<br />

− n2<br />

4a2. (8.38)<br />

This is the fundamental relationship between free space and guided wavelengths.<br />

Associated with the electric fields there will be magnetic fields. Since<br />

c 2 ∇ × B = ∂E<br />

∂t , (8.39)<br />

the lines <strong>of</strong> B will circulate around the regions where ∂E/∂t is largest, that is, between the<br />

maxima and minima <strong>of</strong> E.<br />

Figure 8.9: Variation <strong>of</strong> E and B in the waveguide.<br />

We have assumed that k z has only one root but it should properly have two, a positive and<br />

negative one for the wave travelling in either direction.<br />

√ (ω )<br />

2<br />

k z = ± −<br />

c 2<br />

(<br />

n2 π 2<br />

a 2 )<br />

(8.40)<br />

We have also assumed that k z is real but if<br />

ω < nπc<br />

a<br />

(8.41)<br />

then the term in the square root is negative and we apparently have no solution anymore.<br />

However if we define


102 CHAPTER 8. WAVES IN WAVEGUIDES<br />

with<br />

k z = ±ik ′ (8.42)<br />

k ′ =<br />

√ (n2 ) ( )<br />

π 2 ω<br />

2<br />

−<br />

a 2 c 2<br />

(8.43)<br />

Then for E y we have<br />

E y = E 0 sin(k x x)e i(ωt±ik′ z)<br />

(8.44)<br />

= E 0 sin(k x x)e ±k′z e iωt (8.45)<br />

Thus E y is seen to oscillate with time, but to decay smoothly with z as a real exponential.<br />

This tells us that for<br />

ω c = nπc<br />

a<br />

(8.46)<br />

or below, waves do not propagate down the guide, but decay from the source exponentially<br />

with k ′ . It is for this reason that ω c is called the cut-<strong>of</strong>f frequency and<br />

λ c = 2a n<br />

(8.47)<br />

is called the cut-<strong>of</strong>f wavelength.<br />

We may thus recast the relationship between free space and guide wavelength as<br />

1<br />

λ 2 g<br />

= 1 λ 2 0<br />

− 1 λ 2 c<br />

(8.48)<br />

If operation is below the cut-<strong>of</strong>f frequency the wave is said to be evanescent.<br />

8.3.3 Modes in the guide<br />

We have been working with the transverse electric mode TE n0 . However in general, both<br />

n and m may vary, and we have TE nm , where n denotes the number <strong>of</strong> full variations <strong>of</strong> the<br />

electric field along the broad dimension, and m denote the number <strong>of</strong> full variations along<br />

the narrow dimension. These correspond to other solutions <strong>of</strong> the wave equations.<br />

Pre-lab Question 8.6 Draw the TE 10 , TE 01 , TE 11 ,and TE 20 modes (use figure 8.10 as a<br />

guide to what is required, but remember you are illustrating electrical rather than magnetic<br />

fields).


8.3. MICROWAVES 103<br />

If in a mode TE mn the integers m and n are both non-zero, then the equation for the cut-<strong>of</strong>f<br />

wavelength generalises to<br />

λ mn<br />

c =<br />

√ (n<br />

a<br />

2<br />

) 2<br />

+<br />

( m<br />

b<br />

) 2<br />

(8.49)<br />

All modes with wavelength longer than or equal to λ c the cut-<strong>of</strong>f wavelength exist simultaneously<br />

in the cavity. Thus the dimensions <strong>of</strong> the cavity can be tailored to exclude all modes<br />

but one, or one or two, etc.<br />

TM (transverse magnetic) modes are also possible. Some simple ones are shown in figure<br />

8.10.<br />

Figure 8.10: Simple TM modes.<br />

In TM modes a kind <strong>of</strong> propagation is produced whereby no component <strong>of</strong> the magnetic field<br />

is in the direction <strong>of</strong> propagation, while there is some component <strong>of</strong> the electric field in the<br />

z direction.<br />

8.3.4 Apparatus<br />

8.3.4.1 The Gunn diode<br />

In this experiment, you can identify the Gunn diode as a metal box with a clear plastic top,<br />

inside <strong>of</strong> which you can see some electronics. It is a source <strong>of</strong> microwaves, for details on<br />

how this works, see the appendix (section 8.4.2).<br />

8.3.4.2 The cavity wavemeter<br />

The cavity wavemeter is used to measure the wavelength <strong>of</strong> the standing wave in the<br />

waveguide. The cavity wavemeter consists <strong>of</strong> a cavity terminated by a fixed short circuit and<br />

a moveable short circuit coupled to the main waveguide as shown in figure 8.11.<br />

If a short circuit is placed across a cavity, then at that point the voltage must be zero. Thus<br />

the only wave that can exist in the wavemeter is one that has a voltage minimum at both<br />

ends. Some <strong>of</strong> the allowable modes are shown in figure 8.12.<br />

In this case the wavemeter length is tuned so that only the first mode is possible.<br />

To obtain a reading <strong>of</strong> the wavelength <strong>of</strong> the standing wave in the main guide the micrometer<br />

is adjusted. When this is moved to the correct position the cavity resonates at the frequency


104 CHAPTER 8. WAVES IN WAVEGUIDES<br />

Figure 8.11: Schematic <strong>of</strong> a cavity wavemeter<br />

Figure 8.12: Allowable modes in the cavity wavemeter<br />

<strong>of</strong> the radiation in the guide, sucking energy from the propagation in the main guide. This is<br />

detected as a resonance dip in the standing wavemeter (the reading on the multimeter lowers<br />

suddenly). At this point, the wavelength (in centimetres) may be read <strong>of</strong>f directly from the<br />

red scale on the micrometer, corresponding to a previous calibration. The wavemeter should<br />

be taken <strong>of</strong>f resonance (detuned) after measurement to increase the available power in the<br />

guide.<br />

8.3.4.3 The standing wavemeter<br />

The standing wavemeter is a piece <strong>of</strong> equipment which measures the voltage <strong>of</strong> the standing<br />

wave at different points along the cavity. You can identify it as a rectangular tube with a<br />

BNC cable socket, which can slide up and down the tube, with a ruler attached. To learn<br />

more about the technical detail <strong>of</strong> how this works, see the appendix (section 8.4.3).


8.3. MICROWAVES 105<br />

8.3.5 Experiment<br />

8.3.5.1 Familiarisation<br />

Figure 8.13: The microwave bench<br />

A block diagram <strong>of</strong> the microwave bench is shown in figure 8.13. Each square in the diagram<br />

corresponds with one piece <strong>of</strong> equipment on the bench. The separations in the diagram<br />

correspond to straight pieces <strong>of</strong> hollow waveguide, or in the case <strong>of</strong> the separation between<br />

standing wave meter and twist, or twist and variable end, simply two rings <strong>of</strong> metal screwed<br />

together.<br />

• Ensure the power to the Gunn diode is <strong>of</strong>f, and carefully unscrew a few pieces <strong>of</strong><br />

equipment, and examine inside. Inside the cavity wavemeter, can you see the cross as<br />

in figure 8.11? Carefully reassemble the equipment.<br />

• Connect the DC voltmeter to the standing wave detector. Adjust the cavity wavemeter<br />

until a resonance dip occurs (sudden rapid decrease from reading <strong>of</strong>f the voltmeter).<br />

This corresponds to a reduction in output <strong>of</strong> ∼40% from the standing wavememeter.<br />

(See the appendix (section 8.4.4) for how to read the micrometer).<br />

Record λ g and calculate the frequency <strong>of</strong> the oscillation 5 .<br />

Question 8.5 Is this indeed in the microwave range <strong>of</strong> the electromagnetic spectrum (∼1-<br />

300 GHz)?<br />

8.3.6 Polarisation <strong>of</strong> microwaves in the guide<br />

• Attach a microwave horn to the free end <strong>of</strong> the cavity. Horns simply direct the microwaves<br />

to a narrow beam. Position the other horn with the crystal detector attachment<br />

some distance in front <strong>of</strong> the first horn. Connect the detector to a multimeter.<br />

• Place the wire grille between the two horns and observe the response in the crystal<br />

detector for grille orientations both parallel and perpendicular to the bench surface.<br />

5 See Ohanian [7], chapter 37, page 911.


106 CHAPTER 8. WAVES IN WAVEGUIDES<br />

Question 8.6 What conclusions may you draw from this exercise? (In what direction do the<br />

electric field vectors point in this system?)<br />

8.3.7 Relation between guide and free space wavelengths<br />

Retaining the set-up from the previous exercise replace the other horn with an aluminium<br />

plate ∼20 cm in front <strong>of</strong> the first horn. Attach a multimeter to the standing wave detector<br />

and observe the standing wave pattern by moving the carriage along the slot in the top <strong>of</strong> the<br />

waveguide. Refer to figure 8.14 to see how a standing wave is produced.<br />

Figure 8.14: Standing waves produced using reflection from an aluminium plate<br />

The radiation leaves the main guide via the horn and is reflected back from the plate producing<br />

a standing wave with wavelength λ a between the plate and the horn. Reflected radiation<br />

also enters the guide, producing a standing wave in the guide with wavelength λ g . Moving<br />

the plate simply translates the standing wave pattern laterally by the same amount.<br />

Question 8.7 Determine the guide wavelength λ g by measuring the distance between successive<br />

minima. Now measure λ a , the free space wavelength by keeping the standing wave<br />

detector stationary and moving the aluminium plate. Develop and describe your technique<br />

for minimising error.<br />

Question 8.8 Predict the value <strong>of</strong> λ c using the relationship between λ a and λ g . Compare this<br />

to the theoretical value <strong>of</strong> λ c calculated assuming that the TE 10 mode is being propagated.<br />

Describe the physics <strong>of</strong> what λ c represents.<br />

8.3.8 Impedance measurement<br />

Since microwaves are a wave phenomenon we may use an identical formalism as that described<br />

for acoustics in order to represent them. Thus we may assume identical expressions


8.4. APPENDICES 107<br />

for the phase change, φ = 2kx 0 , the reflection coefficient, r, and the impedance <strong>of</strong> a discontinuity,<br />

Z T /Z 0 , to characterise termination in this case.<br />

• Terminate the line with a short circuit, setting its micrometer to zero.<br />

Question 8.9 Plot the standing wave pattern for one wavelength, that is two maxima and<br />

minima, using the standing wave detector.<br />

Question 8.10 You should record the distance from the flange to the first minimum in order<br />

to calculate the true position <strong>of</strong> the first minimum. Calculate r, φ and K for the short circuit.<br />

Question 8.11 Repeat the measurement for the matched load and open guide.<br />

The short circuit may be described as the perfect mismatch, with almost all the microwave<br />

energy being reflected from the termination. The matched load is the perfect match, containing<br />

a dissipative wedge <strong>of</strong> iron that absorbs most <strong>of</strong> the microwave energy.<br />

Question 8.12 Do your results reflect these differences?<br />

Question 8.13 For electromagnetic waves, there is a π phase shift upon reflection from a<br />

short circuit. That is, φ EM = 2kx 0 − π. Given this, compare your results from this section<br />

with those from the acoustics section.<br />

8.4 Appendices<br />

8.4.1 The complex reflection coefficient K<br />

In the discussion above the complex reflection coefficient was assumed to apply to the displacement<br />

only. This quantity should be more properly labelled K DISP , since reflection is a<br />

process undergone also by the particle velocity and the pressure. From plane wave theory<br />

(and recalling figure 8.1) we may write<br />

∣ ∣ ∣ r ≡<br />

ξ −∣∣∣<br />

∣ =<br />

v −∣∣∣<br />

ξ +<br />

∣ =<br />

p −∣∣∣<br />

v +<br />

∣<br />

(8.50)<br />

p +<br />

Where ξ ± , v ± and p ± are all complex. Let us define the full three complex reflection coefficients.<br />

K DISP = ξ −<br />

ξ +<br />

,<br />

K VEL = v −<br />

v +<br />

,<br />

K PRESS = p −<br />

p +<br />

. (8.51)<br />

We will now show that phase change in acoustic pressure upon reflection is π different to<br />

phase change in both displacement and velocity.


108 CHAPTER 8. WAVES IN WAVEGUIDES<br />

From table 8.2, we have<br />

therefore we have both<br />

v = iωξ (8.52)<br />

and<br />

v − = iωξ − (8.53)<br />

Dividing equation 8.53 by 8.54 we get<br />

v + = iωξ + (8.54)<br />

and since<br />

v −<br />

v +<br />

= ξ −<br />

ξ +<br />

⇒ K VEL = K DISP (8.55)<br />

then from equations 8.14 and 8.13<br />

v − = K VEL v + = K DISP v + ⇒ Z 0 v − = K DISP Z 0 v + (8.56)<br />

−p − = K DISP p + ⇒ K DISP = K VEL = −K PRESS (8.57)<br />

Q.E.D.<br />

Let us now define the form <strong>of</strong> φ = 2kx 0 and K = re iφ that is used in this experiment.<br />

We have an incident plane wave (equation 8.15)<br />

ξ + = a + e i(ωt−kx 0)<br />

(8.58)<br />

and another reflected plane wave (equation 8.16)<br />

ξ − = a − e i(ωt+kx 0)<br />

(8.59)<br />

since r = a −<br />

a +<br />

(equation 8.18)<br />

ξ − = ra + e i(ωt+kx 0)<br />

(8.60)<br />

Now, the reflected wave is given by K (actually K DISP ) times the original wave, so<br />

ξ − = Kξ + = Ka + e i(ωt−kx 0)<br />

(8.61)


8.4. APPENDICES 109<br />

so<br />

and so<br />

K = ra +e i(ωt+kx 0)<br />

a + e i(ωt−kx 0)<br />

(8.62)<br />

where<br />

K = re iφ (8.63)<br />

Q.E.D.<br />

φ = 2kx 0 (8.64)<br />

8.4.2 The Gunn diode<br />

Figure 8.15: Rectification by the crystal detector<br />

You will be using a solid state source to produce electromagnetic radiation <strong>of</strong> microwave<br />

frequencies. The central component <strong>of</strong> a Gunn diode is a piece <strong>of</strong> gallium arsenide crystal.<br />

Generally, in crystals, the repetition <strong>of</strong> a regular array in space results in allowable energy<br />

states for the electrons, depending on their wave number, k. In the absence <strong>of</strong> an applied<br />

electric field, the electrons will choose energy levels as low as possible given a certain temperature.<br />

The net electron flow will be zero since there is no preferred direction. If an<br />

external voltage is applied, the electron velocity will increase linearly with the voltage until<br />

electron collisions with the lattice are energetic enough to damage the crystal.<br />

In GaAs, the linear increase <strong>of</strong> electron velocity with voltage obtains until an internal field<br />

<strong>of</strong> 3 kV/cm is achieved. At this point the electrons gain enough energy to occupy a second<br />

energy level. In this state, the electrons have a higher effective mass and a lower velocity.<br />

Since the velocity is lower, the current is appreciably smaller. Thus even though the voltage is<br />

rising the current is decreasing, meaning that the crystal is behaving as a negative resistance.<br />

Oscillations result.


110 CHAPTER 8. WAVES IN WAVEGUIDES<br />

The oscillation mechanism is complex and involves many modes. The usual mode involves<br />

a dipole distribution <strong>of</strong> charge on either side <strong>of</strong> a non-uniformity in the crystal. This dipole<br />

domain drifts toward the anode and dissipates. The field across the sample rises, causing<br />

a new dipole domain to grow. The frequency <strong>of</strong> this oscillation depends critically on the<br />

domain drift velocity (or applied voltage) and the lateral dimension <strong>of</strong> the sample. These<br />

parameters may be tuned to achieve electromagnetic disturbances <strong>of</strong> a microwave frequency.<br />

8.4.3 The standing wavemeter<br />

The standing wavemeter consists <strong>of</strong> a probe mounted in a slot through the waveguide wall.<br />

This probe is connected to a silicon crystal slab connected directly to a thin tungsten wire.<br />

Passing microwave energy induces a current in the probe. However since the microwave<br />

disturbance is symmetrical about zero, any meter detection <strong>of</strong> this current would read zero<br />

since the fluctuations are so frequent. The crystal-wire configuration rectifies this signal.<br />

Electron flow through the silicon crystal is hampered by the scarcity <strong>of</strong> electrons. Tungsten<br />

as a metal, has surplus electrons, so current can flow readily in this direction only, which<br />

rectifies the signal as shown. The average signal is a measure <strong>of</strong> the microwave energy in the<br />

cavity.<br />

We read <strong>of</strong>f the RMS signal due to the frequency <strong>of</strong> the fluctuations.<br />

8.4.4 Reading the micrometer<br />

This is a foolpro<strong>of</strong> plan to read the micrometer on the cavity wavemeter, the final answer <strong>of</strong><br />

which we shall call M. You will have to adjust the method accordingly for when you can<br />

only just see a line under the thimble - but reading through this guide (including the example<br />

at the end) should give you a good working understanding <strong>of</strong> reading (a) this micrometer in<br />

particular, and (b) a good idea at how to read any other micrometer you may come across.<br />

First <strong>of</strong> all, rotate the micrometer until it is at such a point that you would like to record the<br />

result.<br />

big black line<br />

sleeve<br />

thimble<br />

Figure 8.16: Parts <strong>of</strong> a micrometer


BIBLIOGRAPHY 111<br />

Step 1 What is the biggest red number that you can read, (it is written on the sleeve)? Call<br />

it N. This is the number <strong>of</strong> centimetres. Note that N < M.<br />

Step 2 How many small division markers on the right hand side <strong>of</strong> the big black line can<br />

you count that come after the big division marker corresponding to your N?<br />

If there are 0 then N < M < N + 0.2<br />

If there is 1 then N + 0.2 < M < N + 0.4<br />

If there are 2 then N + 0.4 < M < N + 0.6<br />

If there are 3 then N + 0.6 < M < N + 0.8<br />

If there are 4 then N + 0.8 < M < N + 1.0<br />

Record L, such that (N + L) < M < (N + L + 0.2).<br />

Step 3 Is there a small division marker on the left hand side <strong>of</strong> the big black line after the<br />

final marker on the right hand side?<br />

If the answer is yes, (N + L + 0.1) < M < (N + L + 0.2)<br />

If the answer is no, (N + L) < M < (N + L + 0.1)<br />

Record l, such that l < M.<br />

Step 4 Which division on the thimble is coincident with the big black line? Call the red<br />

number associated with this line (which is written on the thimble) R.<br />

Step 5 The final answer, M, is given by<br />

M = l + R<br />

1000 . (8.65)<br />

Hence, we have measured the number <strong>of</strong> centimetres to 3 decimal places.<br />

Example<br />

Step 1 N = 2, and M > 2.<br />

Step 2 There are 2 ⇒ L = 0.4, and M is between 2.4 and 2.6.<br />

Step 3 Yes ⇒ l = 25, and M is between 2.5 and 2.6.<br />

Step 4 R = 54<br />

Step 5 M = 2.5 + 54<br />

1000<br />

= 2.5 + 0.054 = 2.554 cm<br />

<strong>Bibliography</strong><br />

[1] A W Cross. Experimental Microwaves. Marconi Instruments Ltd., Sanders Division,<br />

1977.


112 CHAPTER 8. WAVES IN WAVEGUIDES<br />

02<br />

25<br />

50<br />

30<br />

60<br />

Figure 8.17: Set the micrometer up so it looks like this.<br />

[2] R Bowers. A solid-state source <strong>of</strong> microwaves. Scientific American, 215, 1966.<br />

[3] P Lorrain and D R Corson. Electromagnetic Fields and Waves. W.H. Freeman and Co.,<br />

2 nd edition, 1970.<br />

[4] R P Feynman, R B Leighton, and M Sands. The Feynman Lectures, volume II. Addison-<br />

Wesley, 1964.<br />

[5] E Hecht. Optics. Addison-Wesley, 2 nd edition, 1987.<br />

[6] D K Cheng. Field and Wave Electromagnetics. Addison-Wesley, 2 nd edition, 1992.<br />

[7] H C Ohanian. <strong>Physics</strong>. 2 nd edition, 1989.


Chapter 9<br />

Electron Spin Resonance<br />

9.1 A guide to background reading<br />

While these notes are designed to be fairly self-contained, students are directed to the cited<br />

references for a more detailed treatment <strong>of</strong> the theory as some results are presented without<br />

detailed pro<strong>of</strong>s. In particular a cursory reading <strong>of</strong> the pertinent sections <strong>of</strong> “Introduction to<br />

Solid State <strong>Physics</strong>” by C. Kittel [6], which is available from the Part 2 <strong>of</strong>fice, may provide<br />

some valuable insights into the phenomena <strong>of</strong> electron diffraction.<br />

Electron spin resonance and Zeeman splitting can be found in any number <strong>of</strong> quantum mechanics<br />

books. Specifically “Modern <strong>Physics</strong>” by Serway, Moses and Moyer [7] or “<strong>Physics</strong><br />

<strong>of</strong> Atoms and Molecules” by B. Bransden and C. Joachain [2] provides detailed treatment<br />

for 2 nd or 3 rd year level.<br />

For an introduction to the quantum mechanics <strong>of</strong> angular momentum and spin, see Quantum<br />

<strong>Physics</strong> <strong>of</strong> Atoms, Molecules, Solids, Nuclei and Particles by R. Eisberg and R. Resnick<br />

[1] or any numer <strong>of</strong> second year textbooks. A discussion on Helmholtz coils can be found<br />

in Experimental Methods in Magnetism by H. Zijlstra [3]. For a detailed discussion <strong>of</strong> the<br />

experimental considerations <strong>of</strong> ESR see Electron Paramagnetic Resonance - Techniques and<br />

Applications by R.S. Alger [4]. For additional information on LC resonator circuits see<br />

Introductory Electronics for Scientists and Engineers by R.E. Simpson, section 2-12 [5].<br />

9.2 Introduction<br />

In 1896, Zeeman observed that atomic spectral lines were split when the atom giving rise<br />

to the spectrum was placed into an external magnetic field. And in 1922, Stern and Gerlach<br />

performed their canonical experiment <strong>of</strong> passing silver atoms through a magnetic field, observing<br />

the original beam split into two in the presence <strong>of</strong> the field. The answer to these<br />

two disparate observations came in 1925, when Uehlenbeck and Goudsmit postulated that<br />

the splitting <strong>of</strong> atomic spectra was due to an intrinsic angular momentum they denoted spin.<br />

This property couples to the orbital angular momentum <strong>of</strong> the electrons giving rise to the<br />

observed splitting in the spectral lines in the presence <strong>of</strong> an external magnetic field. That<br />

spin-orbit coupling is a fundamental force in atomic and subatomic physics. While such a<br />

113


114 CHAPTER 9. ELECTRON SPIN RESONANCE<br />

feature has been incorporated extensively in the Schrödinger equation to describe phenomena<br />

(nuclear physics couldn’t work without it), a true understanding <strong>of</strong> spin-orbit coupling<br />

came in 1929 with Dirac and his equation. Spin was the first quantum observable introduced<br />

which has no classical analogue. It is developed in this section.<br />

Note, however, that Dirac actually sought, as did Schrödinger before him, a correct wave<br />

theory for the electron which incorporated the spin <strong>of</strong> the electron. Spin itself is not a consequence<br />

<strong>of</strong> the Dirac equation; the spin-orbit coupling is.<br />

In this experiment we will study the Zeeman splitting <strong>of</strong> spectra from a molecule, diphenylpicra-hydrazyl<br />

(DPPH), which has an unpaired electron on one <strong>of</strong> the nitrogen atoms. It has<br />

features which allow for the spin <strong>of</strong> the electron to be studied in isolation.<br />

9.3 Introduction: Angular Momentum in Quantum Mechanics<br />

9.3.1 Orbital Angular Momentum<br />

The quantum mechanical analogue <strong>of</strong> classical angular momentum is orbital angular momentum.<br />

For a particle moving in a circular path about a fixed point in space, its angular<br />

momentum is defined as in the classical case, viz.<br />

L = R × P . (9.1)<br />

The quantum mechanical properties <strong>of</strong> orbital angular momentum are, however, different.<br />

The properties <strong>of</strong> momentum, and how they relate to the position <strong>of</strong> a particle, require the<br />

orbital angular momentum take quantized values, which are zero and positive integers, ie<br />

L = 0, 1, 2,....<br />

As angular momentum is a vector, we may define its projection against an arbitrary axis,<br />

with a given direction. In doing so, we find that the values its projection may take are also<br />

integer, but this time including negative values. Picture the positive values as corresponding<br />

to the angular momentum vector pointing in the same general direction as the axis, and the<br />

negative values with the angular momentum pointing in the general opposite one.<br />

9.3.2 Spin<br />

The above picture, however, is incomplete. We may observe the above projections only in<br />

the presence <strong>of</strong> a magnetic field, which splits an energy level corresponding to a given L into<br />

its 2L + 1 levels, from −L to +L in integer steps. For a particle moving in a magnetic field,<br />

such splitting is due to an induced magnetic moment, viz.<br />

µ = e ↕. (9.2)<br />

2mc<br />

For an atom, comprising <strong>of</strong> (an even number <strong>of</strong>) Z electrons, the total magnetic moment is<br />

the sum <strong>of</strong> each magnetic moment induced by each orbiting electron. This amounts to a split


9.4. ELECTRONS IN AN EXTERNAL MAGNETIC FIELD 115<br />

into 2L + 1 levels dictated by<br />

E = MBµ B , (9.3)<br />

where B is the strength <strong>of</strong> the external field and µ B is a constant <strong>of</strong> proportionality known as<br />

the Bohr magneton:<br />

µ B = e<br />

2mc . (9.4)<br />

The M values are known as the magnetic, or projection, quantum numbers and may take<br />

values from −L to +L in integer steps. Thus, for even-Z atoms, there are an odd number <strong>of</strong><br />

split states.<br />

A major problem arises in the case <strong>of</strong> atoms with an odd number <strong>of</strong> electrons: the number<br />

<strong>of</strong> split levels is observed to be even. That leads to the contradictory conclusion that L must<br />

be half -integer. This was observed most strikingly in 1922 by Stern and Gerlach. They<br />

passed a beam <strong>of</strong> silver atoms through a magnetic field and observed that the beam split<br />

into 2. (Remember that silver has Z = 47 which means that there is one odd electron in its<br />

configuration.) If there are 2L+1 = 2 levels into which the angular momentum may project,<br />

then that would mean an L value that is half -integer, specifically L = 1 2 .<br />

The solution came in 1925 when Uehlenbeck and Goudsmit postulated that the electron must<br />

contain an intrinsic angular momentum, which they denoted by spin, taking the value 1 2 for<br />

the electron. That then induces an additional magnetic moment in the presence <strong>of</strong> a magnetic<br />

field given by<br />

µ s = g s<br />

e<br />

2mc s (9.5)<br />

where s is the spin <strong>of</strong> the electron. Agreement with experiment is achieved when g s = 2.<br />

The total angular momentum j is then taken as the vector sum <strong>of</strong> the orbital and spin angular<br />

momenta, and this resolves any discrepancy with experiment.<br />

As with orbital angular momentum, one may then project the spin vector onto an arbitrary<br />

axis. What one finds is that the spin magnetic values may take values anywhere from −s to<br />

+s in integer steps, as with L. For the electron, s = 1 which means it has projection values<br />

2<br />

<strong>of</strong> m s = ± 1 ; the positive we denote as “spin-up” and the negative “spin-down”, refering to<br />

2<br />

the direction the vector takes against the axis. That we have two such values explains the<br />

splitting into two <strong>of</strong> the silver atoms in the Stern-Gerlach experiment.<br />

Two important facets <strong>of</strong> spin should be noted:<br />

1. It is a fundamental property <strong>of</strong> particles;<br />

2. There is no classical analogy for it.<br />

In fact spin was the first observable in the atomic regime for which there was no classical<br />

counterpart.<br />

9.4 Electrons in an external magnetic field<br />

What happens when an electron is placed in a uniform magnetic field? If we passed a beam<br />

<strong>of</strong> electrons through a uniform field we would see the beam split into two consistent with


116 CHAPTER 9. ELECTRON SPIN RESONANCE<br />

E<br />

E 0<br />

0<br />

B=0 B=<br />

E−∆Ε<br />

Ε+ ∆Ε<br />

0<br />

−∆Ε<br />

+∆Ε<br />

Figure 9.1: Energy splitting for an electron in a uniform magnetic field B, with direction as<br />

indicated. Note that the value <strong>of</strong> ∆E is, from Eq. (9.6) negative.<br />

the original Stern-Gerlach experiment. A stationary electron in a uniform field would feel<br />

the magnetic field through the spin magnetic moment and precess about the direction <strong>of</strong> the<br />

field. This will induce a change in the energy <strong>of</strong> the electron, depending on the orientation<br />

<strong>of</strong> the electron with the field, viz.<br />

∆E = −µ s · B. (9.6)<br />

In this situation, therefore, the external field itself defines the axis against which the spin is<br />

measured. To minimise energy, the electron would wish to align its spin with the field. If it<br />

is unable to do so, precession occurs. If no precession is observed, then the electron’s spin<br />

will either be parallel (m s = + 1) or anti-parallel (m 2 s = − 1 ) to the field.<br />

2<br />

Prelab Question (A): Using Eq. (9.6) show that when an electron is placed in an external<br />

field, its energy changes by<br />

∆E = ± 1 2 g sµ B B . (9.7)<br />

The change in energy depends on whether the electron is in the spin-up or spin-down state.<br />

The minus sign in Eq. (9.6) means that the spin-up electron will be in a lower energy state<br />

than that in the spin-down. (Why?). But that change in energy is only effected when there<br />

is an external magnetic field, ie., for a stationary electron the only mechanism by which it<br />

interacts with the external magnetic field is through its spin.<br />

Once the field is applied the electron will either be aligned with or against the field (spin-up<br />

or down). To make the electron switch between these states requires the additional energy<br />

<strong>of</strong> ∆E ′ = g s µ B B. That may be done by the absorption or emission <strong>of</strong> a photon with this<br />

energy. By measuring that energy we may then determine the value <strong>of</strong> g s from experiment.<br />

The whole scenario is illustrated in Fig. 9.1.<br />

9.5 Resonance absorption<br />

We consider the situation in which the only source <strong>of</strong> photons in the system are those that are<br />

controlled in the experiment. The experiment should account for background lighting and


9.6. EXPERIMENT 117<br />

N<br />

N<br />

Figure 9.2: The DPPH molecule showing the isolated unpaired electron in the molecular<br />

configuration.<br />

control the wavelength <strong>of</strong> the photons required in the absorption in the above situation to be<br />

outside that <strong>of</strong> background light.<br />

Prelab Question (B): Calculate the value <strong>of</strong> the external magentic field necessary such that<br />

a photon, wavelength λ = 450 nm, has the required energy to flip the spin <strong>of</strong> the electron.<br />

Why then may we ignore the background light when performing the experiment?<br />

The source <strong>of</strong> electrons in this experiment is the organic molecule diphenyl-picra-hydrazyl,<br />

or DPPH (Fig. 9.2). This molecule is convenient in that it has one valence, unbonded electron<br />

on the second N atom. The interaction <strong>of</strong> that electron with the mean Coulomb field<br />

generated by the other electrons in the molecule ascribe an energy E 0 to it.<br />

Prelab Question (C): Why can’t truly free electrons be used in this experiment? Why is a<br />

beam <strong>of</strong> electrons or a metal inappropriate?<br />

Choosing a sample in which an electron may be unpaired is not the only consideration. Another<br />

may be the situation in which the electron in its natural state in the molecular configuration<br />

is predominantly in the spin-down level. One requires such in order to see a measureable<br />

effect. One would also like the lifetime <strong>of</strong> the spin-down state to be relatively short in order<br />

to eventually decay to the lower energy spin-up state. (This is an important point.)<br />

Under these conditions, we may then provide a source <strong>of</strong> photons, frequency f, in an experiment.<br />

When the photon is at the resonance frequency for a given magnetic field strength,<br />

then the spin-up electron will flip its spin; the resonance energy is given by hf = g s µ B B.<br />

By scanning through a range <strong>of</strong> frequencies, one may then observe that resonance.<br />

Note that in this experiment the electron does not exhibit any orbital motion, hence the<br />

response to the magnetic field <strong>of</strong> the electron will be solely due to the electron’s spin.<br />

9.6 Experiment<br />

Fig. 9.3 shows the experimental configuration required to measure g s .<br />

9.6.1 Helmholtz coils<br />

NB! There are LARGE coils which carry current on the desk.


118 CHAPTER 9. ELECTRON SPIN RESONANCE<br />

inductor with<br />

sample<br />

Oscillator<br />

power<br />

supply<br />

Coil power<br />

supply<br />

AC<br />

I<br />

B<br />

Helmholtz coils<br />

(in series)<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

RF<br />

−12 0 +12<br />

−12 0 +12<br />

Y<br />

000000000<br />

111111111<br />

000000000<br />

111111111<br />

1 Ω<br />

TO CRO<br />

freq.<br />

meter<br />

Oscillator adaptor<br />

Figure 9.3: The apparatus used for the measurement <strong>of</strong> ESR. The AC supply is 50 Hz, and<br />

the voltages indicated on the Oscillator supply and adaptor are in V. The field generated by<br />

the coils is in the direction as indicated.<br />

The Helmholtz coils are connected to an AC power supply (50 Hz), so the current will vary<br />

sinusoidally with time.<br />

Prelab Question (D): What do Helmholtz coils produce? Discuss with your demonstrator,<br />

if unsure.<br />

Prelab Question (E): Let R be the radius <strong>of</strong> a pair <strong>of</strong> Helmholtz coils separated by a distance<br />

R. If x denotes the distance from the centre <strong>of</strong> the left hand coil to any point<br />

along that axis, calculate the magnetic field produced at points x = 0.2R and x =<br />

0.5R. Remember that the field produced by one ccoil with n turns carrying a current<br />

I is given by<br />

µ 0 nIR 2<br />

B =<br />

(9.8)<br />

2 (x 2 + R 2 ) 3/2<br />

Question (a): Draw a voltage vs time graph for the voltage across the resistor for two full<br />

periods <strong>of</strong> the AC signal. Assuming a peak-to-peak voltage <strong>of</strong> 5 V across the coils,<br />

draw the B vs time graph for the coils. Why do we use a 1 Ω resistor?<br />

9.6.2 RF oscillator<br />

The RF oscillator provides the photons needed to examine resonance absorption. It coverts a<br />

signal into a magnetic field and back again. The field produced preiodically bathes anything<br />

within the coil in a sea <strong>of</strong> photons <strong>of</strong> the frequency selected using the knob on the unit.


9.6. EXPERIMENT 119<br />

The oscillator can produce photons with frequencies between 30 and 130 MHz depending<br />

on the coil attached. (The smaller the coil, the higher the frequency.) On the rear <strong>of</strong> the<br />

oscillator you can connect a micro-ammeter to the socket marked “I/µA”. The ammeter<br />

then monitors the current flowiing the through the unit and the frequency.<br />

9.6.3 Investigating the resonance<br />

Included is what is termed a “tank” circuit. It consists <strong>of</strong> a variable capacitor connected to a<br />

coil <strong>of</strong> wire similar to the one on the RF unit. The circuit is actually an LC circuit and will<br />

resonate at the frequency determined by the values <strong>of</strong> capacitance and inductance.<br />

The tank circuit is simply used to demonstrate resonance, and the settings at which such<br />

resonance are NOT used for the remaining sections <strong>of</strong> the experiment.<br />

Turn on the RF unit and use the ammeter to determine the oscillation frequency. Then bring<br />

the tank circuit up to the RF unit such that the coils are ALMOST touching (you will have<br />

to take the RF unit out <strong>of</strong> the Helmholtz coils to do this). Connect the tank circuit up to the<br />

CRO to monitor the voltage across the circuit and then adjust the knob on top fo the trank<br />

circuit until it resonates wth the RF unit. This will occur when the voltage across the CRO<br />

reaches a maximum.<br />

Note: the LC tank circuit has a limited range corresponding to the range <strong>of</strong> the variable<br />

capacitor. If you don’t observe a maximum you may have to change the frequency <strong>of</strong> the RF<br />

unit.<br />

When you have determined the resonance point, use the ammeter to examine the current<br />

through the RF unit. What happens as you move in and out <strong>of</strong> the resonance? Explain.<br />

9.6.4 Electron absorption<br />

The sample <strong>of</strong> DPPH is contained in a vial. Place the sample within the coil <strong>of</strong> the RF unit,<br />

itself in the centre <strong>of</strong> the Helmholtz coils.<br />

Draw the B vs time graph through the coils that indicates the strength <strong>of</strong> the uniform field<br />

seen by the electrons in the DPPH sample. Under this plot draw the current you would expect<br />

to measure through the RF oscillator. Assume that the maximum value for B is 0.50 T and at<br />

a give frequency the resonance condition is satisfied whenever B = ±.025 T. Discuss with<br />

your demonstrator.<br />

Question (b): The relaxation time <strong>of</strong> the electrons back to the ground state should be short,<br />

compared to the frequency <strong>of</strong> the sweeping B field. Why is it so?<br />

Position the Helmholtz coils correctly using the dial caliper ensuring that they are connected<br />

corretly and in series with the resistor and AC supply. “A” identifies the beginning <strong>of</strong> the<br />

coild and Z the end. The mean diameter <strong>of</strong> the coils is 13.6 cm and the number <strong>of</strong> turns in<br />

each is 320.


120 CHAPTER 9. ELECTRON SPIN RESONANCE<br />

Connect the RF unit to the power supply, place the sample in the center <strong>of</strong> the coil. Position<br />

the RF coil as centrally as possible into the Helmholtz coils.<br />

After everything is connected examine the voltage across the resistor and the current through<br />

the RF unit simultaneously. Adjust the current through the coils. What happens to the current<br />

through the RF unit?<br />

Question (c): Why do the features on the plot change? Discuss with your demonstrator.<br />

You should now be able to determine how best to measure B when resonance occurs.<br />

Question (d): How should the maximum current through the coils be chosen such that you<br />

reduce the error when measuring B at resonance?<br />

Measure the voltage at which resonance occurs, as a function <strong>of</strong> RF frequency. Use the<br />

small bar magnet to observe any changes when you change the magentic field by placing<br />

the magnet near the sample such that it is parallel and then perpendicular to the coil axis.<br />

Explain.<br />

Question (e): You should have enough information to determine g s . How does it compare<br />

to the nominal g s = 2?<br />

Question (f): Why to we observe a width on the resonance peak?<br />

9.7 Extension: electron diffraction<br />

This section contains some extra work for efficient or ambitious students. It should only be<br />

attempted when all the work in the previous section has been completed. You can still get<br />

full marks having done only the previous section.<br />

9.7.1 Overview<br />

The postulate that matter may also possess wave-like properties was one <strong>of</strong> the most stunning<br />

and unifying hypotheses <strong>of</strong> this century. In 1924 de Broglie, working from arguments based<br />

on the properties <strong>of</strong> wave packets, hazarded that the wavelength <strong>of</strong> these matter waves could<br />

be obtained from the same relationship that held for light, namely<br />

λ = h p<br />

(9.9)<br />

where λ is the wavelength, p is the momentum and h is Planck’s constant 1 .<br />

Davisson and Germer obtained the first experimental evidence for matter waves by reflecting<br />

slow electrons from the face <strong>of</strong> a single nickel crystal. The wavelength that they were able to<br />

1 Some useful data, including a value for h, are provided in section 9.8.


9.7. EXTENSION: ELECTRON DIFFRACTION 121<br />

determine from using the periodic array <strong>of</strong> nickel atoms as a diffraction grating was strikingly<br />

similar to that calculated from the de Broglie relation. Further experimentation by Thomson<br />

yielded diffraction patterns from the transmission <strong>of</strong> electrons through thin polycrystalline<br />

foils. The fact that non-relativistic electron wavelengths are comparable to interatomic distances<br />

in solids means that electron diffraction experiments (such as Transmission Electron<br />

Microscopy, or TEM) are powerful probes <strong>of</strong> the structure <strong>of</strong> solids.<br />

9.7.2 Basic theory<br />

9.7.2.1 Wavelength<br />

The wavelength <strong>of</strong> a beam <strong>of</strong> electrons is given by<br />

λ = h p = h mv . (9.10)<br />

Thus in the non-relativistic limit the wavelength is inversely proportional to the electron<br />

velocity v. Conservation <strong>of</strong> energy requires that for a charge accelerated through a potential<br />

the change in kinetic energy plus the change in electrical potential energy while traversing<br />

from one point to another must equal zero. Thus we may write<br />

( ) mv<br />

2<br />

2<br />

2 + mv2 1<br />

+ (eV 2 − eV 1 ) = 0 (9.11)<br />

2<br />

where V 1 and V 2 are the potentials at points 1 and 2 respectively. Applied to the electron<br />

transversing an electron gun in an electron diffraction tube V 2 is V a , the anode voltage and<br />

V 1 equals zero. So for negative charges<br />

mv 2<br />

2<br />

= eV a . (9.12)<br />

Hence the wavelength is given by<br />

λ =<br />

h<br />

√ 2meVa<br />

, (9.13)<br />

where m and e are the mass and charge <strong>of</strong> the electron respectively. When the values <strong>of</strong> the<br />

constants are substituted into equation 9.13 and the potential is in volts then the wavelength<br />

in nanometres is given by<br />

λ = 1.2264 √<br />

Va<br />

. (9.14)


122 CHAPTER 9. ELECTRON SPIN RESONANCE<br />

9.7.2.2 Electron diffraction<br />

Pre-lab Question 9.1 Given the situation <strong>of</strong> electrons being reflected from a set <strong>of</strong> planes<br />

and the geometry shown below (figure 9.4) derive the condition for constructive interference,<br />

known as the Bragg condition,<br />

2d sin θ b = nλ (9.15)<br />

Figure 9.4: The reflection <strong>of</strong> light from a set <strong>of</strong> planes. The Bragg condition for constructive<br />

interference may be derived purely from the geometry shown.<br />

The experiment you will undertake replicates Thomson’s method <strong>of</strong> transmitting electrons<br />

through a thin film <strong>of</strong> randomly oriented crystallites to investigate a ring diffraction pattern.<br />

The periodic array <strong>of</strong> atoms in the 3-dimensional solid forms a space lattice. Diffraction from<br />

such a space lattice is somewhat more involved than the situation just depicted, however, a<br />

close analogue <strong>of</strong> the Bragg equation still holds.<br />

2d n sin θ b = λ (9.16)<br />

Here d n is the distinct interplanar distance for a particular set <strong>of</strong> planes described by the<br />

periodic atomic arrangement. A diffraction pattern for a single crystal with a cubic structure<br />

is shown below (figure 9.5). R 1 and R 2 arise from two different sets <strong>of</strong> planes, and when<br />

analysed, will yield two separate d values, d 1 and d 2 .<br />

The angle between the direct and diffracted beam is shown in figure 9.6. You will be measuring<br />

ring radius as a function <strong>of</strong> anode voltage.<br />

Pre-lab Question 9.2 Using the geometry shown in figure 9.6, derive an expression relating<br />

the ring radius to the scattering angle, 2θ b . You will be measuring D = 2R with a pair <strong>of</strong><br />

dial calipers, so take care when you define this quantity.<br />

Pre-lab Question 9.3 Using the Bragg relation and the expression determined above, derive<br />

an expression for d n , the interatomic spacing in terms <strong>of</strong> R.


9.7. EXTENSION: ELECTRON DIFFRACTION 123<br />

Figure 9.5: The electron diffraction pattern from a single crystal <strong>of</strong> a substance with a cubic<br />

structure. The radii shown correspond in real space to two separate sets <strong>of</strong> crystal planes and<br />

yield two values for the interplanar spacing, d.<br />

Pre-lab Question 9.4 The specimen you will use in this experiment is graphitised carbon.<br />

Assume that the form <strong>of</strong> carbon is a simple cubic structure. The atomic mass <strong>of</strong> carbon is 12<br />

and its density is 2.3 g cm −3 . Calculate the spacing <strong>of</strong> adjacent carbon atoms.<br />

Figure 9.6: The geometry <strong>of</strong> the electron diffraction experiment.<br />

9.7.3 Experiment<br />

HIGH VOLTAGES ARE USED IN THIS EXPERIMENT. BEFORE SWITCHING<br />

THE APPARATUS ON CHECK THAT ALL THE VOLTAGES ARE TURNED<br />

DOWN. IF THIS IS UNCLEAR CONSULT YOUR DEMONSTRATOR.


124 CHAPTER 9. ELECTRON SPIN RESONANCE<br />

• Ensure that the apparatus is wired up as shown in figure 9.7.<br />

• Switch on the high voltage supply and wait for one minute for the cathode heater to<br />

stabilise.<br />

• Select “CT”, and slowly increase the high voltage, V a , to the approximate value <strong>of</strong><br />

2.7 kV. This is the knob which you will be adjusting for this experiment. Remember,<br />

this is the anode voltage, and corresponds to equation 9.14.<br />

• You must ensure that the current on the multimeter does not exceed 0.2 mA, and you<br />

can modify this by adjusting V b , the bias voltage supply. Current overload causes<br />

the target to glow a dull red, so inspect the target at intervals during the experiment<br />

to ensure that this is not happening. Adjust the bias voltage supply V b to produce a<br />

clear ring image and uniform central spot. If V b is initially connected the wrong way<br />

reverse it, but TURN DOWN THE HIGH VOLTAGE BEFORE DOING THIS. Note<br />

that adjusting the knob for V b also changes the value <strong>of</strong> V a , so ensure that you are<br />

measuring the ring radius for the correct value <strong>of</strong> V a !<br />

• Sketch the observed ring pattern. Roam around the neck <strong>of</strong> the tube with a bar magnet<br />

and try to increase the clarity <strong>of</strong> the ring pattern. If there is no improvement, remove<br />

the magnet. Sketch the ring pattern with the magnet in different positions. Describe<br />

and explain what you observe.<br />

Question 9.1 Deduce which end <strong>of</strong> the magnet is North.<br />

• Using the dial calipers measure both ring diameters for 12 different anode voltages.<br />

You may note that the rings are distorted. To minimise errors, four measurements<br />

should be taken in the four quadrants and averaged. Watch for drift in the voltage and<br />

adjust when required. Record all your measurements including the voltages.<br />

• Explain why the ring diameters increase with decreasing voltage.<br />

• Graph Va<br />

−1/2 against sin θ b using Excel, and calculate the interatomic spacings from<br />

the gradient. Compare your lattice spacings to the actual values <strong>of</strong> d 11 = 0.123 nm<br />

and d 10 = 0.213 nm. From your derived values <strong>of</strong> interplanar distance suggest how<br />

the carbon atoms are more likely to be arranged. Use a diagram to illustrate your<br />

conclusions.<br />

9.7.4 Discussion<br />

Question 9.2 Why does diffraction from graphitised carbon produce a ring pattern? What<br />

does a single crystal diffraction pattern look like? Explain the difference.<br />

Question 9.3 Could you perform a similar experiment with protons? What parameter would<br />

you have to change?<br />

Question 9.4 Would equation 9.13 still be valid for a 20 MeV beam <strong>of</strong> electrons? What<br />

expression would be more appropriate?


9.8. USEFUL DATA 125<br />

Figure 9.7: The circuit used in the electron diffraction experiment.<br />

9.8 Useful data<br />

Quantity<br />

Value<br />

h<br />

6.626 × 10 −34 J s<br />

µ b 9.273 × 10 −24 A m 2<br />

µ 0 1.2566 × 10 −6 H m −1<br />

Avogadro’s number 6.02217 × 10 23<br />

Nuclear magneton µ N = 5.05 × 10 −27 J T −1<br />

9.9 Wave function for the electron - a brief history<br />

Once the spin <strong>of</strong> the electron had been discovered, Schrödinger actually sought to write down<br />

a wave equation for the electron incorporating spin and wrote down (first) what we now<br />

know as the Klein-Gordon (KG) equation, which is a relativistic version <strong>of</strong> the Schrödinger<br />

equation. (It is important to remember that Schrödinger first wanted to find a relativistic wave<br />

equation for the electron.) Schrödinger noted that the KG equation could not accommodate<br />

in any reasonable way the spin <strong>of</strong> the electron and remain relativistic. (Thus the Klein-<br />

Gordon equation remains the relativistic wave equation for particles with integer spin, such<br />

as the α particle, which has a spin <strong>of</strong> 0.) Dirac resolved the problem <strong>of</strong> a relativistic wave<br />

equation for the electron in 1929 and, with it, found the source <strong>of</strong> the spin-orbit force, which<br />

was known to affect the electron orbits in the atom.


126 CHAPTER 9. ELECTRON SPIN RESONANCE<br />

<strong>Bibliography</strong><br />

[1] R Eisberg and R Resnick. Quantum <strong>Physics</strong> <strong>of</strong> Atoms, Molecules, Solids, Nuclei and<br />

Particles. Wiley, 2 nd edition, 1985.<br />

[2] B H Bransden and C J Joachain. <strong>Physics</strong> <strong>of</strong> Atoms and Molecules. Longman, 1983.<br />

[3] H Zijlstra. Experimental Methods in Magnetism. North-Holland Pub. Co., 1967.<br />

[4] R S Alger. Electron Paramagnetic Resonance - Techniques and Applications. Interscience<br />

Publishers, 1968.<br />

[5] R E Simpson. Introductory Electronics for Scientists and Engineers. Allyn and Bacon,<br />

1987.<br />

[6] C Kittel. Introduction to Solid State <strong>Physics</strong>. Wiley, 7 th edition, 1996.<br />

[7] R A Serway, C J Moses, and C A Moyer. Modern <strong>Physics</strong>. Saunders, 1989.


Part III<br />

Nuclear physics<br />

127


Chapter 10<br />

Rutherford scattering<br />

10.1 Abstract<br />

One <strong>of</strong> the breakthroughs by Einstein in 1905 dealt with Brownian motion, that seemingly<br />

random motion by small particles on the surface <strong>of</strong> water. He demonstrated that this proved<br />

the existence <strong>of</strong> atoms and molecules; the “random” motion was caused by collisions with<br />

the molecules in the water. Together with the discovery <strong>of</strong> the electron 9 years earlier by<br />

J.J. Thompson, the questions were asked as to the structure <strong>of</strong> the atom itself.<br />

The one that was naturally assumed was the so-called “plum pudding” model, where the<br />

electrons were embedded in positively charged matter. (It had been established by chemists<br />

that atoms had to be electrically neutral.) While this assumption had not been put to the test,<br />

it was already known to have at least one flaw: would not the electrons and positive charges<br />

cancel each other out through interaction? A test came in 1911 by Rutherford and his team,<br />

who conducted an experiment by elastically scattering α particles 1 from a thin gold (Au) foil.<br />

As the α particles are positively charged, it was assumed that the scattering would proceed<br />

via the Coulomb interaction from the positively charged centres in the atom.<br />

There were two surprises: first, some <strong>of</strong> the α particles were elastically scattered at an angle<br />

<strong>of</strong> 180 ◦ . Second, and far more surprisingly, most <strong>of</strong> them did not scatter at all. The only<br />

interpretation, and the correct one, was that the positive charge was concentrated in the very<br />

centre <strong>of</strong> the atom, with the electrons in orbit at a great distance from it, and that most <strong>of</strong> the<br />

atom was, in fact, free space.<br />

This experiment reproduces that first experiment, with the view to measuring the distribution<br />

<strong>of</strong> elastically scattered α particles from which one may infer the existence <strong>of</strong> the nucleus. We<br />

may also, as a result, find the thickness <strong>of</strong> the gold foil we use.<br />

1 Remember that an α particle consists <strong>of</strong> two protons and two neutrons, i.e. an He 2+ ion.<br />

129


130 CHAPTER 10. RUTHERFORD SCATTERING<br />

10.2 Introduction/background reading - importance <strong>of</strong> scattering<br />

“. . . is devoted to the theory <strong>of</strong> scattering and, more generally, collision processes. It is<br />

impossible to overemphasise the importance <strong>of</strong> this subject.” [1], chapter 7.<br />

Imagine yourself looking at the entrance to a tunnel at the side <strong>of</strong> a mountain. It is dark, it<br />

is wet, and you are wondering if there is anything in the tunnel. You can’t see directly into<br />

it because <strong>of</strong> the lack <strong>of</strong> light and you don’t want to go near it as you’ll be soaked. So you<br />

send in a car (no driver, this one’s robotic...). If shortly thereafter you hear a loud crash and<br />

observe a wheel flying out <strong>of</strong> the tunnel back towards you, then you may safely infer that<br />

there was something in the tunnel. So you send in another car. And another. And another<br />

after that. After picking up the pieces <strong>of</strong> debris lying in front <strong>of</strong> you from each collision, a<br />

picture soon emerges <strong>of</strong> what that object may have been. The velocity at which the pieces<br />

came flying back tells you <strong>of</strong> the impulse involved in the collision and hence you may infer<br />

something <strong>of</strong> the object’s mass. The time you record when you hear the crash tells you<br />

something <strong>of</strong> where that object may have been in the tunnel.<br />

This type <strong>of</strong> experiment is typical <strong>of</strong> scattering experiments at the quantum level. We can’t<br />

see directly the object about which we wish to find information, so we send something in to<br />

scatter from it and pick up the pieces. Doing this repeatedly and controlling the kinematic<br />

conditions then allows us to work backwards and infer the structure <strong>of</strong> the object. This has<br />

direct correlation to diffraction studies in optics, where studying diffraction patterns allows<br />

us to work backwards to the structure <strong>of</strong> the diffracting object.<br />

Rutherford, and his students Geiger and Marsden, scattered low-energy α particles from a<br />

gold foil to find out about the structure <strong>of</strong> atoms; in particular, the structure <strong>of</strong> the positively<br />

charged half <strong>of</strong> them. Most <strong>of</strong> nuclear and particle physics has been done as variations on<br />

this theme. The quarks were discovered in 1964 by deep inelastic scattering <strong>of</strong> high energy<br />

electrons <strong>of</strong>f the proton, which translated to elastic scattering <strong>of</strong> electrons <strong>of</strong>f the quarks<br />

inside.<br />

10.3 Geometry<br />

In order to perform this experiment it is first necessary to understand the geometry <strong>of</strong> the<br />

experiment. This then provides the framework by which we may understand the scattering.<br />

Figure 10.1 defines the variables for the scattering, with the relevant lengths for the<br />

arrangement given in table 10.1.<br />

10.3.1 Cross section<br />

The quantity we will measure is the number <strong>of</strong> scattered α particles reaching the detector.<br />

However, that number is affected by variables whose values are unique to this experiment:<br />

Number <strong>of</strong> nuclei The number <strong>of</strong> Au nuclei from which scattering may take place is defined


10.3. GEOMETRY 131<br />

only for this experiment. It depends on the thickness <strong>of</strong> the foil, the area <strong>of</strong> the annulus,<br />

and the density <strong>of</strong> Au.<br />

Flux <strong>of</strong> α particles The number <strong>of</strong> α particles produced by the source must be known. This<br />

depends on the activity <strong>of</strong> the source.<br />

Counting time The number is affected by the time alloted for counting. This has bearing<br />

also on the statistics needed.<br />

θ<br />

Scattering angle<br />

D S−F<br />

αsource<br />

θ R<br />

D S−0<br />

θ D<br />

000 111<br />

000 111<br />

000 111<br />

000 111<br />

detector<br />

Au foil<br />

l = 6.90 cm<br />

d = 1 − 20 cm<br />

Figure 10.1: The geometry <strong>of</strong> the experiment. The various distances are given in the table<br />

10.1.<br />

Apparatus Quantity Value<br />

Au foil Inner radius r 1 2.05 cm<br />

Outer radius r 2 2.50 cm<br />

α source diameter 0.63 cm<br />

detector diameter 0.8 cm<br />

Table 10.1: The values <strong>of</strong> the parameters defined in figure 10.1.<br />

In order to produce a measurable quantity (or semi-measurable) that may be compared to<br />

results <strong>of</strong> other experiments, and also to theory, we need to define a quantity that is more<br />

general, and does not depend on the quantities defined above. Defining such a measurable<br />

quantity also allows us to make the experiment reproducible, even for ourselves. This comes<br />

from the problem that when we perform the experiment the conditions under which the<br />

experiment is performed is unique to that time. Even if the experiment is performed with the<br />

same equipment at a later date, the results may be different as the thickness <strong>of</strong> the foil may<br />

have changed (deformities in the metal), the activity <strong>of</strong> the source may have changed (decay<br />

laws playing their part), or the time taken may be different.


132 CHAPTER 10. RUTHERFORD SCATTERING<br />

The first thing we can do is instead to think <strong>of</strong> the probability per unit time <strong>of</strong> scattering. The<br />

probability is found by dividing the number <strong>of</strong> particles detected by the number given from<br />

the source. Taking it per unit time removes the problem involving the counting time. If we<br />

now normalise it to the probability per unit time per nucleus, then we obtain an observable<br />

which we can compare to results <strong>of</strong> other experiments. That observable is the cross section.<br />

The cross section may be thought <strong>of</strong> as an “area” within the nucleus as the particle sees it. If<br />

the particle hits that area then, on average, it will scatter through the angle θ and reach the<br />

detector. As the name suggests, it has the units <strong>of</strong> area. When a particle passes through a<br />

thin target, the probability for a scattering event is, in terms <strong>of</strong> the size <strong>of</strong> the beam,<br />

blocked area in beam<br />

Probability for scattering =<br />

beam area<br />

− dN (nA dx)σ<br />

=<br />

N A<br />

= nσ dx. (10.1)<br />

The blocked area in beam indicates the number <strong>of</strong> particles that have actually been scattered<br />

away from the beam direction. dN is the number <strong>of</strong> scattered particles, N is the number <strong>of</strong><br />

incident particles, n is the number <strong>of</strong> nuclei per unit volume in the target, dx is the thickness<br />

<strong>of</strong> the target, and A is the area <strong>of</strong> the target struck by the beam. σ is the scattering cross<br />

section, and, from equation 10.1, it is clear that it must have the dimension <strong>of</strong> area. In fact, it<br />

is customary to quote the cross section in units <strong>of</strong> barn (1 barn = 10 −28 m 2 ), which comes<br />

straight from the saying “couldn’t hit the side <strong>of</strong> a ...”<br />

10.3.2 Solid angle<br />

The solid angle comes from considering how position is defined using spherical coordinates.<br />

The effective area <strong>of</strong> the detector viewed from the target is dependent on the distance to the<br />

detector from the target. This is illustrated in figure 10.2. The solid angle, dΩ, subtended at<br />

the surface from the point P may be calculated as<br />

dΩ =<br />

dS · ˆr<br />

r 2 , (10.2)<br />

where the vector dS is the (infinitesimal) area vector normal to the surface. In terms <strong>of</strong> the<br />

angular coordinates (θ,ϕ) <strong>of</strong> a spherical system, the solid angle is<br />

dΩ = sin θ dθdϕ. (10.3)<br />

The units <strong>of</strong> solid angle are steradians (abbreviated “sr”). There are 4π steradians in a sphere.<br />

P<br />

dΩ<br />

r<br />

dS<br />

Figure 10.2: Solid angle.


10.3. GEOMETRY 133<br />

Pre-lab Question 10.1 Write down the expression for ∆Ω, the solid angle subtended by the<br />

detector in the foil, in terms <strong>of</strong> the following quantities<br />

(a) d, the distance from the centre <strong>of</strong> the annulus to the detector;<br />

(b) ∆S, the area <strong>of</strong> the detector; and<br />

(c) r = (r 1 + r 2 )/2, the mean radius <strong>of</strong> the annulus.<br />

Remember that the solid angle may be defined as the dot product <strong>of</strong> two vectors. You may<br />

wish to draw a diagram showing the directions <strong>of</strong> those vectors.<br />

Solid angle comes into its own when one considers that a form <strong>of</strong> the cross section depends<br />

on angle. . .<br />

10.3.3 Differential cross section<br />

The cross section, as discussed above, is known as the total (or sometimes integral) cross<br />

section. The scattering depends on energy and, especially, angle, and when we wish to<br />

discuss the angular dependence the concept <strong>of</strong> differential cross section comes into being.<br />

This is the cross section associated with the scattering <strong>of</strong> particles through an angle θ into<br />

an infinitesimal solid angle, dΩ, subtended at the detector. Following the same argument as<br />

before, we find that the total number <strong>of</strong> particles scattered into a solid angle ∆Ω from a thin<br />

target ∆N sc (θ,ϕ), is given by<br />

∆N sc (θ,ϕ)<br />

N 0<br />

= nx dσ (θ,ϕ)∆Ω (10.4)<br />

dΩ<br />

where dσ/dΩ is the differential cross section. The relationship between the differential and<br />

total cross sections is<br />

∫ 2π ∫ π<br />

dσ<br />

σ = dϕ (θ,ϕ) sin θ dθ . (10.5)<br />

0 dΩ<br />

0<br />

Normally, the particle beam reaching the target is unpolarised, which means the differential<br />

cross section is not dependent on ϕ, hence<br />

∫ π<br />

dσ<br />

σ = 2π (θ) sin θ dθ . (10.6)<br />

0 dΩ<br />

That is the case in this experiment. The unit <strong>of</strong> the differential cross section is barn/steradian<br />

(b/sr).<br />

In terms <strong>of</strong> quantities that can be measured in this experiment, the differential cross section<br />

is<br />

dσ<br />

dΩ (θ) = ∆n s<br />

Nn i ∆Ω , (10.7)<br />

where n i and n s are the incident and scattering flux <strong>of</strong> particles per unit area, N is the number<br />

<strong>of</strong> Au nuclei, and ∆Ω is the solid angle subtended at the detector.


134 CHAPTER 10. RUTHERFORD SCATTERING<br />

Pre-lab Question 10.2 The number <strong>of</strong> atoms, N, in the foil must be known. This can be<br />

found by calculating the weight <strong>of</strong> Au in the scattering annulus and using Avogadro’s number.<br />

As the (exact) thickness is unknown, we approximate the number by<br />

N = ct atoms, (10.8)<br />

where c is a constant and t is the foil thickness in µm. Calculate the constant and check the<br />

value with your demonstrator.<br />

10.3.4 Theoretical cross section<br />

One <strong>of</strong> the earliest major successes <strong>of</strong> quantum mechanics was the confirmation that the<br />

theoretical differential cross section derived from quantum mechanics using the Coulomb<br />

interaction was the same as that derived by Rutherford classically. That cross section is<br />

given by<br />

( )<br />

dσ<br />

dΩ (θ) = Z1 Z 2 e 2 2<br />

1<br />

, (10.9)<br />

16πǫ 0 E K sin 4 θ 2<br />

where Z 1 is the charge (atomic number) <strong>of</strong> the target nucleus, Z 2 is the charge <strong>of</strong> the projectile,<br />

and E K is the kinetic energy <strong>of</strong> the projectile. e and ε 0 are the usual electron charge and<br />

permittivity <strong>of</strong> free space.<br />

Pre-lab Question 10.3 For your experiment<br />

dσ<br />

dΩ (θ) =<br />

k<br />

sin 4 θ 2<br />

. (10.10)<br />

Determine the constant k for your experiment to give the final value in m 2 . Check that value<br />

with your demonstrator.<br />

Pre-lab Question 10.4 Write down an expression for the scattering angle θ in terms <strong>of</strong> the<br />

foil-detector distance d.<br />

10.4 Experiment<br />

First, some words about safety...<br />

10.4.1 Safety precautions<br />

At all times, be wary <strong>of</strong> the following:<br />

10.4.1.1 Gold foil<br />

The Au foil is extremely fragile and must on no account be touched. Be sure to handle the<br />

foil frame only by the edges, and be very careful not to scratch or tear it.


10.4. EXPERIMENT 135<br />

10.4.1.2 Vacuum system<br />

Working with vacuum systems is a highly specialised art (and the subject <strong>of</strong> another experiment).<br />

While the system you will use is very simple, general rules apply:<br />

1. Do not touch the inside <strong>of</strong> the evacuated detector chamber. Any contamination will<br />

result in a loss <strong>of</strong> vacuum.<br />

2. Ensure that the exhaust from the pump is adequately ventilated. (The exhaust hose<br />

should go outside.)<br />

3. Finally, make sure that the whole vacuum chamber is sealed when the pump is running.<br />

It is inefficient and stresses the pump if a valve is left open or there is an air leak in the<br />

system.<br />

10.4.1.3 Detector<br />

The surface <strong>of</strong> the detector is extremely sensitive and must never be touched. Damage to<br />

the surface would render the detector useless and its replacement cost is several thousand<br />

dollars.<br />

The bias should be applied and taken <strong>of</strong>f in slow, even steps to prevent dielectric breakdown.<br />

Take at least 30 seconds to raise or lower the bias.<br />

There should be no bias on the detector during large changes in pressure. Make sure the bias<br />

is OFF whenever the chamber is pumped down or let up to air. Also turn the bias <strong>of</strong>f when<br />

moving the detector arm in case <strong>of</strong> leaks around the sliding O-ring.<br />

ALWAYS check the detector arm is clamped in place BEFORE the chamber is pumped<br />

down.<br />

10.4.1.4 Radiation<br />

There is a low risk <strong>of</strong> radiation exposure in this experiment. The decay length <strong>of</strong> α particles<br />

in air is around 3 cm - less than the distance from the source to the open face <strong>of</strong> its holder<br />

- and the source itself is almost always sealed inside the vacuum system, out <strong>of</strong> which no<br />

α particles can leak. Common sense in handling the source (especially during the spacer<br />

plate to foil changeover) should still be the norm, and don’t forget to wash your hands after<br />

handling radioactive sources.<br />

10.4.2 Relationship between counts and source to detector distance<br />

In the first part <strong>of</strong> the experiment, you will study the relationship between the distance from<br />

the source to a detector and N, the number <strong>of</strong> counts detected in a given time. Actually, this<br />

should look familiar to you....<br />

Pre-lab Question 10.5 What relationship do you expect this to follow? Why?


136 CHAPTER 10. RUTHERFORD SCATTERING<br />

10.4.2.1 Detector and electronics<br />

The electronics <strong>of</strong> the detection system are given in the block diagram shown in figure 10.3.<br />

Detector Pre−amp Amp SCA<br />

Bias<br />

Counter<br />

Figure 10.3: Block diagram <strong>of</strong> the electronics. SCA = Single Channel Analyser. The scaler<br />

and clock together form a counter, which is a single unit.<br />

The detector is a solid-state Si detector and is, in essence, a type <strong>of</strong> miniature ionisation<br />

chamber. The active volume is formed by the depletion region <strong>of</strong> a large area p-n junction.<br />

It is shown schematically in figure 10.4. The detector is a p-n junction where the charged<br />

particle enters the detector through the p-type layer. This layer is as thin as possible in order<br />

to minimise the energy loss through the material from ionisation. A reverse bias voltage<br />

applied as shown in figure 10.4 increases the volume <strong>of</strong> the depletion region.<br />

A charged particle entering the depletion region creates electron-hole pairs (free charges)<br />

which then move in the electric field induced by the positive and negative charges created<br />

by the application <strong>of</strong> the bias voltage in the p- and n-type layers. The mean energy loss <strong>of</strong> a<br />

charged particle required to produce one electron-hole pair in Si is 3.5 eV, and so 2.87 × 10 5<br />

such pairs are produced on average per MeV energy deposited.<br />

The effect <strong>of</strong> the total energy lost in the depletion layer is to produce a potential V dependent<br />

on the total amount <strong>of</strong> charge collected in the capacitor: V = Q/C, where Q is the charge<br />

collected. The capacitor recharges through a resistor after the discharge. A charge-sensitive<br />

preamplifier produces a signal whose amplitude is proportional to the charge collected and<br />

hence also to the energy <strong>of</strong> the detected charged particle.<br />

p−type<br />

n−type<br />

incident<br />

particle<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

−<br />

−<br />

−<br />

−<br />

−<br />

−<br />

−<br />

−<br />

−<br />

−<br />

EMF<br />

depletion<br />

layer<br />

Figure 10.4: Detector and response as a function <strong>of</strong> time.


10.4. EXPERIMENT 137<br />

That signal is then passed onto the (high-gain) amplifier which produces a pulse with amplitude<br />

0 − 10 V, still proportional ultimately to the energy <strong>of</strong> the detected particle, and<br />

whose shape characteristics are such that it may be analysed in subsequent hardware. In<br />

our case, that is the Single Channel Analyser. The Single Channel Analyser (SCA) acts as<br />

a filter transmitting only those pulses whose amplitudes fall within a predefined window to<br />

the counter. The two modes <strong>of</strong> operation are “window” and “normal”. The Lower-Level<br />

Discriminator (LLD) sets the lower voltage level <strong>of</strong> the pulses to accept. In normal mode,<br />

the Upper-Level Discriminator (ULD) sets the upper voltage level, while in window mode,<br />

the ULD scale is reduced by a factor <strong>of</strong> 10, and that new level then acts as the window above<br />

the LLD for the acceptable pulses.<br />

The Counter simply counts the number <strong>of</strong> signals it receives, incrementing by one whenever<br />

a signal reaches it. In conjunction with the SCA it forms an Analogue to Digital Converter.<br />

A timer connected to the counter allows for counting to be done within preset times.<br />

The distribution <strong>of</strong> counts in this detection system is Gaussian, and for N counts, the uncertainty<br />

is ± √ N.<br />

10.4.2.2 Foil and source<br />

The source <strong>of</strong> α particles is 241 Am which produces α particles <strong>of</strong> general discrete energy.<br />

The ones <strong>of</strong> interest to us are those with 5.48 MeV <strong>of</strong> energy. The source is secured at the<br />

end <strong>of</strong> the scattering chamber by a plate which exposes a limited area <strong>of</strong> the source. As<br />

protection from contamination, the active material is enclosed between a backing layer <strong>of</strong><br />

silver and a 3 mm thick gold window layer. The window reduces the energy <strong>of</strong> the emitted<br />

particles to 4.66 MeV with a spread <strong>of</strong> 0.46 MeV.<br />

The Au foil that you will use is evaporated onto a mylar backing (mylar is essentially all C<br />

and H). For the purposes <strong>of</strong> scattering, we can ignore the mylar due to its much lower mass<br />

and number compared to the Au.<br />

The Au foil is extremely fragile and must ON NO ACCOUNT be touched. Great care must<br />

be taken in handling it.<br />

There are two inserts that fit between the source and the detector. One has the actual foil<br />

attached, with an annulus cut into an Al backing plate. The second is simply an open annulus<br />

which acts as a spacer, to ensure that the source is placed at the same relative distances as<br />

when the foil is in place.<br />

10.4.2.3 Vacuum system<br />

The scattering chamber is shown schematically in figure 10.5. The pressure in the scattering<br />

chamber must be reduced since the range <strong>of</strong> α particles in air at atmospheric pressure is only<br />

about 3 cm. The pump provided will reduce the pressure in the chamber to about 10 −2 Torr<br />

in 5 minutes, greatly increasing their range.<br />

As shown in figure 10.5, O-ring seals are used at various locations. The shaft slides through<br />

O-rings and care should be taken when moving the shaft under vacuum to avoid deformation,<br />

which would cause leaks. At all times turn the detector bias DOWN before moving the


138 CHAPTER 10. RUTHERFORD SCATTERING<br />

To vacuum<br />

pump<br />

Air intake<br />

valve<br />

Protective clamp<br />

Detector<br />

To electronics<br />

241<br />

Am source<br />

Au foil<br />

annulus<br />

O−ring seals<br />

Figure 10.5: The scattering chamber.<br />

shaft.<br />

10.4.2.4 Data collection<br />

Place the spacer plate in the chamber and evacuate the apparatus.<br />

Question 10.1 Why is the experiment done under vacuum?<br />

Once the chamber is evacuated, SLOWLY raise the bias voltage <strong>of</strong> the detector to 30 V and<br />

view the signal out <strong>of</strong> the detector on the CRO, with the detector at a distance <strong>of</strong> 1 cm. Set<br />

the LLD and ULD on the SCA and check that the electronics are working.<br />

Question 10.2 To what did you set the LLD and ULD? Why?<br />

Measure the counts reaching the detector over a wide range <strong>of</strong> detector-plate distances, from<br />

1 cm to 20 cm, using a suitable counting time.<br />

Question 10.3 What determines a suitable count time? Check with your demonstrator before<br />

proceeding.<br />

With the count time reset to 5 minutes, record the number reaching the detector at a distance<br />

<strong>of</strong> 1 cm.<br />

10.4.2.5 Analysis<br />

Plot the inverse root <strong>of</strong> the count versus distance. Make sure to include the errors you have<br />

already calculated.


10.4. EXPERIMENT 139<br />

Question 10.4 What do you expect this relationship to be?<br />

Using this calibration graph, determine the expected number <strong>of</strong> counts when the detector is<br />

sitting in the plane <strong>of</strong> the foil for 5 minutes.<br />

We must determine the incident flux on the foil. This can be calculated after the number <strong>of</strong><br />

α particles passing to the detector have been measured. The flux incident on the foil is<br />

n i = C 0<br />

A DET<br />

[<br />

DS−O<br />

D S−F<br />

] 2<br />

cos θ R , (10.11)<br />

where C 0 is the number <strong>of</strong> counts per second at zero distance, A DET is the area <strong>of</strong> the detector<br />

in units <strong>of</strong> m 2 , D S−O is the distance from the source to zero distance, D SF is the source-foil<br />

distance, and θ R is defined in figure 10.1.<br />

Pre-lab Question 10.6 With the aid <strong>of</strong> diagrams, derive equation 10.11.<br />

From your 5 minute measurement, determine the average incident flux <strong>of</strong> α particles on the<br />

foil, as described in Pre-lab question 10.6.<br />

10.4.3 Rutherford scattering<br />

We now turn our attention to recreating the Rutherford experiment. The results will be able<br />

to distinguish between the plum pudding and nuclear models <strong>of</strong> the atom (we hope!), and,<br />

by assuming the correct model, we may be able to determine the thickness <strong>of</strong> the Au foil.<br />

Lower the bias on the detector to zero, SLOWLY, and let the chamber up to air. Replace the<br />

spacer plate with the plate holding the Au foil, and evacuate the chamber. Restore the bias<br />

to the detector for measurements. Note that due to the thin foil and low source strength, long<br />

count times will be required. In fact, this point is so important, we take the rare opportunity<br />

to repeat a question for your consideration:<br />

Question 10.5 What determines a suitable count time? Check with your demonstrator before<br />

proceeding.<br />

At least 4 data points must be taken (the more the better), and so it is strongly recommended<br />

that at least one point be taken in the first session, one point early in the morning, one around<br />

lunch time, and two during the second session 2 .<br />

Consult with your demonstrator and select a range <strong>of</strong> distances for your measurement. Use<br />

figure 10.6 to determine suitable distances and count times. As you take the data, consider<br />

the analysis and prepare the calculations using Excel.<br />

2 This assumes that the first session is in the afternoon, which is nearly always the case.


140 CHAPTER 10. RUTHERFORD SCATTERING<br />

1.0<br />

Counts (arb units)<br />

0.8<br />

0.6<br />

0.4<br />

total<br />

detector orientation<br />

inverse square law<br />

0.2<br />

0.0<br />

0 5 10 15 20<br />

d (cm)<br />

Figure 10.6: Expected number <strong>of</strong> counts as a function <strong>of</strong> detector-foil distance. Use this<br />

diagram to decide on counting times.<br />

10.4.3.1 Analysis<br />

By equating the data with the theoretical (Rutherford) cross section, one may determine the<br />

thickness <strong>of</strong> the foil. Assuming that the measured and calculated cross sections are equal<br />

dσ<br />

dΩ∣ = dσ<br />

∣<br />

THEORY<br />

dΩ<br />

∣<br />

EXP<br />

k<br />

sin 4 (θ/2) = 1 ∆n s<br />

t cn i ∆Ω . (10.12)<br />

By plotting the ratio <strong>of</strong> the theoretical cross section to t times the experiment cross section,<br />

as a function <strong>of</strong> d, one may infer an average value for the thickness <strong>of</strong> the foil.<br />

Be sure to include a sample calculation showing how you converted from the raw data to the<br />

cross section.<br />

Question 10.6 What value have you obtained for the thickness <strong>of</strong> the foil?<br />

Question 10.7 Using this value, re-calculate the experimental cross section and plot as a<br />

function <strong>of</strong> the scattering angle. How does it compare with Rutherford’s formula?


10.5. USEFUL DATA 141<br />

10.4.3.2 Further considerations<br />

You may wish to discuss, qualitatively, the effects <strong>of</strong> the following on the results:<br />

• the finite width <strong>of</strong> the scattering foil, and the corresponding spread in the scattering<br />

angle;<br />

• the effect <strong>of</strong> alpha particles striking the walls <strong>of</strong> the chamber;<br />

• the change in the alpha particle energy as it passes through the foil; and<br />

• the variation in mean path length <strong>of</strong> the particles through the foil as a function <strong>of</strong><br />

scattering angle.<br />

10.5 Useful data<br />

Quantity<br />

Value<br />

Permittivity <strong>of</strong> free space, ε 0<br />

8.8542 × 10 −12 F/m<br />

Energy <strong>of</strong> α particles emitted by 241 Am 5.48 MeV<br />

Mean energy and spread <strong>of</strong> 5.48 MeV α particles<br />

after passing through 3 µm <strong>of</strong> Au<br />

4.66 ± 0.23 MeV<br />

Atomic weight <strong>of</strong> Au<br />

196.97 g/mole<br />

Density <strong>of</strong> Au<br />

19.3 g.cm −3<br />

Atomic number <strong>of</strong> Au 79<br />

Atomic number <strong>of</strong> α 2<br />

Avogadro’s number<br />

6.02205 × 10 25 mole −1<br />

1 eV 1.60219 × 10 −19 J<br />

e<br />

1.60219 × 10 −19 C<br />

d<br />

)<br />

x<br />

a<br />

dx a<br />

a 2 +x 2<br />

<strong>Bibliography</strong><br />

[1] J J Sakurai. Modern Quantum Mechanics. Addison-Wesley, 1994.


Chapter 11<br />

β spectroscopy<br />

11.1 Abstract<br />

Of the four fundamental forces in nature, the most beguiling is the weak nuclear force. Its<br />

sole purpose is to mediate β decay and electron capture processes in the nucleus, and is the<br />

only mechanism by which the neutrino may (or may not, as is much more likely) interact.<br />

The purpose <strong>of</strong> this experiment is to observe β decay, which has become one <strong>of</strong> the most<br />

fundamental processes <strong>of</strong> nature by which we can, amongst other things, understand the<br />

standard model <strong>of</strong> particle physics.<br />

11.2 Introduction<br />

The three (main) nuclear decays were established by the 1930s. The first was designated α,<br />

the second β, the third γ. Rutherford, in a very elegant experiment, proved that α particles<br />

were 4 He nuclei. The ubiquitous high energy photon proved to be the γ radiation. The β<br />

particles are the electron (β − ) and positron (β + ) and only one or the other may be produced<br />

as a result <strong>of</strong> β decay.<br />

The early studies <strong>of</strong> β decay were most controversial. β decay is the process by which<br />

a nucleus decays leaving the daughter nucleus with the same mass but changing the proton<br />

and neutron numbers by 1 (Z ±1,N ∓1) after the emission <strong>of</strong> either the electron or positron.<br />

Figure 11.1: Early beta spectrum from the decay <strong>of</strong> Ra. The measurement is from 1927 [1].<br />

143


144 CHAPTER 11. β SPECTROSCOPY<br />

Figure 11.1 shows an early beta spectrum (1927) as measured from radium 1 . The astounding<br />

observation was that this spectrum was continuous. Only two particles were known in nature<br />

at the time and, with the neutron around the corner, the only known processes for the decay<br />

were<br />

p → n + e +<br />

n → p + e − .<br />

The two-body kinematics dictated that the spectrum <strong>of</strong> β particles should be discrete at a<br />

fixed value. The observation <strong>of</strong> a continuous spectrum led to a crisis in the physics community:<br />

how could this be? It appeared that conservation <strong>of</strong> energy was violated. A famous<br />

exchange <strong>of</strong> letters between Heisenberg and Pauli discussed whether conservation <strong>of</strong> energy<br />

was finally observed to be violated at the nuclear level; after all, quantum physics gave us<br />

uncertainty which was revolutionary. Heisenberg was uncertain 2 and thought that energy<br />

conservation might have been violated. Pauli was, however, adamant: energy conservation,<br />

he said, was inviolate at all levels - there had to be a third, as yet unobserved, particle taking<br />

some energy away, which would then lead to a continuous β spectrum. He was first to postulate<br />

the neutrino. He postulated the neutrino in 1931; it wasn’t discovered until 1953 by F.<br />

Reines and C. L. Cowan. Reines was awarded the Nobel Prize in 1995 for their discovery.<br />

As a result, the corrected underlying processes <strong>of</strong> nuclear β decay are 3<br />

For nuclear β ± decay itself the processes are<br />

p → n + e + + ν<br />

n → p + e − + ¯ν .<br />

A<br />

ZM N → A Z∓1M ′ N±1 + e ± +<br />

{ ν¯ν .<br />

(In this equation, M is the decaying nucleus <strong>of</strong> mass A with Z protons and N neutrons.)<br />

The decay can only occur if the conditions are energetically favourable, i.e., that the binding<br />

energy <strong>of</strong> M is higher than that <strong>of</strong> the daughter nucleus M ′ . While the mass number is<br />

conserved, there is a very slight difference in the rest masses <strong>of</strong> the parent and daughter<br />

nuclei from the change <strong>of</strong> a proton to a neutron or vice-versa.<br />

11.3 Background concepts<br />

This brief section is based on Fermi’s theory <strong>of</strong> β decay. However, all descriptions (Fermi,<br />

Gamow-Teller, field-theoretic, electroweak, etc.) <strong>of</strong> β decay are beyond the scope <strong>of</strong> second<br />

year level and we only present the concepts involved and which variables are defined to<br />

facilitate measurement.<br />

The Fermi model assumes that the decay occurs according to the model presented in figure<br />

11.2. The interaction effecting the decay is in this case considered a contact interaction,<br />

1 The neutron was discovered by Chadwick in 1932.<br />

2 Sorry.<br />

3 where ν is a neutrino, which is chargeless. It is important to note that until very recently the neutrino was


11.3. BACKGROUND CONCEPTS 145<br />

n (p)<br />

p (n)<br />

000 111<br />

000 111<br />

000 111<br />

e + (e − )<br />

ν (ν)<br />

Figure 11.2: Diagrammatic representation <strong>of</strong> the Fermi model <strong>of</strong> β ± decay. The brackets<br />

denote the process <strong>of</strong> neutron decay.<br />

i.e. where the interaction between the particles participating in the reaction all interact at a<br />

point 4 .<br />

By considering this picture, one can define a new variable <strong>of</strong> measurement, which is derived<br />

from a knowledge <strong>of</strong> the probability <strong>of</strong> the process, and the kinematics involved. That<br />

variable is called the Kurie variable and is defined by<br />

K(Z,p) =<br />

√<br />

w(p)<br />

p 2 F(Z,p) =<br />

8π<br />

h 7/2 c 3/2 |M if| (E 0 − E), (11.1)<br />

where<br />

2πη Ze2<br />

F(Z,p) =<br />

1 − e−2πη,η =<br />

4πǫ 0 v , (11.2)<br />

v is the velocity <strong>of</strong> the electron, and w(p) and |M if | are measures <strong>of</strong> the probability <strong>of</strong> the<br />

decay for a particular β momentum p and F(Z,p) is known as the Fermi function. E is the β<br />

energy. One then expects that a plot <strong>of</strong> K against E to be a straight line with negative slope.<br />

As w(p) is related to the number <strong>of</strong> counts at each β energy and momentum, one can replace<br />

the probability by the number <strong>of</strong> counts for a given time and define an experimental Kurie<br />

variable by<br />

K(Z,p) =<br />

√<br />

N(p)<br />

p 2 F(Z,p) . (11.3)<br />

Pre-lab Question 11.1 What is the value <strong>of</strong> K when the β energy is equal to E 0 ?<br />

Pre-lab Question 11.2 What is then the probability <strong>of</strong> a β particle being emitted with an<br />

energy E 0 ?<br />

assumed to have zero mass. The neutrino mass was finally discovered (∼ O(0.1eV)) in measurements taken<br />

with the very large neutrino detectors at Kamiokande (Japan) and Sudbury (Canada). For all practical purposes,<br />

a zero mass neutrino is still a good approximation.<br />

4 the more modern description involving the full weak theory <strong>of</strong> β decay involves the exchange <strong>of</strong> the W ±<br />

and Z particles.


146 CHAPTER 11. β SPECTROSCOPY<br />

Evacuated chamber<br />

r<br />

e<br />

−<br />

To electronics<br />

147<br />

Pm source<br />

Geiger tube<br />

Figure 11.3: Magnetic spectrometer showing the configuration with the 147 Pm source and<br />

detector. The radius <strong>of</strong> the path the β traverses is r = 3.80 ± 0.01 cm.<br />

Pre-lab Question 11.3 How does the intercept with the energy axis change with neutrino<br />

mass?<br />

11.4 Experiment - the β decay <strong>of</strong> promethium<br />

You will be studying the β decay <strong>of</strong> 147<br />

61 Pm which decays into 147<br />

62 Sm. The process is<br />

147 Pm → 147 Sm + e − + ¯ν ,<br />

so, for this decay the β particles are electrons. The decay from promethium may lead to<br />

several states in samarium, with the total number <strong>of</strong> counts from the decay being distributed<br />

among the various states that are accessible. This will be reflected in the energy spectra<br />

collected.<br />

Pre-lab Question 11.4 What do you think might influence the count rate going into each<br />

state? Discuss with your demonstrator.<br />

11.4.1 Calibrating the magnet<br />

To measure the energy <strong>of</strong> the electrons in the experiment we use a magnetic spectrometer<br />

connected to a Geiger-Müller tube. The spectrometer is shown in figure 11.3. A particle <strong>of</strong><br />

mass m, charge q, and velocity v, moving in a uniform magnetic field B, experiences a force<br />

F where<br />

F = qv × B. (11.4)


11.4. EXPERIMENT - THE β DECAY OF PROMETHIUM 147<br />

If B is perpendicular to v then the particle will move in a circle <strong>of</strong> radius r where<br />

mv 2<br />

r<br />

= qvB (11.5)<br />

with momentum p = qrB. One can therefore select the momentum <strong>of</strong> the charged particles<br />

by adjusting the magnetic field given that the path taken from source to detector is fixed.<br />

In order to measure the magnetic field, we use a duplicate spectrometer which has identical<br />

dimensions to the actual spectrometer but without the source or detector present. Mount the<br />

duplicate spectrometer between the poles <strong>of</strong> the electromagnet. Initialise the Hall probe:<br />

• Place the tip <strong>of</strong> the probe in the zero field case and switch on the meter.<br />

• When the number 1973 appears on the display, press the ENTER/RESET button.<br />

• At this point three dashes should appear on the display.<br />

• Press the PB button.<br />

• Press the orange UP/DOWN keys until KGAUSS is selected.<br />

• The Hall probe is now ready for use.<br />

• NB! Do NOT switch <strong>of</strong>f the meter from this point until ALL <strong>of</strong> the data have been<br />

taken. If the meter is switched <strong>of</strong>f it will have to be re-initialised.<br />

Once the Hall probe is ready for use, insert it into the duplicate chamber without switching<br />

<strong>of</strong>f the meter. Remove any remnant field by degaussing the chamber. If the magnitude <strong>of</strong> the<br />

field is still above 10 G, degauss again. If the magnetic field persists, see your demonstrator.<br />

Question 11.1 Explain how the degausser works.<br />

Connect the magnet to the stabilised DC supply, making sure the current is flowing into the<br />

red plug <strong>of</strong> the magnet and out <strong>of</strong> the black. Use the multimeter to measure the current.<br />

Before switching on the supply check that:<br />

• The current is turned to a maximum.<br />

• The voltage is at a minimum.<br />

• The multimeter is on the 0 − 2 A DC scale.<br />

Measure the magnetic field as a function <strong>of</strong> current from 0 to 500 mA in 20 mA steps. Note<br />

that you should:<br />

• Never reduce the current: backtracking at any stage will render the calibration useless.<br />

• Never change scales on the hall probe or multimeter.<br />

• Allow a few seconds for the field to stabilise before taking a measurement.


148 CHAPTER 11. β SPECTROSCOPY<br />

~ 1 kV<br />

cathode<br />

β<br />

anode<br />

electron cascade<br />

Figure 11.4: Operation <strong>of</strong> a Geiger-Müller tube.<br />

Question 11.2 Why would backtracking render the calibration useless?<br />

You may notice a small black button on the side <strong>of</strong> the Hall probe. When the magnetic field<br />

is entering the Hall probe from this side, the gaussmeter will give a positive reading.<br />

Question 11.3 What is the direction <strong>of</strong> the magnetic field for this direction <strong>of</strong> current?<br />

After you have completed the measurement, slowly reduce the current to zero and remove<br />

any remnant field. Find the calibration <strong>of</strong> the Hall probe and check the slope and intercept<br />

<strong>of</strong> the calibration with your demonstrator.<br />

Question 11.4 What is the maximum kinetic energy you will measure?<br />

Question 11.5 To what magnetic field, and thus current, does this energy correspond?<br />

11.4.2 Measuring the spectrum<br />

Now that the Hall probe has been calibrated, you are now able to measure the magnetic field<br />

inside the spectrometer by measuring the electromagnet’s current. By measuring the number<br />

<strong>of</strong> counts as a function <strong>of</strong> current one finds the β spectrum.<br />

11.4.2.1 The Geiger-Müller Tube<br />

A schematic <strong>of</strong> the detector used in this experiment is given in figure 11.4. The tube consists<br />

<strong>of</strong> an easily ionisable gas contained within a thin-walled chamber. A high voltage is applied


11.4. EXPERIMENT - THE β DECAY OF PROMETHIUM 149<br />

between the central anode and the walls <strong>of</strong> the chamber creating a very high electric field<br />

within the chamber. After entering the detector, a β − ionises the gas. Each collision liberates<br />

a secondary electron which is accelerated towards the anode, together with the original β − .<br />

The high electric field provides extra energy to the electrons as they travel towards the anode.<br />

The mean free path <strong>of</strong> electron in the gas is very small, so the electrons undergo many<br />

collisions before reaching the anode. So the original β − initiates an electron cascade which<br />

results in a large charge at the anode. This charge is collected by the detector electronics.<br />

As the electrons move towards the anode, the positively charged ions move towards the<br />

cathode but much more slowly as they are much, much heavier. Hence if another β − enters<br />

the detector before a significant number <strong>of</strong> positive ions have been discharged at the cathode<br />

no cascade is possible as the ions will recombine with the secondary electrons induced by<br />

the second β − . So the tube is rendered useless until the event has been cleared from the<br />

chamber. This is known as dead time, denoted by τ. It comes about as a combination <strong>of</strong> the<br />

detector and electronics, which require a finite time to collect the charge from the detector<br />

coming from each detected particle.<br />

Pre-lab Question 11.5 Show that if the tube has a dead time τ, and N counts are observed<br />

in a time t, the true number <strong>of</strong> events, N ′ , is given by<br />

N ′ =<br />

N<br />

1 − τ t N . (11.6)<br />

The dead time for the detector you will be using has been measured, at a count rate <strong>of</strong><br />

∼ 1000 counts/sec, to be 20.0 ± 0.5 µs. To a good approximation, the dead time may be<br />

considered constant for the spectra you will measure.<br />

11.4.3 The spectrometer<br />

Ensure that you have disconnected the power supply and degaussed the magnet. Carefully<br />

remove the Hall probe, switch it <strong>of</strong>f, and replace it in its case. Replace the duplicate chamber<br />

with the actual spectrometer containing the source and detector.<br />

NB!! All the data for this part <strong>of</strong> the experiment must be taken at once. It cannot be split<br />

over two days. It is recommended that you take the data on the first day and do the analysis<br />

on the second day, leaving time on the second day to take more data if the need so arises.<br />

Question 11.6 Why must all the data be taken at once?<br />

Connect the DC supply to the magnet.<br />

Question 11.7 In which direction are the β particles moving as a result <strong>of</strong> the magnetic<br />

field?<br />

Ensure that the vacuum pump is connected to the chamber and switch on the pump and<br />

gauge. After a few minutes the pressure in the chamber should be below 10 −2 Torr. Connect<br />

up the circuit shown in figure 11.5. Set the electronics to the following settings and then<br />

switch on the NIM bin:


150 CHAPTER 11. β SPECTROSCOPY<br />

DET.<br />

AMP<br />

SCA<br />

COUNT<br />

H.V.<br />

CRO<br />

TIMER<br />

Figure 11.5: Electronics for the detection <strong>of</strong> β particles.<br />

Amplifier Course gain = 16; fine gain = 6; input = positive; output = unipolar.<br />

SCA ULD = 10 V; LLD = 0.20 − 0.30 V.<br />

Counter Discriminator = 3.0 V.<br />

Timer 100 s.<br />

Start measuring the number <strong>of</strong> counts per 100 s in the current range 0 to 400 mA, taking<br />

suitable steps across the range. Remember NEVER to backtrack, always increasing the<br />

current, and allowing for a few seconds before counting each time you change the current to<br />

allow the field to stabilise.<br />

As the data are taken, observe the output pulses from the amplifier on the CRO.<br />

Question 11.8 Draw, on scaled axes, the shape <strong>of</strong> the output. At which current do they<br />

appear smallest?<br />

Check, for a current ∼ 100 mA, that the output does not exceed the upper level on the SCA.<br />

Discuss the shape <strong>of</strong> the pulses in terms <strong>of</strong> the operation <strong>of</strong> the detector.<br />

Question 11.9 Are the pulses consistent with the quoted dead time?<br />

After the first set <strong>of</strong> measurements, increase the current to 500 mA and take one more datum.<br />

This will serve as the background measurement, which must be subtracted from each <strong>of</strong> your<br />

previous measurements.<br />

Question 11.10 Why is this a suitable current at which to take such a background reading?<br />

11.4.4 Analysis<br />

You should now have a table <strong>of</strong> the measurements along with the appropriate uncertainties.<br />

The analysis <strong>of</strong> these data requires a series <strong>of</strong> calculations and great care must be taken to<br />

ensure the transformations are done correctly. The steps you make should be made clear; be<br />

sure to comment on all aspects <strong>of</strong> the analysis.<br />

Use Excel to correct the number <strong>of</strong> counts for background and dead time, and calculate the<br />

electron kinetic energies from the deduced momenta. Plot the corrected counts against the


11.5. USEFUL DATA 151<br />

kinetic energy (in keV). Explain the shape <strong>of</strong> the spectrum, discussing its features in terms<br />

<strong>of</strong> both the physics and the apparatus used.<br />

From this spectrum, and again using Excel, transform the spectrum into a Kurie plot. Fit the<br />

data to a straight line in the region <strong>of</strong> interest and obtain a value for E 0 .<br />

Question 11.11 How does your deduced value for E 0 compare with the quoted value for the<br />

decay <strong>of</strong> 147 Pm?<br />

Question 11.12 Is the experiment sensitive enough to allow for a determination <strong>of</strong> the neutrino<br />

mass? (??) Discuss, if you are game enough. . .<br />

11.5 Useful data<br />

Quantity<br />

Accepted value for E 0 , for the decay <strong>of</strong> 147 Pm<br />

h<br />

c<br />

ε 0<br />

Value<br />

224.5 ± 0.4 keV<br />

6.6256 × 10 −34 Js<br />

2.997925 × 10 8 ms −1<br />

8.85416 × 10 −12 Fm −1<br />

1.6022 × 10 −19 C<br />

e<br />

1 eV 1.6022 × 10 −19 C<br />

electron rest mass E e<br />

511 keV c −2<br />

<strong>Bibliography</strong><br />

[1] K Heyde. Basic Ideas and Concepts in Nuclear <strong>Physics</strong>. Institute <strong>of</strong> <strong>Physics</strong> Publishing,<br />

2 nd edition, 1999.


Chapter 12<br />

γ ray detection and spectroscopy<br />

12.1 Abstract<br />

How does one detect radiation? This series <strong>of</strong> experiments is designed to give you some<br />

familiarity with detectors, in this case, a gamma-ray detector. Along the way, you may learn<br />

a little bit about how a nucleus decays. . .<br />

12.2 Introduction<br />

A piece <strong>of</strong> “something” is placed in front <strong>of</strong> you. Placing something else next to it elicits a<br />

response in that something else. At once you may infer that something had been transferred<br />

between the first something and the other something, giving that measurable response in the<br />

other. Depending on our understanding <strong>of</strong> the other we may gain an understanding <strong>of</strong> what<br />

was transferred.<br />

This is, in essence, what happens in a detector when a radioactive source is placed next to it.<br />

Our ability to detector whichever radiation being produced depends on the response <strong>of</strong> the<br />

detector to that radiation, and our understanding <strong>of</strong> the processes involved in that response.<br />

No experiment may be done without this knowledge.<br />

For this experiment, we will examine the response <strong>of</strong> a typical γ ray detector to incident γ<br />

rays. In achieving this goal, we need to understand how photons interact with matter, and<br />

what happens when it does. The techniques and principals involved are adaptable to all cases<br />

<strong>of</strong> particle detection.<br />

12.3 γ radiation. How?<br />

We need to understand how a γ source works, in order to understand the spectrum <strong>of</strong> photons<br />

produced by it. First, and most importantly, γ radiation is the product <strong>of</strong> another nuclear<br />

decay and/or reaction - ALWAYS! A γ is produced as a result <strong>of</strong> a transition between an<br />

excited state <strong>of</strong> a nucleus to a lower energy state in that nucleus, proceeding down to the<br />

153


154 CHAPTER 12. γ RAY DETECTION AND SPECTROSCOPY<br />

706.41<br />

57<br />

Co<br />

0.18%<br />

EC<br />

366.74<br />

57<br />

Fe<br />

136.47<br />

14.41<br />

99.82%<br />

Figure 12.1: Decay from 57 Co by electron capture to 57 Fe showing the three possible γ rays<br />

from 57 Fe produced: 136.47, 122.06, and 14.41 keV. All energies in the figure are in keV.<br />

ground state. It is impossible to produce a lump <strong>of</strong> matter in which all, or indeed any, <strong>of</strong> the<br />

nuclei are in an excited state.<br />

To produce a γ ray within a source, one starts <strong>of</strong>f with a radioactive source which would<br />

decay in one <strong>of</strong> the other modes (α decay, β decay or electron capture) leaving the daughter<br />

nucleus in an excited state which then decays to its ground state by γ emission. A radioactive<br />

source is defined as a source <strong>of</strong> nuclei which may decay by those modes with a suitably long<br />

half-life. That half-life must be long enough (at least a few hours) in order to produce a<br />

mass <strong>of</strong> it which can be stored and used. Figure 12.1 shows the production <strong>of</strong> a 14.41 keV γ<br />

from 57 Fe as a result <strong>of</strong> the decay <strong>of</strong> 57 Co. The half-life <strong>of</strong> 57 Co is 271 days so this is a very<br />

suitable candidate <strong>of</strong> very low energy γ rays. The other sources in this experiment operate<br />

by similar decays and transitions.<br />

12.4 Calibration <strong>of</strong> the detector<br />

You will be using NaI(Tl) scintillators as γ detectors in this experiment. In order to interpret<br />

the results accurately, we must understand how photons interact with matter. This is not so<br />

easy: the photon is not charged and so it does not interact with matter by ionising the atoms<br />

in the material (as is the case with charged particles). The thallium (Tl, Z = 81) dopant is<br />

critical in the detection process.<br />

The circuit for this part <strong>of</strong> the experiment is shown in figure 12.2.<br />

12.4.1 Detecting photons - the NaI(Tl) crystal<br />

The photon interacts with matter in three principle ways, depending on its energy.


12.4. CALIBRATION OF THE DETECTOR 155<br />

NaI<br />

Pre−<br />

amp Amp MCA<br />

CRO<br />

Figure 12.2: Circuit for detecting γ rays and calibrating the detection system.<br />

hν’<br />

hν<br />

θ<br />

Figure 12.3: Kinematics <strong>of</strong> Compton scattering.<br />

φ<br />

T<br />

12.4.1.1 Photoelectric effect<br />

In this process 1 , a photon is absorbed by an atomic electron and imparts all <strong>of</strong> its energy<br />

to it. The electron is ejected from the atom with most <strong>of</strong> its kinetic energy coming from<br />

the photon. (Some <strong>of</strong> it is absorbed in the recoil <strong>of</strong> the nucleus, but for the most part this<br />

is negligible given the mass <strong>of</strong> the nucleus.) The kinetic energy <strong>of</strong> the ejected electron is<br />

thus E = E γ − E B , where E B is the binding energy <strong>of</strong> that electron to the atom. A little<br />

energy also comes from the X-ray that is produced from the transition when the hole left by<br />

the (inner shell) electron is filled by an outer shell electron. The probability <strong>of</strong> this process<br />

∼ Z 4 , so a high-Z material is most useful in the detection. Hence the Tl dopant.<br />

The primary photoelectron then ionises the surrounding material producing light in the visible<br />

spectrum from the subsequent transitions.<br />

12.4.1.2 Compton scattering<br />

In this process, the photon undergoes scattering on free electrons. (In the detector material,<br />

the process is on the outermost electrons which are very loosely bound and as such may be<br />

considered “free” to a good approximation.) The process is illustrated in figure 12.3.<br />

The energy <strong>of</strong> the scattered electron depends on the angle <strong>of</strong> scattering and is given by<br />

E e = E γ<br />

α(1 − cos θ)<br />

1 + α(1 − cos θ) , (12.1)<br />

1 for which Einstein won the Nobel Prize coming from his three masterworks <strong>of</strong> 1905 (the other two were<br />

Brownian motion and Special Relativity - for which he did NOT win a prize!!)


156 CHAPTER 12. γ RAY DETECTION AND SPECTROSCOPY<br />

Phosphor screen<br />

Dynodes<br />

1100<br />

01<br />

01<br />

00 11 01<br />

00 11 01<br />

1100<br />

01<br />

01<br />

00 11 01<br />

00 11 01<br />

To base<br />

and electronics<br />

NaI(Tl)<br />

crystal<br />

Cathode<br />

Light guide<br />

Anode<br />

Figure 12.4: Configuration <strong>of</strong> NaI(Tl) detector with photomultiplier tube. The dynodes <strong>of</strong><br />

the photomultiplier are shown.<br />

where E γ is the photon energy, E e is the scattered electron energy, and α = E γ /(m e c 2 );<br />

m e = 511 keV/c 2 is the rest mass energy <strong>of</strong> the electron.<br />

The scattered electron then ionises the surrounding material and also liberates visible-light<br />

photons. The scattered γ undergoes more and more scattering events, scattering more electrons<br />

which produce more light, until the original photon expends all its energy.<br />

12.4.1.3 Pair production<br />

If the incident photon has an energy <strong>of</strong> more than twice the rest mass energy <strong>of</strong> an electron,<br />

the third process is possible. Pair production takes place in a region <strong>of</strong> a high Coulomb<br />

field, where the photon is converted into an electron-positron pair. The Tl nucleus provides<br />

that strong Coulomb field. The electron then ionises the surrounding material as in the other<br />

cases while the positron collides with another electron causing annihilation into two 511 keV<br />

photons, which then also may be detected.<br />

The upshot <strong>of</strong> all three processes is that an electron is produced which ionises the surrounding<br />

material producing visible light [E ∼ O(1 eV)] photons. The NaI crystal has a refractive<br />

index <strong>of</strong> ∼ 2 so the probability <strong>of</strong> total internal reflection from the mirrored sides <strong>of</strong> the<br />

detector is quite high. The far side <strong>of</strong> the detector has a glass window which is optically<br />

coupled to a photomultiplier tube. The configuration is shown in figure 12.4.<br />

The light produced in the crystal is guided to the phosphor screen at the entrance to the<br />

photomultiplier tube. Each photon liberates an electron via the photoelectric effect. That<br />

electron then hits the cathode, which is the first dynode in the chain. Each dynode is under<br />

high voltage (V ∼ 1 kV, typically) and so each collision has the effect <strong>of</strong> creating an electron<br />

shower. That shower <strong>of</strong> electrons proceeds to the next dynode, producing another shower per<br />

electron. There are typically, 10 to 14 dynode stages in a photomultiplier tube and, overall,<br />

this gives a cascade <strong>of</strong> electrons which provides a gain <strong>of</strong> ∼ 10 7 electrons at the anode per<br />

electron produced at the phosphor. The charge at the anode is then collected and passed<br />

to the electronics through the base. As the number <strong>of</strong> visible-light photons entering the


12.4. CALIBRATION OF THE DETECTOR 157<br />

photomultiplier tube is proportional to the energy <strong>of</strong> the original γ being detected, the charge<br />

collected at the end <strong>of</strong> the tube is also proportional to the energy <strong>of</strong> that γ.<br />

When detecting the photons as a function <strong>of</strong> energy, one obtains a spectrum that is quite<br />

structured. As all the energy is deposited by a photoelectric event in the detector, one obtains<br />

a peak (called the “photopeak”) at the energy <strong>of</strong> the incident photon. Likewise, one obtains<br />

a peak at the energy <strong>of</strong> the incident photon after pair production. In that case, however,<br />

two more peaks are observed: a) when one <strong>of</strong> the annihilation 511 keV photons escapes<br />

the detector, at 511 keV below the full energy peak; and b) when both annihilation photons<br />

escape the detector, at 1022 keV below the full energy peak. The energy <strong>of</strong> the electron<br />

after a Compton scattering is continuous from zero energy up to a maximum, and thus it is<br />

reflected in the photon spectrum as a continuous distribution up to an edge (the Compton<br />

edge) which lies just below the photopeak.<br />

Pre-lab Question 12.1 Calculate the maximum energy given to the scattered electron as<br />

a result <strong>of</strong> Compton scattering <strong>of</strong> a photon energy E γ . To what scattering angle does this<br />

energy correspond? Calculate the maximum energy <strong>of</strong> the electron for the case <strong>of</strong> an incident<br />

662 keV photon.<br />

Pre-lab Question 12.2 Calculate the energy <strong>of</strong> a Compton scattered photon for an incident<br />

photon <strong>of</strong> energy E γ when the scattering angle is 180 ◦ . Is it possible only for the scattered<br />

photon to be detected?<br />

Pre-lab Question 12.3 Show that as E γ → ∞, the energy <strong>of</strong> photons scattered through<br />

180 ◦ approaches 255 keV, and is independent <strong>of</strong> E γ .<br />

12.4.2 The Multi-Channel Analyser<br />

The centre-piece <strong>of</strong> the electronics is the Multi-Channel Analyser (MCA). It may sometimes<br />

be a stand-alone piece <strong>of</strong> equipment but is more likely to be a card in a PC, as it is here.<br />

When the signal is produced as a result <strong>of</strong> the detection <strong>of</strong> a photon, it is amplified by the<br />

pre-amplifier/amplifier system to produce a positive (unipolar) or positive/negative (bipolar)<br />

pulse whose amplitude is proportional to the energy <strong>of</strong> the photon. The MCA accepts this<br />

pulse and performs a comparison <strong>of</strong> the voltage <strong>of</strong> the signal with an internal voltage stepping<br />

up from zero volts until agreement. The number <strong>of</strong> steps defines the channel and the counter<br />

in that channel is incremented by 1. The results <strong>of</strong> such counting are displayed on the screen<br />

for analysis. It is possible then to take an entire spectrum at once with such a system.<br />

All MCAs accept pulses to 10 V full-scale, and divide the scale into the channels defined<br />

by the Analogue-to-Digital-Converter (ADC). These are usually in 2 k steps; for our system<br />

there are 1024 channels, defining 10 mV channels.<br />

12.4.3 The reference sources<br />

To calibrate the detector, we need sources for which we know the energies <strong>of</strong> the γ rays<br />

produced. Those sources are listed in table 12.1.


158 CHAPTER 12. γ RAY DETECTION AND SPECTROSCOPY<br />

Table 12.1: Radioactive sources used for the calibration, and the energies <strong>of</strong> the photons<br />

produced.<br />

Source E γ (MeV)<br />

137 Cs 0.032<br />

0.662<br />

22 Na 0.511<br />

1.275<br />

60 Co 1.173<br />

1.332<br />

12.4.4 Gain <strong>of</strong> the detection system and MCA<br />

Starting with the detector, identify each piece <strong>of</strong> equipment and follow the wiring as per<br />

figure 12.2. The electronics, including the high voltage (HV) power supply, are modules<br />

contained in the NIM bin. Ensure that the wiring from the detector to the MCA is complete,<br />

and that the HV is connected to the detector. With the HV set to zero switch on the bin, the<br />

CRO and the computer (MCA). Slowly increase the voltage on the HV to 800 V.<br />

Place the 137 Cs source in front <strong>of</strong> the detector. Observe the signal produced by the preamplifier<br />

on the CRO to ensure that a signal is produced. Reconnect the amplifier to the preamplifier<br />

and observe the signals output from the amplifier.<br />

Question 12.1 Describe the pulses emerging from the amplifier and note any differences,<br />

giving possible reasons.<br />

Recall that the MCA accepts voltages between 0 and 10 V. The gain on the amplifier should<br />

be set such that the voltages <strong>of</strong> the pulses produced fall within this range. Note that the<br />

photon energies you may encounter in this experiment are no more than 2 MeV.<br />

Question 12.2 The highest energy photon emitted by the 137 Cs source is 0.662 MeV. Assuming<br />

a linear calibration <strong>of</strong> the MCA (1024 channels), at approximately which channel will<br />

the photopeak appear if we wish to measure energies <strong>of</strong> ∼ 2 MeV?<br />

Adjust the amplifier gain such that you affect this situation. Note the amplifier settings.<br />

The overall gain <strong>of</strong> the detection system is also affected by the HV applied to the photomultiplier<br />

tube.<br />

Pre-lab Question 12.4 Why does the high voltage affect the gain <strong>of</strong> the system?<br />

Investigate the effect <strong>of</strong> the HV supply on the gain by making very gradual changes to the<br />

applied voltage and observe the effect on the output voltages on the CRO.<br />

Question 12.3 What is the effect <strong>of</strong> small drifts in the applied HV on the overall spectrum?<br />

Question 12.4 What action may you take to minimise the effect <strong>of</strong> drift?


12.5. ENERGY RESOLUTION 159<br />

There is one more quantity to consider. When one particle is detected it triggers <strong>of</strong>f the<br />

electronics in sequence to the MCA to record the event. As this happens the electronics are<br />

working to process the event. Another particle entering the detector during this time will not<br />

be detected - the detector is dead for this time. Dead time is recorded at the bottom <strong>of</strong> the<br />

MCA screen. Generally, this will not be a major problem but you should keep a record <strong>of</strong> it.<br />

Dead time is discussed in detail in the β spectroscopy experiment.<br />

12.4.5 Measuring spectra<br />

The commands for operating the MCA s<strong>of</strong>tware are given in appendix 12.7. Reset the HV<br />

to 800 V and allow a few minutes for the power supply to stabilise. Collect a spectrum for<br />

around 10 min using the 137 Cs source. The collection time is nothing special, but it leads to<br />

a very important concept and a visiting question from the α scattering experiment:<br />

Question 12.5 What determines a suitable count time?<br />

Save the spectrum and print one copy for each <strong>of</strong> the partners in the group. Record the<br />

channel numbers <strong>of</strong> all <strong>of</strong> the features in the spectrum. Collect spectra also for the 22 Na and<br />

60 Co sources.<br />

12.4.6 Calibration<br />

Once you have identified the features in each spectrum, and especially the positions <strong>of</strong> the<br />

peaks and the errors therein, you may calibrate the detector. Using the known energies <strong>of</strong><br />

the photons in the three sources given in table 12.1 plot a graph <strong>of</strong> energy against channel<br />

number.<br />

Question 12.6 What is the relationship between energy and channel number? Use this to fit<br />

an appropriate function to the data and hence find the calibration <strong>of</strong> the detector.<br />

Use the calibration to determine the energies <strong>of</strong> the other features in the spectra: Compton<br />

edges and pair production peaks. Compare the measured energies with your expectations.<br />

12.5 Energy resolution<br />

The accuracy <strong>of</strong> the measurement <strong>of</strong> the spectrum is dictated by the energy resolution <strong>of</strong><br />

the system. That accuracy determines the amount <strong>of</strong> detail that may be identifiable. The<br />

resolution is defined as the width <strong>of</strong> the peaks in the spectrum, and is embodied in the full<br />

width half maximum (FWHM) - the width as measured at half the number <strong>of</strong> counts in the<br />

peak. The resolution is then<br />

Resolution = FWHM<br />

E γ<br />

. (12.2)


160 CHAPTER 12. γ RAY DETECTION AND SPECTROSCOPY<br />

Measure the FWHM <strong>of</strong> the photopeaks in your spectra and calculate the resolution for each<br />

<strong>of</strong> the six peaks, including errors as you go. Plot a function <strong>of</strong> resolution against E γ and<br />

comment. What function <strong>of</strong> the detection system causes the resolution observed? Think<br />

about it, and remember how the detector works to observe a photon.<br />

12.6 Unknown sources<br />

Now that the system is calibrated, and potential sources <strong>of</strong> error identified, you are in a<br />

position to use it to identify unknown sources. Collect the spectra <strong>of</strong> the unknown sources<br />

given to you and measure the energies <strong>of</strong> any features you identify in the spectra. Compare<br />

those energies to the ones listed in table 12.2. What are the sources?<br />

Note that you should use those features <strong>of</strong> the spectra which are readily identifiable.<br />

12.7 MCA commands<br />

Alt-F1 Start data acquisition.<br />

Alt-F2 Stop data acquisition.<br />

Alt-F5 Move data from the MCA card into the computer’s memory buffer.<br />

Alt-F, Alt-S Save the data stored in the memory buffer into a file. Only a spectrum stored in<br />

the memory buffer can be saved.<br />

Arrow keys Left and right keys move the cursor. Channel number <strong>of</strong> the cursor’s position<br />

is indicated at the bottom <strong>of</strong> the screen. Page-up and Page-dn are fast movements.<br />

Up arrow changes the vertical scale from logarithmic to linear; the down arrow key is<br />

the reverse.<br />

F3 Changes to expanded view. The expansion region is changed by the +/- key on the<br />

keypad.<br />

<strong>Bibliography</strong><br />

[1] G F Knoll. Radiation Detection and Measurement. Wiley, 3 rd edition, 2000.<br />

[2] W R Leo. Techniques for Nuclear and Particle <strong>Physics</strong> Experiments. Springer-Verlag,<br />

1987.


BIBLIOGRAPHY 161<br />

Table 12.2: Table <strong>of</strong> various γ sources with the energies in keV.<br />

Energy Source Energy Source Energy Source<br />

32 137 Ba 303 133 Ba 779 152 Eu<br />

35 125 Sb 344 152 Eu 812 56 Ni<br />

53 133 Ba 353 226 Ra 894 142 Ba<br />

53 226 Ra 356 133 Ba 898 88 Y<br />

77 142 Ba 364 142 Ba 949 142 Ba<br />

80 131 I 364 131 I 964 152 Eu<br />

81 133 Ba 381 381 Sb 1001 142 Ba<br />

97 75 Se 401 75 Se 1064 207 Bi<br />

121 75 Se 425 142 Ba 1078 142 Ba<br />

122 152 Eu 428 125 Sb 1086 152 Eu<br />

136 75 Se 463 125 Sb 1120 226 Ra<br />

176 125 Sb 511 e + − e − 1173 60 Co<br />

186 226 Ra 570 207 Bi 1204 142 Ba<br />

232 142 Ba 600 142 Ba 1238 226 Ra<br />

242 226 Ra 601 125 Sb 1277 22 Na<br />

245 152 Eu 607 125 Sb 1333 60 Co<br />

225 142 Ba 609 226 Ra 1378 226 Ra<br />

265 75 Se 636 125 Sb 1408 1408 Eu<br />

270 56 Ni 637 131 I 1562 56 Ni<br />

273 133 Ba 662 137 Cs 1764 226 Ra<br />

280 75 Se 723 131 I 1770 207 Bi<br />

284 131 I 750 56 Ni 1836 88 Y<br />

295 226 Ra 769 226 Ra


Chapter 13<br />

Energy loss <strong>of</strong> α particles through matter<br />

13.1 Abstract<br />

Nature has been both kind and unkind to us for giving us charged particles with which to<br />

perform experiments in our endeavour to understand it. Charged particles interact with matter<br />

through the Coulomb force with the greatest ease, making detection easy, but which also<br />

means that such particles interact in air. In order to perform experiments with charged particles<br />

in a controlled environment one needs to create vacuums.<br />

The experiment is thus tw<strong>of</strong>old: a) to understand how charged particles, in our case α particles,<br />

interact with matter; and b) how to create high vacuums.<br />

13.2 Introduction<br />

Much <strong>of</strong> what we know about the atomic and subatomic universes comes from studying<br />

the interactions <strong>of</strong> charged particles with the subject (call it “target”) under investigation.<br />

The problem is, that needs to be done in a controlled way such that the charged particles<br />

are interacting ONLY with that target. The very thing that allows us to easily study the<br />

results <strong>of</strong> the interactions and therefore the structure <strong>of</strong> the targets, the Coulomb force, is<br />

also responsible for making it incessantly difficult to perform experiments in air. It is not<br />

possible to control the flux under those conditions as:<br />

• the charged particles lose energy through inelastic collisions with atomic electrons;<br />

and,<br />

• the direction <strong>of</strong> the flux changes through elastic scattering with the surrounding nuclei.<br />

Other reactions can occur, <strong>of</strong> course, (Cherenkov radiations, nuclear reactions, bremsstrahlung 1 ).<br />

For our purposes, the most probable interaction effecting loss <strong>of</strong> energy is the interaction between<br />

the charged particles and the atomic electrons, causing either ionisation or atomic<br />

excitation.<br />

1 braking radiation<br />

163


164 CHAPTER 13. ENERGY LOSS OF α PARTICLES THROUGH MATTER<br />

Energy loss itself is embodied in the quantity dE/dx, the energy loss per unit distance (It<br />

is, by definition, negative.) Two calculations <strong>of</strong> this quantity exist: that <strong>of</strong> Bohr, derived<br />

classically, and that <strong>of</strong> Bethe and Bloch, which is derived from quantum mechanics and is<br />

the correct formulation. The classical formulation is sufficient for low energy α particles. It<br />

is:<br />

{ }<br />

dE<br />

dx = e 4 γ 2<br />

−4πz2 m e v N mv 3<br />

2 e ln<br />

ze 2¯ν , (13.1)<br />

where<br />

ze is the charge <strong>of</strong> the α particle;<br />

m e is the mass <strong>of</strong> the electron;<br />

m is the mass <strong>of</strong> the α;<br />

γ is the relativistic boost;<br />

v is the velocity <strong>of</strong> the α; and,<br />

ν is the average orbital frequency <strong>of</strong> the bound electron in the atom.<br />

Taking E α = mv 2 /2, we can subsume most <strong>of</strong> the constants (including the introduced mass<br />

<strong>of</strong> the α) into a single beast and obtain<br />

dE<br />

dx = −c 1<br />

E ln [c 2E]. (13.2)<br />

If the material is sufficiently thin that the energy loss is small compared to the total energy,<br />

the log factor is almost constant over the foil thickness and we can rewrite the equation as<br />

∆E<br />

T ≈<br />

k<br />

E m , (13.3)<br />

where ∆E is the energy lost over thickness <strong>of</strong> foil T ; E is the energy <strong>of</strong> the incident α<br />

particle; k and m are constants <strong>of</strong> the foil.<br />

Taking the log <strong>of</strong> both sides, one obtains<br />

log ∆E = log k + log T − m log E . (13.4)<br />

Thus, if we plot a curve <strong>of</strong> ∆E as a function <strong>of</strong> E on a log-log scale, we should get a straight<br />

line.<br />

13.3 Measuring foil thickness<br />

With a knowledge <strong>of</strong> energy loss <strong>of</strong> charged particles through a known thickness <strong>of</strong> foil, one<br />

may use equation 13.4 to determine the actual function, and from that determine thicknesses<br />

<strong>of</strong> other foils <strong>of</strong> the same material.<br />

To do this, we need to measure the energy loss as a function <strong>of</strong> energy through foils <strong>of</strong> known<br />

thickness. The apparatus for such a measurement is shown in figure 13.1. The geometry <strong>of</strong>


13.3. MEASURING FOIL THICKNESS 165<br />

O−ring seal<br />

Air intake<br />

valve<br />

Foil holder<br />

226<br />

Ra<br />

α<br />

Solid state<br />

detector<br />

To electronics<br />

Foil<br />

To vacuum pump<br />

Figure 13.1: Detection system for energy loss <strong>of</strong> α particles in foil.


166 CHAPTER 13. ENERGY LOSS OF α PARTICLES THROUGH MATTER<br />

SSD<br />

Pre−<br />

amp Amp MCA<br />

Bias<br />

CRO<br />

Figure 13.2: Circuit for analysing the detector signals.<br />

the system is such that all α particles detected must have passed through a thickness <strong>of</strong> foil.<br />

The detector itself is the same as that used in the Rutherford Scattering experiment. As<br />

with the Rutherford scattering, the experiment must be performed under high vacuum. The<br />

source is a sample <strong>of</strong> 226 Ra which produces α and β particles on its way to decay to 206 Pb,<br />

the spectrum <strong>of</strong> which is shown in Table 13.1.<br />

Table 13.1: Spectrum <strong>of</strong> α and β particles produced by the 226 Ra source. Energies are in<br />

keV.<br />

Particle Energy<br />

α 7687.1<br />

6002.6<br />

5489.7<br />

5305<br />

4748.5<br />

β 3280<br />

1161<br />

690<br />

170<br />

The signal from the detector is passed to the circuit shown in figure 13.2.<br />

13.3.1 Calibration <strong>of</strong> the system<br />

Identify all the components in figures 13.1 and 13.2 and set it up. Make sure that the bias<br />

voltage is <strong>of</strong>f when handling the detector in air, and ensure that for this part <strong>of</strong> the experiment<br />

there is no foil in the foil holder. Assemble the detecting chamber and, with the air inlet<br />

valve closed, pump down the chamber until the pressure is below at least 0.1 Torr. Once it<br />

has dropped below this pressure, slowly turn up the bias voltage to 70 V. NB! Do NOT adjust<br />

the bias voltage under pressures GREATER than 0.1 Torr.<br />

The MCA system is as with the other experiments, measuring voltages in the 0−10 V range,<br />

and is based on a 1024-channel ADC. The MCA commands are given in appendix 13.6.<br />

Given that we are measuring α particles with energies up to 7.8 MeV, adjust the gain <strong>of</strong><br />

the amplifier such that the highest energy α produces a pulse under the maximum voltage.<br />

Record your setting.<br />

Collect a spectrum with no foil in place. Ensure a long enough time for collection.


13.4. MAKING A THIN FOIL 167<br />

Question 13.1 What determines a suitable count time?<br />

Determine which peaks are due to detection <strong>of</strong> α particles and which are due to the β’s.<br />

Record the channel numbers <strong>of</strong> only the α peaks and use those to calibrate the system. Be<br />

sure to take note <strong>of</strong> the dead time (see β and γ experimental notes) and ensure that it is<br />

minimised.<br />

13.3.2 Measurement <strong>of</strong> known foils<br />

Take various energy spectra <strong>of</strong> α particles as they pass through known thicknesses <strong>of</strong> foils,<br />

using the foils supplied. Track any changes to the spectrum with foil thickness using a couple<br />

<strong>of</strong> representative peaks. There are two things for you to consider:<br />

1. Any changes in the positions <strong>of</strong> peaks;<br />

2. Any changes to the full-width-half-maximum (FWHM) <strong>of</strong> said peaks.<br />

Plot a curve <strong>of</strong> ∆E against E for a few points on a log-log scale and determine the energyloss<br />

function (equation 13.4) for your foils.<br />

Question 13.2 Discuss possible reasons why there may be changes to the FWHM with foil<br />

thickness. Can a point be reached where energy resolution becomes too poor for the recognition<br />

<strong>of</strong> α peaks?<br />

13.4 Making a thin foil<br />

As all experiments involving charged particles involve the use <strong>of</strong> vacuum systems, this part<br />

<strong>of</strong> the experiment requires the use <strong>of</strong> ultra-high vacuum pumps. The application is to make<br />

your own thin foil, <strong>of</strong> the same material used in the previous section, and use the results <strong>of</strong><br />

α energy loss to measure the thickness <strong>of</strong> the foil you make. This takes all <strong>of</strong> the second day<br />

- make sure you hit the ground running!<br />

13.4.1 Background reading<br />

The book by Roth, Vacuum Technology [2], is available from the part 2 <strong>of</strong>fice and all answers<br />

to the questions may be found there. Chapters 5, 6, and section 7.1, 7.2, address the<br />

techniques studied in this part <strong>of</strong> the experiment and you should take the time to read/skim,<br />

and familiarise yourself with, those sections. NB! Do NOT transcribe large chunks <strong>of</strong> the<br />

book into the prac: the knowledge gained should form part <strong>of</strong> the report as is relevant to the<br />

experiment being done!


168 CHAPTER 13. ENERGY LOSS OF α PARTICLES THROUGH MATTER<br />

Bell jar<br />

Baffle<br />

valve<br />

Roughing<br />

valve<br />

Air inlet<br />

Backing<br />

valve<br />

Diffusion<br />

pump<br />

To outside<br />

Rotary<br />

pump<br />

Figure 13.3: Diagram <strong>of</strong> the evaporation system.<br />

13.4.2 Foil manufacture<br />

The aluminium foil you will make will utilise the techniques <strong>of</strong> evaporation. The source<br />

material (Al) is placed on a metallic boat made <strong>of</strong> heat resistant material, such as titanium.<br />

That boat is heated by passing an electric current through it. The material sits within a<br />

depression in the boat. The depression is thinner than the surrounding material, creating a<br />

region <strong>of</strong> higher resistance relative to the rest <strong>of</strong> the circuit and so heats up the fastest. The<br />

source material is heated to boiling point at which time the resultant vapour expands in a<br />

hemisphere above the boat condensing on whatever surface it first strikes. The procedure is<br />

done under high vacuum to ensure that no oxidation <strong>of</strong> the material takes place.<br />

The evaporation system is shown in figure 13.3. The process takes place inside a glass<br />

bell jar which is evacuated by a system <strong>of</strong> a rotary pump and diffusion pump. Next to the<br />

bell jar is a high-current power supply consisting <strong>of</strong> variable (240 V, 10 A) to (3 V, 80 A)<br />

step down transformer connected to the boat clamps via 150 A cables. NB! The bell jar is<br />

very expensive and delicate. Only the demonstrator should handle it, and the perspex shield<br />

should always be placed around the jar when it is evacuated to guard against shrapnel in the<br />

unlikely event <strong>of</strong> an implosion.<br />

Take note <strong>of</strong> the rules for operating a vacuum system as given in appendix 13.5. Identify all


13.4. MAKING A THIN FOIL 169<br />

the components <strong>of</strong> the vacuum system and have the demonstrator remove the bell jar if not<br />

already done. The titanium boat will have been handled prior to use so it is necessary for<br />

you to clean it before placing it in situ. Clean the boat using isopropyl alcohol and lint-free<br />

tissues. (Remember those latex gloves!) After cleaning, carefully place the boat in its holder<br />

in the evaporation chamber, being careful not to touch it with bare hands.<br />

Pre-lab Question 13.1 Why is it necessary to keep the inside <strong>of</strong> a vacuum chamber clean?<br />

How can contamination be minimised?<br />

Pre-lab Question 13.2 What is outgassing?<br />

Place the 0.5 g Al slug into the boat, making sure it is clean. Clean the mylar film holder,<br />

again with isopropyl alcohol, and mount it in the clamp ∼ 7 cm above the boat in the chamber.<br />

Inspect the inside <strong>of</strong> the vacuum chamber for any contamination. If you find any, remove<br />

them with the alcohol before replacing the bell jar and perspex shield.<br />

Pre-lab Question 13.3 Note the rubber seals on the jar. Discuss the function <strong>of</strong> the vacuum<br />

seal and the main factors involved.<br />

Pre-lab Question 13.4 Briefly discuss the function <strong>of</strong> rotary and diffusion pumps and their<br />

operating vacuums.<br />

Pre-lab Question 13.5 How does a thermistor gauge work? Over what range <strong>of</strong> pressures<br />

can such thermal gauges operate?<br />

Pre-lab Question 13.6 What is the definition <strong>of</strong> a Torr?<br />

To reduce the pressure to below 10 −4 Torr, we use the diffusion pump. Before we can use the<br />

diffusion pump, we must evacuate the bell jar to a pressure <strong>of</strong> 10 −2 Torr. Close all <strong>of</strong> the air<br />

inlet valves, checking that both the baffle and backing valves are closed, and then open the<br />

roughing valve. Switch on the rotary pump and wait until the pressure is below 10 −2 Torr.<br />

Once reached, open the backing valve and switch on the diffusion pump heater.<br />

Once heated (about 15 minutes), close the roughing valve, open the baffle valve (and worry<br />

about why the roughing valve should be closed FIRST), and observe the change in pressure.<br />

Comment on any rapid changes during this step.<br />

When the pressure is below 10 −5 Torr you can start the evaporation. Turn on the power<br />

supply and raise the current until the boat glows bright red (∼ 50 − 70 A). Once the Al has<br />

melted, slowly increase the current until the boat glows white hot and you can the bell jar<br />

being coated with Al. Continue evaporation until you can see that all <strong>of</strong> the Al source has<br />

evaporated (roughly 2 minutes).<br />

You have made the foil and can now begin the process <strong>of</strong> bringing the bell jar back up to air.<br />

First, close the baffle valve and then switch <strong>of</strong>f the diffusion pump heater. The bell jar is now<br />

isolated from the pumps so you can use the air inlet closest to the jar to bring the chamber<br />

back up to air. When the diffusion pump has cooled (∼ 20 minutes), close the backing valve,<br />

turn <strong>of</strong>f the rotary pump, and bring the vacuum side <strong>of</strong> the rotary pump to air using the right<br />

most air inlet valve.


170 CHAPTER 13. ENERGY LOSS OF α PARTICLES THROUGH MATTER<br />

13.4.3 Thickness measurement<br />

You have a new foil. Now measure its thickness using the energy loss information you<br />

measured for aluminium foils. Ensure that the mylar backing is facing the 226 Ra α source.<br />

Compare your measured value with what you may expect if all <strong>of</strong> the slug had been evaporated<br />

(Al density is 2.7 g/cm 3 ) during which it uniformly coats a hemisphere <strong>of</strong> radius equal<br />

to the boat-mylar distance.<br />

Note that the α particles have also passed through mylar (a compound <strong>of</strong> C and H). If time<br />

permits, take measurements to correct for that absorption using the second mylar-only foil<br />

available for use. Would you expect there to be a large correction?<br />

13.5 Rules for operating a high vacuum system 2<br />

• Always vent a mechanical pump to air when the power is <strong>of</strong>f. The presence <strong>of</strong> a<br />

residual vacuum in the pump causes oil to be drawn into the pump contaminating its<br />

chambers.<br />

• Do not permit a mechanical pump to exhaust a high vacuum system below a pressure<br />

<strong>of</strong> about 10 −2 Torr, unless the pump is separated from the high-vacuum chamber with<br />

a trap stopping the oil vapour from entering the chamber.<br />

• Always keep the inside <strong>of</strong> a high vacuum system clean. In particular, grease from<br />

fingers can result in considerable contamination and gloves should ALWAYS be used<br />

when handling anything destined for the inside <strong>of</strong> high vacuum equipment.<br />

• Do not run a mechanical pump at high vacuum for long periods <strong>of</strong> time. The motors<br />

are not generally designed for long term use with the additional load, and the pump<br />

will eject much oil, in addition to gas, into the system.<br />

• Diffusion pumps should be cooled to a safe intermediate temperature before being<br />

vented to air. Venting at too high a temperature results in oxidation <strong>of</strong> the pump fluid<br />

and excessive carryover <strong>of</strong> the fluid into the mechanical pump.<br />

• Always check that the cooling water supply to diffusion and turbomolecular pumps is<br />

turned on prior to operation. It is also a good idea to have thermal protection to guard<br />

against interruption to the cooling water supply.<br />

• In liquid-N 2 trapped systems the trap should be cool enough to condense the diffusion<br />

pump oil prior to turning on the diffusion pump. Maximum backstreaming <strong>of</strong> pump<br />

oil occurs during startup and shutdown <strong>of</strong> the diffusion pump.<br />

• Do not vent a liquid N 2 trap to air whilst cold.<br />

• When the chamber is vented to atmosphere it is advisable to use a dry, inert, gas such as<br />

nitrogen to minimise moisture adsorption on the surface <strong>of</strong> the vacuum system. Never<br />

vent from the fore-line <strong>of</strong> the diffusion pump.<br />

2 Not all <strong>of</strong> the rules apply here; for example, we do not have a liquid N 2 trap on our diffusion pump.<br />

However, we mention it for the sake <strong>of</strong> reference.


13.6. MCA COMMANDS 171<br />

• Ionisation gauges, in particular hot cathode gauges, should not be turned on until the<br />

pressure is below 10 −3 Torr.<br />

13.6 MCA commands<br />

Alt-F1 Start data acquisition.<br />

Alt-F2 Stop data acquisition.<br />

Alt-F5 Move data from the MCA card into the computer’s memory buffer.<br />

Alt-F, Alt-S Save the data stored in the memory buffer into a file. Only a spectrum stored in<br />

the memory buffer can be saved.<br />

Arrow key Left and right keys move the cursor. Channel number <strong>of</strong> the cursor’s position is<br />

indicated at the bottom <strong>of</strong> the screen. Page-up and Page-dn are fast movements. Up<br />

arrow changes the vertical scale from logarithmic to linear; the down arrow key is the<br />

reverse.<br />

F3 Changes to expanded view. The expansion region is changed by the +/- key on the<br />

keypad.<br />

<strong>Bibliography</strong><br />

[1] W R Leo. Techniques for Nuclear and Particle <strong>Physics</strong> Experiments. Springer-Verlag,<br />

1987.<br />

[2] A Roth. Vacuum Technology. Elsevier Science, 1990.


Chapter 14<br />

Errors<br />

14.1 Introduction<br />

This chapter has been constructed to show you that error analysis really is quite simple and<br />

is <strong>of</strong> great importance to your report and your development as an experimental physicist.<br />

Topics covered include sources and types <strong>of</strong> error, how to calculate and combine errors and<br />

how to display your final result. Errors are <strong>of</strong>ten thought <strong>of</strong> as troublesome, but given the<br />

appropriate background knowledge they are really quite easy. So please take the time to read<br />

through the following guide and master those errors once and for all.<br />

When scientists present the results <strong>of</strong> experimental measurements, they should almost always<br />

specify a possible error associated with the quoted results. It is very poor experimental<br />

technique to say that, “this is my best value, but . . . I have no idea, nor do I care about how<br />

accurate it is.” In first year laboratory you were introduced to the concept <strong>of</strong> errors. In<br />

second year laboratory we not only expect you to think about and mention the errors, but<br />

you are required to understand the errors, estimate their magnitude using some relatively<br />

simple formulae, reduce them by developing good experimental technique, and discuss them<br />

at length throughout your report. The failure to consider, estimate, reduce or discuss the<br />

error associated with any quantity you obtain is very poor experimental technique. Hence,<br />

you should always state your final result plus a confidence limit.<br />

x + △x units<br />

Your final result is your best estimate for the value being investigated, whereas the confidence<br />

limits are the upper and lower limits between which you declare that you confidently expect<br />

the true value to lie.<br />

The confidence limit or ‘error’ conveys very significant information about the experiment and<br />

the result obtained, thus its consideration is an indication <strong>of</strong> a sound experimental technique.<br />

We shall now consider the different types <strong>of</strong> errors, how to estimate and combine them and<br />

how to present them appropriately.<br />

173


174 CHAPTER 14. ERRORS<br />

14.2 Sources <strong>of</strong> error in experimental work<br />

14.2.1 Mistakes (!!)<br />

First, let us dispose with human errors that are just silly blunders; misreading a scale, recording<br />

a number incorrectly or a making a mistake in calculation. These are not really errors in<br />

the scientific sense at all. The best advice we can give is - don’t make them!<br />

14.2.2 Random errors<br />

Random errors become evident when one takes repeated readings <strong>of</strong> a physical quantity and<br />

a small spread <strong>of</strong> readings is obtained, with readings equally likely to be in either direction<br />

from the average. The cause could be fluctuations in experimental conditions, unavoidable<br />

instrument variations, or your own unconscious subjective bias in the setting or reading <strong>of</strong><br />

instruments.<br />

Random errors can be investigated by taking repeated measurements.<br />

14.2.2.1 Repeated readings<br />

If you need to know the value <strong>of</strong> a quantity as accurately as possible, it is obvious that you<br />

need to check the result a few times. Repeating the measurement allows you to calculate the<br />

best value for the quantity being investigated and investigate the magnitude <strong>of</strong> the random<br />

errors by considering the statistical variation in the data.<br />

The best estimate <strong>of</strong> the true value <strong>of</strong> the result is the arithmetic mean <strong>of</strong> the measured result,<br />

〈X〉 = 1 N<br />

∑<br />

X i . (14.1)<br />

i<br />

An estimate in the uncertainty <strong>of</strong> the result comes from the standard deviation σ <strong>of</strong> the data,<br />

σ 2 =<br />

∑i (X i − 〈X〉) 2<br />

. (14.2)<br />

(N − 1)<br />

where N is the number <strong>of</strong> points in the data set. For errors with a normal distribution we<br />

expect 68% (= 1 − 1/e) <strong>of</strong> the recorded measurements to be within one standard deviation<br />

<strong>of</strong> the mean [2].<br />

14.2.3 Reading errors<br />

Both the accuracy and precision <strong>of</strong> any real measuring device is limited 1 . In first year laboratories<br />

you learnt that the precision is limited to perhaps to a quarter <strong>of</strong> the smallest scale<br />

1 The distinction between accuracy and precision can be illustrated by making an analogy with target archery.<br />

Precision is how close the arrows are together; accuracy is how close they are to the bullseye.


14.2. SOURCES OF ERROR IN EXPERIMENTAL WORK 175<br />

on the instrument (ruler). The instruments used to measure quantities include rulers with<br />

Vernier scales, calipers and digital voltmeters. Each <strong>of</strong> these improves the precision with<br />

which values can be obtained; however, a reading error is still always present.<br />

14.2.3.1 Parallax<br />

A quick little reminder <strong>of</strong> a thing called parallax. It is important to align yourself with the<br />

instrument appropriately if you wish your results to be accurate.<br />

Figure 14.1: Parallax errors arising from the measurement <strong>of</strong> the length <strong>of</strong> an object. The<br />

apparent length, r 2 , in this case is shorter than the real length, r 1 . The apparent length would<br />

be the real length if the measurement was taken with the ruler in contact with the object.<br />

14.2.3.2 Limit <strong>of</strong> accuracy<br />

Figure 14.2: The first scale can be read to the nearest 0.5 mm, while the second scale can be<br />

read to the nearest 0.025 mm.<br />

Limit <strong>of</strong> accuracy = a quarter <strong>of</strong> the smallest division<br />

The accuracy <strong>of</strong> the result will depend upon the accuracy <strong>of</strong> the measuring device you use. In<br />

figure 14.2 we consider the length <strong>of</strong> a block <strong>of</strong> material. It is quite obvious that the accuracy


176 CHAPTER 14. ERRORS<br />

<strong>of</strong> our result is greater in the second case. In the Second Year laboratories the accuracy <strong>of</strong><br />

our result should be down to a quarter <strong>of</strong> the smallest division. This applies to both rulers<br />

and needle metres. If the results depend greatly upon the accuracy <strong>of</strong> this result you will find<br />

a Vernier scale or digital meter (both discussed below) more appropriate.<br />

14.2.3.3 The Vernier scale<br />

A Vernier scale 2 is used whenever one needs to make a measurement <strong>of</strong> a distance or angle<br />

to a greater accuracy than that obtainable through direct visual reading <strong>of</strong> a linear scale.<br />

The Vernier uses the linear reading scale <strong>of</strong> the measuring apparatus (i.e. a ruler) along with<br />

another scale which is scaled by a factor <strong>of</strong> 9:10 compared to the linear scale.<br />

Reading a Vernier scale is best seen with an example (figure 14.3).<br />

In order to read this scale:<br />

Figure 14.3: Vernier scale<br />

• Read the linear part <strong>of</strong> the measurement, in this case, the ruler marking imediately to<br />

the left <strong>of</strong> the ‘zero’ <strong>of</strong> the Vernier. This is the first approximation to the measurement<br />

and yields a result <strong>of</strong> 2.5 mm.<br />

• Next, determine which marking on the Vernier scale most nearly lines up with the<br />

markings on the linear scale. Do this by scanning your eyes across the scale and<br />

judging whether the Vernier scale is to the left or the right <strong>of</strong> the linear scale. When<br />

we cannot say left or right, we know that we have the alignment. In this case the 2.5<br />

on the Vernier. We add this to the first approximation to find the reading. So the result<br />

is 2.75 mm.<br />

• Finally, the smallest scale on the Vernier scale is 0.005 mm. This is the uncertainty in<br />

the measurement. Thus our final result would be<br />

(2.750 ± 0.005) mm<br />

For further details on reading a Vernier, see section 8.4.4, “Reading the micrometer”.<br />

2 Invented by the French mathematician and inventor Pierre Vernier (1850-1637).


14.2. SOURCES OF ERROR IN EXPERIMENTAL WORK 177<br />

14.2.3.4 Digital voltmeter<br />

If you are using a digital voltmeter to record a voltage, it might be reading 1.004 volts, and<br />

the limit <strong>of</strong> the reading is 0.001 volts (which is the smallest difference between readings<br />

which the device can resolve). We would then be tempted to claim the result to be (1.004 ±<br />

.001)V. Which would seem reasonable. However the accuracy <strong>of</strong> the digital metres is less<br />

than the number <strong>of</strong> digits displayed would suggest, typically the accuracy is 1 digit plus some<br />

fraction <strong>of</strong> the reading (typically 0.3%). In this example the error is calculated as follows:<br />

△V = 3 × V + limit <strong>of</strong> reading<br />

1000<br />

e.g. △V = 3/1000 × 1.004 + 0.001 = 0.004012 ∴ error = 0.004.<br />

Thus the final result plus confidence limit is (1.004 ± 0.004) V.<br />

14.2.4 Systematic errors<br />

It is fairly easy to assess the size <strong>of</strong> the random errors that we have been discussing. Averaging<br />

a number <strong>of</strong> readings tends to cancel out the effects <strong>of</strong> such errors, and the spread <strong>of</strong><br />

the readings allows us to estimate the size <strong>of</strong> the random errors involved. There is however<br />

another type <strong>of</strong> error - known as systematic errors - the effects <strong>of</strong> which are much more<br />

difficult to assess. Systematic errors are errors that systematically shift the measurements in<br />

one direction away from the true value. Thus repeated readings do not show up the presence<br />

<strong>of</strong> systematic errors, and no amount <strong>of</strong> averaging will reduce their effects. Both systematic<br />

and random errors are <strong>of</strong>ten present in the same measurement, and the different effects they<br />

have are contrasted in figure 14.4.<br />

There are no rules for eliminating systematic errors, or even detecting them, but the following<br />

points may be <strong>of</strong> some assistance.<br />

• Apparatus<br />

– If at all feasible, meters should be checked against a standard, or at least, against<br />

another and hopefully better meter. Zero settings should be checked and adjusted<br />

if necessary.<br />

– Given values should be treated with suspicion, and checked if possible; especially<br />

resistors that can vary 10% or more from their nominal value.<br />

– Instructions for use <strong>of</strong> the apparatus should be read and followed. Make sure<br />

you know how to read the scales on the instruments. Identifying features <strong>of</strong> the<br />

apparatus should be recorded.<br />

– If possible, vary the conditions at measurement slightly to avoid systematic errors.<br />

Use different rulers or section <strong>of</strong> a ruler to measure lengths or swap rulers if<br />

possible. The variation <strong>of</strong> conditions should be used in a discussion <strong>of</strong> systematic<br />

errors in your report.<br />

• Observations


178 CHAPTER 14. ERRORS<br />

Figure 14.4: Schematic diagram showing the different types <strong>of</strong> errors and how they are<br />

manifest in the results.<br />

– Measurements should be repeated by different observers to detect experimenter<br />

bias.<br />

– Corrections for instrument bias should be made if necessary (e.g. incorrect zero<br />

setting on a metre).<br />

– Any necessary precautions (warm up times, step to avoid backlash in screw transports,<br />

temperature controls etc.) should be observed.<br />

• Analysis<br />

– The experiment should be carefully analysed for likely sources <strong>of</strong> error and steps<br />

should be taken to minimise these.<br />

– Theoretical arguments used in the derivation <strong>of</strong> relations for indirectly measured<br />

quantities should be checked for assumptions, and the assumptions checked for<br />

validity in the particular experiment.<br />

– Results should be checked for feasibility, and silly results 3 recalculated. If the<br />

results are still unreasonable then measurements must be taken again.<br />

3 Note we say recalculated, not discarded - sometimes the “silly” results contain a lot <strong>of</strong> physics. Be careful<br />

- if you torture the data enough, it will confess to anything!


14.3. GRAPHS AND ERRORS 179<br />

14.3 Graphs and errors<br />

14.3.1 Error bars<br />

If estimates <strong>of</strong> uncertainties in measured values are available then these should be indicated<br />

on the plot in the form <strong>of</strong> error bars. These are bars passing through the estimated point<br />

running from the lower limit to the upper limit <strong>of</strong> the range covered by the uncertainties.<br />

Thus if (〈x〉, 〈y〉) is the estimated point and △x and △y are the uncertainties then the error<br />

bar in the x direction runs from 〈x〉 − △x to 〈x〉 + △x, and the bar in the y directions runs<br />

from 〈y〉 − △y to 〈y〉 + △y as shown in figure 14.5.<br />

One may choose error bars with a magnitude <strong>of</strong> one or two standard deviations, remember<br />

to indicate the chosen convention in the description <strong>of</strong> the plot. In many practical cases the<br />

uncertainty <strong>of</strong> one <strong>of</strong> the quantities (usually the independent variable <strong>of</strong>ten placed on the<br />

x-axis) is much less than that <strong>of</strong> the other. In such cases the smaller error bar is <strong>of</strong>ten omitted<br />

and a note made in the report. If more than one data set is plotted on the same graph, a<br />

different coloured pen or different symbols should be used.<br />

(,+ ∆y)<br />

(− ∆x,)<br />

(,)<br />

(+ ∆x, )<br />

(, − ∆y)<br />

Figure 14.5: Error bars<br />

14.3.2 Estimates <strong>of</strong> gradients and uncertainties<br />

In many experiments the result is obtained from the gradient <strong>of</strong> a straight line. The straight<br />

line is usually a line <strong>of</strong> best fit to a set <strong>of</strong> experimental points. However, random errors will<br />

most probably cause the plotted points not to lie precisely on any one straight line, although<br />

the trend <strong>of</strong> the data may be clear.<br />

There are thus two questions to be answered:<br />

• What is the best estimate for the gradient <strong>of</strong> the line?<br />

• What is the uncertainty in the estimated gradient?<br />

There are a number <strong>of</strong> ways to estimate the gradient <strong>of</strong> a straight line from a plot <strong>of</strong> experimental<br />

points. We shall consider two such methods; best fit by eye, and least squares<br />

fitting.


180 CHAPTER 14. ERRORS<br />

14.3.2.1 Best fit by eye<br />

Figure 14.6: Line <strong>of</strong> best fit by eye.<br />

In many cases it is sufficient to draw the single straight line which seems to lie closet to the<br />

largest number <strong>of</strong> experimental points.<br />

If error bars <strong>of</strong> length one standard deviation have been drawn, then such a line should pass<br />

through at least 2/3 <strong>of</strong> these error bars. If the error bars are two standard deviations long,<br />

the line should pass through 90% <strong>of</strong> the bars.<br />

The gradient can then be calculated using “rise over run” and the equation <strong>of</strong> a straight line:<br />

y = mx + c<br />

m = rise<br />

run = y 2 − y 1<br />

x 2 − x 1<br />

The uncertainty in the gradient can be calculated using the ‘maximum and minimum gradient<br />

approach’. Two lines are drawn, a steepest and most shallow line that fit the data, (with fit<br />

meaning that they pass through 67% or 90% <strong>of</strong> the error bars for one and two standard<br />

deviation error bars respectively). The gradient is calculated for each line and the average <strong>of</strong><br />

these is used as the estimate <strong>of</strong> the gradient.<br />

gradient = 1 2<br />

(steepest gradient + shallow gradient)<br />

The uncertainty can be taken as<br />

gradient = 1 2<br />

(steepest gradient - shallow gradient)<br />

14.3.2.2 Least squares fit<br />

This technique is one <strong>of</strong> the most accurate techniques for obtaining the gradient <strong>of</strong> a line <strong>of</strong><br />

best fit. Excel uses this technique to find the line <strong>of</strong> best fit in its curve fit algorithm. This


14.4. EXCEL 181<br />

technique was developed by Gauss and assumes that all <strong>of</strong> the error is concentrated in the y<br />

co-ordinate, and asks<br />

“For a line <strong>of</strong> the form y = mx + c, what values <strong>of</strong> m and c yield the smallest<br />

value for the sum <strong>of</strong> errors (differences between the point it predicts and the<br />

experimental point) squared?”<br />

This technique minimises the total distance between the points on a line <strong>of</strong> best fit and the<br />

experimental point to obtain the best fit.<br />

14.4 Excel<br />

Excel is a powerful and elegant s<strong>of</strong>tware program used with the 2 nd year physics laboratories<br />

within the school <strong>of</strong> physics. Excel, as you can see in the tutorial (chapter 15), enables you<br />

to take a data set and fit an equation to it.<br />

Figure 14.7 is obtained in the tutorial. In this experiment, particles with particular energy<br />

are binned in a particular ADC channel number. Excel is used to find out the relationship.<br />

Clearly this is roughly linear so we guess that the equation to be used in the transform is <strong>of</strong><br />

the form y = mx + c.<br />

Figure 14.7: Line <strong>of</strong> best fit obtained using Excel in the tutorial, chapter 15.<br />

14.5 Combining errors<br />

In the majority <strong>of</strong> experiments there is more than one source <strong>of</strong> error. For instance, several<br />

random and systematic errors may be present in measurements <strong>of</strong> a single quantity. Further-


182 CHAPTER 14. ERRORS<br />

more, it is likely that a number <strong>of</strong> measurements <strong>of</strong> different quantities (each <strong>of</strong> which has<br />

an error associated with it) will have to be combined to calculate the required results. Thus<br />

it is important to know how errors are combined to get the overall error in an experiment.<br />

The basis <strong>of</strong> this combination is the following formula, which really makes a lot <strong>of</strong> sense.<br />

Suppose that your final result Z is a function <strong>of</strong> some experimentally determined quantities<br />

(x,y,...). It seems reasonable that if x and y have small errors in them, then the final result,<br />

being a function <strong>of</strong> those variables will also have some error in it. In the mid 1600’s Sir<br />

Isaac Newton developed calculus (and partial derivatives) to consider such variations. The<br />

general formulae for the error in Z due to experimental uncertainties (△x, △y,...) in the<br />

parameters is<br />

(△Z) 2 ≈<br />

( ) 2 δZ<br />

(△x) 2 +<br />

δx<br />

( ) 2 δZ<br />

(△y) 2 + ... (14.3)<br />

δy<br />

The squared terms in this function arise as a result <strong>of</strong> statistical theory, which is beyond the<br />

scope <strong>of</strong> these notes see [2] for more detail.<br />

Every confidence limit we quote can be calculated using the parameters in equation<br />

14.3. The errors △x, △y,... can be random errors such as the standard<br />

deviation from repeated measurements, reading errors from a Vernier scale, tolerance<br />

limits from a digital voltmeter or uncertainties in the gradient <strong>of</strong> a graph.<br />

All these ‘errors’ combine together to give an error in the final result.<br />

The example below illustrates the use <strong>of</strong> this equation and calculation <strong>of</strong> the final error.<br />

14.5.0.3 Worked example<br />

Consider the use <strong>of</strong> a diffraction grating to determine the wavelength <strong>of</strong> an unknown emission<br />

line. The formula is:<br />

Thus<br />

mλ = d sin θ. (14.4)<br />

λ = d sin θ<br />

m . (14.5)<br />

We can now work out the errors. We are trying to find out, if there are small errors (standard<br />

deviation, reading errors . . . ) in the variables m, d and θ, what the resultant error in λ is. We<br />

calculate the partial derivatives <strong>of</strong> λ with respect to d, m and θ and then use equation 14.3)<br />

δλ<br />

δd = sin θ<br />

m ,<br />

δλ<br />

δm<br />

=<br />

−d sin θ<br />

m 2 ,<br />

δλ<br />

δθ = d cos θ<br />

m<br />

(14.6)<br />

Thus


14.6. SIGNIFICANT FIGURES AND DATA/ERROR PRESENTATION 183<br />

(△λ) 2 = sin2 θ<br />

m 2 (△d)2 + d2 sin 2 θ<br />

(△m) 2 + d2 cos 2 θ<br />

( π<br />

) 2<br />

(△q) 2 (14.7)<br />

m 4 m 2 180<br />

The last term was converted into radians. This allows you to enter the angle in degrees.<br />

Alternatively you could not include the (π/180) 2 term and enter the angle in radians. If we<br />

are given the quantities<br />

d = 1710 ± 2 nm,<br />

m = 1,<br />

q = 17.30 ± 0.01 degrees,<br />

note that m is a DISCRETE variable and so has no associated error with it. When all the<br />

terms are evaluated we get λ = 508.511 ± 0.6595 nm = (508.5 ± 0.7) nm. Note also that we<br />

only quote the error to one significant figure. This leads us to a discussion about. . .<br />

14.6 Significant figures and data/error presentation<br />

The correct presentation <strong>of</strong> the final result from the previous problem is 508.5 ± 0.7 nm.<br />

Hence we are confident that our result is between 507.8 nm and 509.2 nm. Please take note<br />

<strong>of</strong> the following:<br />

• Results can be rounded up or down depending upon the magnitude <strong>of</strong> the error. Ask<br />

your demonstrator if you are unsure.<br />

• The ‘best guess’ can only be as accurate as the error. There is no point is quoting<br />

508.5110 ± 0.6. The 0.0110 is not important.<br />

• Usually only the first significant figure in the error is <strong>of</strong> any use. Stating the wavelength<br />

is 508.5 ± 0.7 will suffice. Clearly giving details <strong>of</strong> the error to the n th degree is not<br />

required. i.e. ± 0.6595.<br />

• Displaying the result as 5.085×10 2 ±6.595×10 −1 nm indicates a lack <strong>of</strong> appreciation<br />

for what the result is actually telling us. This error presentation method is far from<br />

elegant and indicates that we can calculate the error but do not really know what the<br />

important points to take from the calculation are. The form <strong>of</strong> (5.085 ± 0.007) ×10 2<br />

nm would be more appropriate.<br />

Calculating an error is only half the exercise. Once the error has been calculated<br />

we must then discuss not only the magnitude <strong>of</strong> the error and its significance but<br />

sources <strong>of</strong> error and what was done to minimise them. As we said earlier, a best<br />

result without an error is next to useless, but an error calculation without an<br />

explanation <strong>of</strong> what it means is almost as pointless.


184 CHAPTER 14. ERRORS<br />

14.7 General expressions<br />

Our final result Z is a function <strong>of</strong> experimental parameters (x,y,...). These formulae reduce<br />

to a simple form in many common cases. We encourage everyone to derive the following so<br />

we can appreciate that the following formulae are based upon the idea that if the variables x<br />

and y differ then our final result Z(x,y) will differ.<br />

14.7.1 General expression<br />

If the result Z depends upon (x,y,...), the corresponding error is calculated as follows:<br />

(△Z) 2 ≈<br />

( ) 2 dZ<br />

(△x) 2 +<br />

dx<br />

( ) 2 dZ<br />

(△y) 2 + ... (14.8)<br />

dy<br />

14.7.2 Sums and differences<br />

If Z = x + y + ... or Z = x − y − ...<br />

then<br />

(△Z) 2 ≈ (△x) 2 + (△y) 2 + ... (14.9)<br />

14.7.3 Products and quotients<br />

If Z = x × y × ... or Z = x ÷ y ÷ ...<br />

then<br />

( ) 2 △Z<br />

=<br />

〈Z〉<br />

( ) 2 △x<br />

+<br />

〈x〉<br />

( ) 2 △y<br />

+ ... (14.10)<br />

〈y〉<br />

14.7.4 Powers and functions<br />

If Z = x N<br />

then<br />

( ) ( )<br />

△Z △x<br />

= |N|<br />

〈Z〉 〈x〉<br />

(14.11)<br />

14.7.5 Simple and trigonometric functions<br />

If Z = f(x)<br />

then


14.8. CONCLUDING COMMENTS 185<br />

( ) △Z<br />

=<br />

〈Z〉<br />

df<br />

∣dx∣<br />

( ) △x<br />

〈x〉<br />

(14.12)<br />

Note: this last category includes all the trigonometric functions. Substituted angles must be<br />

in radians unless we have included the (π/180) 2 term in the error.<br />

A few <strong>of</strong> the errors to be calculated in the second year laboratories will require a combination<br />

<strong>of</strong> the above formulae. If you wish to use these formulae, then you must go through your<br />

specific case step by step. Alternatively you could use equation 14.3.<br />

14.8 Concluding comments<br />

We have looked at the source and types <strong>of</strong> errors, and how to record, reduce and calculate<br />

them. Error analysis is neither mysterious nor tricky 4 . As we have mentioned before, we<br />

must always state our final result plus confidence limit with correct accuracy and significant<br />

figures.<br />

It is a bad look to obtain a value without considering a confidence limit. The confidence limit<br />

or ‘error’ conveys very significant information about the experiment and the result obtained.<br />

Once the error has been calculated it should be displayed appropriately and finally discussed<br />

in depth.<br />

<strong>Bibliography</strong><br />

[1] L Kirkup. Experimental methods. Wiley, 1994.<br />

[2] P Bevington and D K Robinson. Data Reduction and Error Analysis for the Physical<br />

Sciences. McGraw-Hill Science/Engineering/Math, 3 rd edition, 2002.<br />

4 Although it can be very tedious.


Chapter 15<br />

Excel tutorial<br />

Welcome to the 2 nd year labs. This semester you will encounter a number <strong>of</strong> new concepts<br />

and experiments that will enable you to improve your experimental technique. In addition to<br />

becoming confident with your ability as a experimental physicist, you will hopefully develop<br />

an appreciation <strong>of</strong> the physical significance <strong>of</strong> the results you are taking, the associated errors<br />

and how to analyse these results.<br />

To this ends, we shall be using Excel. Though you should have already used Excel in Part I<br />

labs, in this tute you shall be introduced to the more powerful aspects <strong>of</strong> this program.<br />

15.1 Simple example, energy vs. ADC number<br />

Let us enter the data from table 15.1 and create a graph <strong>of</strong> it. We can then create a linear fit<br />

for this data, and plot this over the top <strong>of</strong> the experimental data to observe the accuracy <strong>of</strong><br />

the fit. Finally we will transform this data to obtain another quantity.<br />

15.1.1 Entering data into the worksheet<br />

After entering Windows, open Excel. A data worksheet will then open - enter the data<br />

contained in table 15.1 into two columns in the worksheet. Save your worksheet (File →<br />

Save As → yourFilename.xls).<br />

15.1.2 Making the plot<br />

We shall now fit the data. Select the two columns (including titles) and choose Insert →<br />

Chart, or click the graph icon on the toolbar. This will begin a chart wizard - your first<br />

choice is almost always XY (Scatter) as graph type (the 5 th choice in the list), and the next<br />

screen ensures you have the correct data in the correct arrangement (does it look like you<br />

expect?). Now, enter the chart title, and axes labels. The final screen asks you if you would<br />

prefer to place your new chart as a “A new sheet” or “As object in”. The latter choice is<br />

187


188 CHAPTER 15. EXCEL TUTORIAL<br />

ADC Number Energy (MeV)<br />

313 18.207<br />

399 19.186<br />

452 20.653<br />

538 21.632<br />

574 23.100<br />

611 24.155<br />

613 24.079<br />

713 26.602<br />

814 29.048<br />

907 31.495<br />

1069 33.941<br />

1130 36.387<br />

1227 38.833<br />

1347 41.279<br />

1430 43.725<br />

1509 46.171<br />

1600 48.617<br />

Table 15.1: Data.<br />

pre-selected, however the former, I find, always makes your program neater (try both, and<br />

see how the latter choice may cover your data, and generally gets in the way).<br />

15.1.3 Fitting the data<br />

To fit the plot, choose Chart → Fit the trendline. Within this window, there are two tabs,<br />

Type (which in this case is linear) and Options (please select both Display equation on chart,<br />

and Display R-squared value 1 on chart. I found that E = 0.0239ADC + 9.5876, with<br />

r 2 = 0.9969. Note how the chart is linked in a ‘live’ fashion to the data - observe this by<br />

changing one point, and checking to see how the point on the graph changes.<br />

15.1.4 Absolute referencing<br />

When formulas are used in Excel, it applies special rules to copy these formulas around the<br />

worksheet. For example, try typing =B1*A4 into cell C1. Both B1 and A4 can be entered<br />

using the keyboard, or by using the arrow keys or clicking to highlight the necessary cells,<br />

after the equals sign has been entered. Now copy cell C1 (Ctrl+c or Edit → Copy) and paste<br />

it into cell D1 (Ctrl+v or Edit → Paste). Note how the formula is no longer what it was<br />

originally, but instead has changed to =C1*B4. This is because Excel reads the formula not<br />

as “B1*A4”, but more as “multiply the cell one to the left with the cell two to the left and<br />

three down”. The application <strong>of</strong> dollar signs to formulae is summarised in table 15.2.<br />

The addition <strong>of</strong> dollar signs ($) into the formula will change the way Excel views the formula.<br />

1 The R 2 value gives the quality <strong>of</strong> the fit, it is called the (linear) correlation coefficient. If R 2 is 1, the fit is<br />

perfect, if R 2 is 0, there is absolutely no correlation. For more detail, see [1], section 11.2, page 198 - 201.


15.2. NONLINEAR FITTING 189<br />

For example, change the contents <strong>of</strong> cell C1 to =$B1*A4. Now the column B is fixed in B1<br />

- try this by again copying cell C1, and pasting into cell D1. The new formula in D1 should<br />

be =$B1*B4. The formula in C1 now reads “multiply the cell in column B, the same row<br />

with the cell two to the left and three down”.<br />

Formula Excel Reads<br />

=B1*A4 “multiply the cell one to the left with. . . ”<br />

=$B1*A4 “multiply the cell in column B, the same row with. . . ”<br />

=B$1*A4 “multiply the cell in row 1, one column to the left, with. . . ”<br />

=$B$1*A4 “multiply the cell B1, with. . . ”<br />

Table 15.2: Absolute referencing.<br />

15.1.5 Using formulas to transform the data<br />

We shall now calculate the momentum associated with each channel as follows. This is<br />

achieved using a transform. Select cell E1, and type 938.3 (this is E 0 ). Select the cell in the<br />

first row <strong>of</strong> column C, and type =SQRT(B1*B1+2$E$1*B1). ‘Fill down’ by selecting cells<br />

C1 through C17, and either pressing Ctrl + d, selecting Edit → Fill down, or clicking on the<br />

small crosshairs at the bottom right corner <strong>of</strong> cell C1, and dragging down to the end <strong>of</strong> C17.<br />

Excel evaluates the formulae so now column C displays the momentum.<br />

15.2 Nonlinear fitting<br />

In simple experiments, a measurable quantity is recorded for various values <strong>of</strong> an experimental<br />

parameter. This information is then plotted, and usually a straight line results. The slope<br />

and vertical intercept are then determined, either by hand (drawing a graph and then using a<br />

ruler to draw a straight line through it) or by some computer program.<br />

But what if the data were more complicated than your average straight line? What if the<br />

theory predicted a slightly curved graph, which couldn’t be simply factored out by squaring<br />

one <strong>of</strong> the variables, or even predicted a wildly squiggly plot? Then, a non-linear curve fit is<br />

needed, which is extremely difficult to do by hand. Excel is good at this, and in this section,<br />

we will use the example <strong>of</strong> a damped sinusoid to demonstrate Excel’s ability to perform<br />

nonlinear fitting. This ability is not a standard feature <strong>of</strong> Excel, however it is easy to install.<br />

15.2.1 The Solver Add-In<br />

If you have already installed the solver add-in, then a quick check in the Tools menu will<br />

show you that the option Solver exists. If it is not there, select Add-Ins (also under the Tools<br />

menu). A window should appear with a number <strong>of</strong> add-ins available; select Solver Add-In,<br />

and click OK. Your computer should take less than a minute to install this, and no disk should<br />

be required. On <strong>University</strong> computers, you may need to repeat this process each time you<br />

log in - let your demonstrator know if this is the case.


190 CHAPTER 15. EXCEL TUTORIAL<br />

Figure 15.1: Solver parameters<br />

15.2.2 Example<br />

Figure 15.2: Solver results<br />

Open damped sinusoid.xls, (you can download it from the part 2 labs website 2 ) and you<br />

should see a number <strong>of</strong> columns <strong>of</strong> data on the left with titles t, y obs, y fit, and error;<br />

parameters on the top right, and two graphs with data points and a curve through them. This<br />

worksheet gives a good example <strong>of</strong> how non-linear fitting works. The function<br />

y = Ae −γt cos(ωt/100 + φ) − B (15.1)<br />

will be used, where t is the independent variable, y is the dependent variable, and the parameters<br />

γ, ω, φ and B are to be obtained by fitting.<br />

There are a number <strong>of</strong> features <strong>of</strong> this spread sheet with which you should be familiar<br />

• Columns I and J show the value <strong>of</strong> the parameters for which the data was generated<br />

from<br />

• Columns L and M show the value <strong>of</strong> the parameters subsequently determined by Excel’s<br />

non-linear fit<br />

• Column A contains the time data<br />

• Column B uses equation 15.1 and the parameters in column J to determine an ‘observed<br />

y’<br />

• Column C gives Column B, rounded to 2 significant figures<br />

2 www.ph.unimelb.edu.au/˜part2


BIBLIOGRAPHY 191<br />

• Column D show the values <strong>of</strong> y as a result <strong>of</strong> the fit, i.e. equation 15.1 using column<br />

M. (Column B could be used, but then we would end up with zero error).<br />

• Cell M7: yields the χ 2 , which is given by the sum <strong>of</strong> the squares <strong>of</strong> the errors (column<br />

E) (This must be squared, to cancel out the possibility <strong>of</strong> some negative and some<br />

positive errors yielding exactly zero error.)<br />

Open the Solver (Tools menu). In the target cell, you should see the cell ‘fit chi sq’ (which is<br />

M7). This cell contains the value we would like to minimise for; ensure that ‘Equal To: Min’<br />

is selected. The cells to be modified to ensure that this is minimised are the fit parameters,<br />

namely fit w, fit phi, fit A, fit gamma, and fit B. Click Solve, and Excel quickly brings up a<br />

new window titled “Solver Results”, which should announce that “Solver found a solution.<br />

All constraints and optimality conditions are satisfied”. Ensure that “Keep Solver Solution”<br />

is selected, and check OK. This first run isn’t so exciting, as the fit was probably almost<br />

perfect in the first place.<br />

Now try and make Excel do some actual work. Start by modifying the values in column M<br />

by small amounts; note how the fit on the graphs no longer matches the data points. It should<br />

become clear at this stage why there are two graphs, not one - the smaller range has enough<br />

space to see individual data points, and the bigger range shows the whole exponentially<br />

damped cosine at once.<br />

If you modify the fit parameters by larger amounts, Excel may not find a solution. This can<br />

easily be checked by eyeballing the graphs to check whether the fit matches the points.<br />

Other ways to play with this could be to change the function entirely. What about sine instead<br />

<strong>of</strong> cosine? What about log rather than exponential?<br />

Obviously, in a real life situation, there would be little reason to generate data using a function,<br />

and then try and fit to it, as we have done here. It is more likely that you could take<br />

some data, (something more complicated than which would yield your average straight line<br />

plot), get a function from theory, calculate the rough parameters, then use Excel to fit to that.<br />

A final suggestion - use your partners! Many <strong>of</strong> you will have much stronger computer skills<br />

than your partners. Take the opportunity to teach them - they can then teach you some physics<br />

in return. Avoid simply doing it for them - it is bad karma, and infuriates demonstrators.<br />

<strong>Bibliography</strong><br />

[1] P Bevington and D K Robinson. Data Reduction and Error Analysis for the Physical<br />

Sciences. McGraw-Hill, 2 nd edition, 1992.


Chapter 16<br />

The Safe Use <strong>of</strong> Lasers<br />

16.1 General comments<br />

A characteristic <strong>of</strong> all laser beams is the low divergence and high power density over long<br />

distances. A small (5 mW) helium/neon laser can have an apparent brightness 1,000 times<br />

that <strong>of</strong> the sun! Laser wavelengths range from ultraviolet through the visible to the far infrared<br />

regions <strong>of</strong> the spectrum.<br />

Many lasers are capable <strong>of</strong> inflicting biological damage, principally through the heat generated<br />

by the interaction <strong>of</strong> light with matter. Very high power lasers may also produce a<br />

thermally induced sonic shock wave which may damage tissue some distance from the site<br />

<strong>of</strong> beam exposure.<br />

The most vulnerable organ is the eye, although skin is also prone to damage. Lasers which<br />

produce light in the visible (400 to 700 nm) and near IR (700 to 1400 nm) regions <strong>of</strong> the<br />

spectrum are particularly hazardous since the eye will focus the beam onto the retina where<br />

a burn could cause impairment <strong>of</strong> vision or even blindness. For wavelengths in the UV (100<br />

to 400 nm) and far IR (∼1400 nm) regions which cannot penetrate the lens, the lens and<br />

cornea are most at risk.<br />

Although the skin can tolerate a great deal more exposure than the eye, the greatest hazards<br />

arise from the IR and near UV regions. Skin burns may result from continued exposure,<br />

leading to the formation <strong>of</strong> cancers.<br />

In addition to the hazards <strong>of</strong> direct exposure, exposure to the specular reflections pose similar<br />

risks.<br />

Finally, lasers usually require high voltage power supplies with the attendant risk <strong>of</strong> electric<br />

shock<br />

193


194 CHAPTER 16. THE SAFE USE OF LASERS<br />

16.2 Precautions with lasers<br />

NEVER LOOK DIRECTLY AT THE BEAM OR<br />

REFLECTED BEAMS.<br />

Avoid exposing the skin.<br />

When viewing the pattern on a diffuse screen, view from the same side as the laser to avoid<br />

direct illumination through pin-hole leaks.


Multiplier Prefix Symbol<br />

10 −24 yocto y<br />

10 −21 zepto z<br />

10 −18 atto a<br />

10 −15 femto f<br />

10 −12 pico p<br />

10 −9 nano n<br />

10 −6 micro µ<br />

10 −3 milli m<br />

10 −2 centi c<br />

10 −1 deci d<br />

10 deca da<br />

10 2 hecto h<br />

10 3 kilo k<br />

10 6 mega M<br />

10 9 giga G<br />

10 12 tera T<br />

10 15 peta P<br />

10 18 exa E<br />

10 21 zetta Z<br />

10 24 yotta Y<br />

10 36 undecillion -<br />

10 100 googol -<br />

10 googol googolplex -<br />

A list <strong>of</strong> multipliers.<br />

Those without symbols are not SI approved.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!