Ebook Electromagnetic Waves and Lasers Second Edition Kimura Wayne D Online PDF All Chapter

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

Electromagnetic Waves and Lasers

Second Edition Kimura Wayne D


Visit to download the full and correct content document:
https://ebookmeta.com/product/electromagnetic-waves-and-lasers-second-edition-kim
ura-wayne-d/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Fundamentals of Optical Waves and Lasers 1st Edition


Sanichiro Yoshida

https://ebookmeta.com/product/fundamentals-of-optical-waves-and-
lasers-1st-edition-sanichiro-yoshida/

Sensors for Ranging and Imaging (Electromagnetic


Waves), 2nd Edition Graham Brooker

https://ebookmeta.com/product/sensors-for-ranging-and-imaging-
electromagnetic-waves-2nd-edition-graham-brooker/

Spin Waves Problems and Solutions 1st Edition Daniel D.


Stancil

https://ebookmeta.com/product/spin-waves-problems-and-
solutions-1st-edition-daniel-d-stancil/

Consumer Behavior, 8e 8th Edition Wayne D. Hoyer

https://ebookmeta.com/product/consumer-behavior-8e-8th-edition-
wayne-d-hoyer/
Electromagnetic Sources and Electromagnetic Fields 1st
Edition Gaobiao Xiao

https://ebookmeta.com/product/electromagnetic-sources-and-
electromagnetic-fields-1st-edition-gaobiao-xiao/

LASERS AND THEIR APPLICATION IN THE COOLING TRAPPING


AND BOSE EINSTEIN CONDENSATES OF ATOMS SECOND EDITION
Richard A. Dunlap

https://ebookmeta.com/product/lasers-and-their-application-in-
the-cooling-trapping-and-bose-einstein-condensates-of-atoms-
second-edition-richard-a-dunlap/

Consumer Behavior 7e 7th Edition Wayne D Hoyer Deborah


J Macinnis Rik Pieters

https://ebookmeta.com/product/consumer-behavior-7e-7th-edition-
wayne-d-hoyer-deborah-j-macinnis-rik-pieters/

Consuming Empire in U S Fiction 1865 1930 1st Edition


Heather D Wayne

https://ebookmeta.com/product/consuming-empire-in-u-s-
fiction-1865-1930-1st-edition-heather-d-wayne/

Surface flute waves in plasmas eigenwaves excitation


and applications Second Edition Manfred Kaspar A Thumm
Igor Girka

https://ebookmeta.com/product/surface-flute-waves-in-plasmas-
eigenwaves-excitation-and-applications-second-edition-manfred-
kaspar-a-thumm-igor-girka/
Electromagnetic Waves and
Lasers (Second Edition)
Electromagnetic Waves and
Lasers (Second Edition)
Wayne D Kimura
STI Optronics, Inc., 1809 130th Ave NE, Suite 118, Bellevue, WA 98005-2201, USA

IOP Publishing, Bristol, UK


ª IOP Publishing Ltd 2020

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system
or transmitted in any form or by any means, electronic, mechanical, photocopying, recording
or otherwise, without the prior permission of the publisher, or as expressly permitted by law or
under terms agreed with the appropriate rights organization. Multiple copying is permitted in
accordance with the terms of licences issued by the Copyright Licensing Agency, the Copyright
Clearance Centre and other reproduction rights organizations.

Permission to make use of IOP Publishing content other than as set out above may be sought
at permissions@ioppublishing.org.

Wayne D Kimura has asserted his right to be identified as the author of this work in accordance
with sections 77 and 78 of the Copyright, Designs and Patents Act 1988.

ISBN 978-0-7503-3523-2 (ebook)


ISBN 978-0-7503-3521-8 (print)
ISBN 978-0-7503-3524-9 (myPrint)
ISBN 978-0-7503-3522-5 (mobi)

DOI 10.1088/978-0-7503-3523-2

Version: 20201101

IOP ebooks

British Library Cataloguing-in-Publication Data: A catalogue record for this book is available
from the British Library.

Published by IOP Publishing, wholly owned by The Institute of Physics, London

IOP Publishing, Temple Circus, Temple Way, Bristol, BS1 6HG, UK

US Office: IOP Publishing, Inc., 190 North Independence Mall West, Suite 601, Philadelphia,
PA 19106, USA
To my wife, Yuka, and children, Kevin and Anna, for all their love and support.
Contents

Preface to second edition ix


Preface to first edition x
Author biography xi
Abbreviations xii

1 What are electromagnetic waves? 1-1


1.1 Electromagnetic radiation 1-1
1.2 Characterization of EM waves 1-6
1.2.1 Maxwell’s equations 1-7
1.2.2 Helmholtz wave equation 1-9
1.2.3 Poynting vector 1-10
1.2.4 Wave–particle duality of EM radiation 1-10
1.3 Basic properties of EM waves 1-11
1.3.1 Phase and group velocity 1-11
1.3.2 EM wave propagation 1-14
1.3.3 Focusing of Gaussian beams 1-16
1.3.4 Non-ideal Gaussian beams 1-18
1.3.5 Beam coherence 1-19
1.3.6 Interferometry using laser beams 1-22
1.3.7 Diffraction 1-25
1.3.8 Light scattering 1-28
1.4 Summary 1-29
References 1-30

2 Lasers and applying EM wave theory 2-1


2.1 How does a laser work? 2-1
2.2 Laser resonator 2-4
2.3 Types of lasers 2-7
2.4 Tuning of laser output wavelength 2-9
2.5 Optical modes 2-11
2.6 Creating short laser pulses 2-14
2.7 Harmonic generation and conversion of laser wavelength 2-17
2.8 Beam control and manipulation 2-18
2.8.1 Reflective optics 2-18
2.8.2 Transmissive optics 2-22

vii
Electromagnetic Waves and Lasers (Second Edition)

2.8.3 Optical filters 2-23


2.8.4 Optics related to polarization 2-25
2.8.5 Dispersive optics 2-31
2.8.6 Optical fibers 2-32
2.9 Optical system design 2-35
2.9.1 ABCD matrix 2-35
2.9.2 General considerations when utilizing lasers for an application 2-37
2.10 Summary 2-40
References 2-41

3 Working with lasers and associated optical components 3-1


3.1 Laser safety 3-1
3.1.1 Laser beam hazard 3-2
3.1.2 Electrical hazards 3-4
3.1.3 Other hazards 3-6
3.1.4 Personal safety and good practices 3-6
3.2 Laser damage 3-8
3.3 Cleaning laser optics 3-11
3.4 Brief overview of optical hardware 3-16
3.4.1 Optomechanics 3-16
3.4.2 Laser diagnostics 3-26
3.4.3 Laser setup 3-30
3.5 Aligning optics 3-36
3.6 Designing and assembling an optical system 3-39
3.6.1 Optical system design and layout 3-39
3.6.2 Assembling optical system 3-42
3.6.3 Operating the optical system 3-54
3.7 Summary 3-55
References 3-56

Appendix A: Radially polarized laser beams and Bessel beams A-1


A.1 Introduction A-1
A.2 Radially polarized laser beams A-5
A.3 Focusing radially polarized laser beams for ICA A-7
A.4 Bessel beams A-8
A.5 Summary A-10
References A-10

viii
Preface to second edition

In keeping with the theme of the first edition, this second edition emphasizes the
practical aspects of working with lasers and their associated optical components. A
third chapter has been added that discusses in extensive detail how to design and
assemble a laser optical system. It covers issues such as laser safety, laser damage,
cleaning optics, commonly used optomechanical devices and equipment, design
strategies and considerations, and step-by-step procedures for assembling the system
and aligning laser beams. Although, it is written assuming the reader has no or little
experience working with lasers and optics, those with more experience may find
some of the techniques that I share to be helpful. In that sense, one can view the third
chapter as a tutorial on using lasers in the laboratory. It contains the same
information that I would teach anyone who works in my laser laboratory.
Also, in this second edition is an appendix intended to provide an opportunity for
the reader to practice using the knowledge gained on electromagnetic waves from
this book in order to understand and appreciate an unusual type of laser beam that is
radially polarized. At the same time, the reader will be introduced to several diverse
technologies and applications, including radiation from relativistic electrons, laser
particle acceleration, and diffraction-free Bessel beams. Thus, the appendix is a way
for the reader to test how well they comprehend the contents presented in this book.
Lastly, new figures and discussions were added to chapter 2.
Dr Wayne D Kimura
STI Optronics, Inc.
University of Washington—Bothell

ix
Preface to first edition

This book was written for a specific target audience of readers. There are many
textbooks written on electromagnetic waves, lasers, and related topics. With regard
to lasers, probably the most comprehensive and authoritative book is Lasers by A E
Siegman (University Science Books, Mill Valley, CA, 1986). However, at 1281
pages, this book can be formidable reading, even for the most ardent student of the
subject matter. And, while other textbooks may not be as lengthy, the details that
need to be covered to thoroughly explain the technical concepts still entail a
concentrated reading effort. This can be a quandary for readers who are interested
in learning more about electromagnetic waves and lasers, but who do not wish to
expend or cannot afford the time and effort required to read these textbooks. This
book is written for those readers.
An abbreviated overview of the basic physics and devices related to electro-
magnetic waves and lasers is presented without going into detailed derivations.
Although equations are presented based upon calculus, this is meant to inform the
reader of the mathematical basis for the other equations given in the book. It does
not require understanding calculus to utilize these other equations. The book does
assume a basic understanding of algebra and trigonometry. Hence, this book is
suitable for undergraduates, graduate students, and professionals in technical fields
that typically do not deal with electromagnetic waves and lasers. Its aim is to provide
these readers with a quick basic understanding of key concepts to help them not only
become familiar with this technology area, but also identify where they would like to
gain more detailed information by consulting the references in the book’s bibliog-
raphy. In that regards, this book is not intended to be a substitute for these other
references, but rather a guide for the reader to help them navigate through the
copious amount of information that is available in books, journal papers, and the
Internet.
This book also attempts to emphasize the practical aspects of this technology
area. Theory is often based upon idealized assumptions and real-life is rarely
idealized. Hence, this book shares various ‘tricks of the trade’ that can be useful to
adapt the theory when working with electromagnetic waves, lasers, and the
associated optics.
The author wishes to thank Dr Albert H Jeans and Dr Kevin W Kimura for their
helpful comments in reviewing the manuscript for this book, and Professor
Sanichiro (Sany) Yoshida for his encouragement and guidance.
Dr Wayne D Kimura
STI Optronics, Inc.

x
Author biography

Dr Wayne D Kimura
Dr Wayne D Kimura received his BS in electrical engineering
(magna cum laude) from the University of Washington in 1976,
and his MS and PhD degrees in electrical engineering from
Stanford University in 1977 and 1981, respectively. His first
experience working with lasers was in 1975 as an undergraduate
using an argon ion laser to stop bleeding ulcers. At Stanford, the
focus of his graduate studies was in quantum electronics and
lasers. His PhD thesis was on the first demonstration of acceleration of relativistic
electrons using a high-power laser beam. He joined STI Optronics, Inc., in 1982, and
he is now President of the company. He is also an Affiliate Professor in the
Department of Electrical Engineering at the University of Washington—Bothell. He
has performed extensive research in the area of lasers and their applications. This
has included work on excimer, Nd:YAG, and CO2 lasers; laser-triggered spark-gap
switches; glancing incidence mirror measurements; ultrashort-pulse CO2 lasers; laser
acceleration of electrons; magnetic field measurements using optical fibers; and
advanced accelerator research and development.

xi
Abbreviations

Acronyms and symbols

2D two-dimensional
3D three-dimensional
Å angstrom = 10−10 m
ABCD matrix method for calculating performance of optics
ANSI American National Standards Institute
APD avalanche photodiode
AR anti-reflection
at attoseconds = 10−18 s
B magnetic flux
c speed of light ≈ 3 × 108 m s−1
CCD charge-coupled device as in CCD camera
cm centimeters
CO2 carbon dioxide as in CO2 laser
COIL chemical oxygen iodine laser
cw continuous wave
d grating groove spacing typically in grooves/mm
D displacement field
D diameter
E, E0 electric field
Emax maximum or peak electric field
Ephoton energy of photon
EM electromagnetic
ELF extremely low frequency
ε0 vacuum permittivity = 8.854 × 10−12 farads m−1
εr relative permittivity
Er radial component of electric field
Er erbium
eV electron-volt
Ez longitudinal component of electric field
η intrinsic impedance
f focal length of lens
f frequency of EM wave, same as ν
f# f-number
FBG fiber Bragg grating
FEL free electron laser
FIR far-infrared
FHG fourth harmonic generation
fs femtosecond = 10−15 s
FWHM full width at half maximum
γ-ray gamma ray
Ge germanium
GeV billion (109) electron-volts
GW gigawatts = 109 W
h Planck’s constant = 6.626 × 10−34 m2 kg s−1
ℏ h/2π

xii
Electromagnetic Waves and Lasers (Second Edition)

H magnetic field
HeNe helium-neon as in HeNe laser
HgCdTe mercury cadmium telluride
Hmax maximum magnetic field
HR high reflection
HV high voltage
HWHM half width at half maximum
Hz hertz ≡ s−1
ICA inverse Cerenkov acceleration
I intensity
IR infrared
J joules
J0 Bessel function of the first kind, order 0
J1 Bessel function of the first kind, order 1
J5 Bessel function of the first kind, order 5
k wave number = 2π/λ
KD*P highly deuterated potassium dideuterium phosphate
kHz kilohertz = 103 Hz
km kilometers
kW kilowatts
λ wavelength of EM wave
L length
LED light emitting diode
LF low frequency
LIA Laser Institute of America
M magnification factor
MeV million (106) electron volts
MW megawatts = 106 W
m meters
MPE maximum permissible exposure
ms milliseconds = 10−3 s
μm microns = 10−6 m
μs microseconds = 10−6 s
MIR mid-infrared
MHz megahertz = 106 Hz
M2 M-squared factor
MM multi-mode
μ0 vacuum permeability = 4π × 10−7 henrys m−1
μr relative permeability
n index of refraction
n number of modes
NA numerical aperture
NaCl sodium chloride
ND neutral density
Nd:YAG neodymium yttrium-aluminum-garnet as in Nd:YAG laser
Nd:YVO4 neodymium yttrium orthvanadate as in Nd:YVO4 laser
NIR near infrared
nm nanometers = 10−9 m
ns nanoseconds = 10−9 s
ν frequency of EM wave, same as f

xiii
Electromagnetic Waves and Lasers (Second Edition)

ω radian frequency, i.e., ω = 2πf


OC output coupler
OD optical density
OPA optical parametric amplifier
OPO optical parametric oscillator
P Poynting vector
Pave average power density
PM polarization maintaining
PMT photomultiplier tube
p-polarization polarized parallel to plane of incidence
PRF pulse repetition frequency
ps picoseconds = 10−12 s
PRF pulse repetition frequency
PW petawatt = 1012 W
Q-switch method for creating short laser pulse by quickly dumping energy
from excited states
σ 1/e point of Gaussian beam profile
r radius
R resistance
s seconds
SHG second harmonic generation
Si silicon
SiO2 silicon dioxide
s-polarization polarized perpendicular to plane of incidence
SM single-mode
SPD spiral-phase-delay
T time period of oscillation
T transfer matrix
TEA transverse excited atmospheric
TEM transverse-electromagnetic
θ divergence of laser beam
θ angular resolution
θ angle of incidence
θc Cerenkov angle
θm diffraction angle for order m
THG third harmonic generation
THz terahertz = 1012 Hz
τspon spontaneous lifetime
TW terawatt = 1012 W
UHF ultrahigh frequency
ULF ultralow frequency
UV ultraviolet
v velocity
vEM velocity of EM wave
vL velocity of light wave
ve velocity of electron
vphase phase velocity
vphoton photon velocity
vgroup group velocity
V volts

xiv
Electromagnetic Waves and Lasers (Second Edition)

VHF very high frequency


W watts ≡ Joules s−1
w0 waist (radius) of focus
Yb ytterbium
ZnS zinc sulfide
ZnSe zinc selenide
zR Rayleigh range

xv
IOP Publishing

Electromagnetic Waves and Lasers (Second Edition)


Wayne D Kimura

Chapter 1
What are electromagnetic waves?

1.1 Electromagnetic radiation


Electromagnetic (EM) radiation refers to a type of energy that is able to propagate
through space. If this space is a vacuum, then this radiation travels at the speed of
light, i.e., approximately 3 × 108 m s−1. As we will explain later, this radiation is
characterized as having both electric and magnetic fields associated with it, which is
why this radiation is referred to as ‘electromagnetic’.
The concept of waves is something we are all familiar with as we walk along the
ocean beach and watch the waves breaking on the sand. It is readily apparent there
are all different kinds of water waves—fast ones, slow ones, ripples on a pond, and
giant tsunamis. Electromagnetic radiation shares many of the same traits as water
waves; thus, it is highly convenient to combine the two concepts into the theory of
electromagnetic waves (or EM waves for short).
This book will provide a broad overview of EM waves, both theory and practical
applications with a special emphasis on lasers. This will not only give the reader
tangible examples of how the theory is manifested in real life, but also practical
knowledge about lasers, and their operation and usage. The latter can be useful to
those involved with using lasers. As a short treatise on this subject matter, this book
is not intended to delve deeply into the details of EM waves nor lasers. References
are provided at the end of each chapter in this book and cited throughout the book
for those who wish to explore in more depth the topics covered in this book. Rather,
the aim of this book is to provide a quick synopsis, which will allow the reader to
gain a competent general understanding of EM waves and lasers.
While lasers are rather specialized devices that are becoming more common place
(e.g., laser pointers), EM waves permeate our everyday lives, 24/7, and are literally
penetrating your body as you are reading this book. In order to gain an appreciation
for the ubiquity of EM waves, it is helpful to first define two of their most
fundamental characteristics—wavelength, represented by λ, and frequency, repre-
sented by ν or f. Figure 1.1 depicts a sinusoidal wave where the z-axis denotes the

doi:10.1088/978-0-7503-3523-2ch1 1-1 ª IOP Publishing Ltd 2020


Electromagnetic Waves and Lasers (Second Edition)

Figure 1.1. Schematic representation of an EM wave.

direction of propagation of the wave and typically has the units of either time or
position. As explained later, this direction of propagation is often indicated by the
so-called k-vector of the wave. EM waves consist of both oscillating electric and
magnetic field components, which are plotted along the x-axis and y-axis in
figure 1.1, respectively. Note, the magnetic field is oriented orthogonal to the
electric field.
The distance between two adjacent peaks or two adjacent valleys in figure 1.1 is
the wavelength of the wave λ in units of length if the z-axis is distance, or the period
of the wave T in units of time if the z-axis is time. How many times per second the
wave oscillates is the frequency of the wave ν, and this frequency depends on how
fast the wave is moving, i.e., its velocity v, where v = νλ. Unfortunately, the
characters v and ν look almost the same, so this can be quite confusing. Fortunately,
when dealing with EM waves, we typically use the velocity of light in vacuum,
denoted by c ≈ 3 × 108 m s−1, as our reference velocity. Therefore, the frequency of
the EM wave can be expressed as ν = c/λ. Since c is a large number, this implies that
for small values of λ, ν can be huge; or, conversely, for low values of ν, λ can be huge.
As we will show shortly in table 1.1, the wavelengths for EM waves can range from
10−11 m to thousands of kilometers. That is over 19 orders of magnitude!
As mentioned, the z-axis in figure 1.1 can have units of time or position. In the
latter form, it is easier to visualize the concept of water waves moving along, say, a
wharf where distance represents the position along the wharf. However, if you were
standing still on the wharf, then the waves would be passing by you as a function of
time. Note, in either case, the wave is the same; it is just how it is being plotted that is
different. When characterizing EM waves, you will often encounter plots where the
abscissa is either time or distance depending on what is being discussed.

1-2
Electromagnetic Waves and Lasers (Second Edition)

Table 1.1. The EM spectrum and example of laser sources.

Name of EM Typical Example of sources Example of lasers


wave wavelength or applications (see text for details)

Gamma rays 10−11 m Radioactive materials


Hard x-rays 10−10 m (1 Å) Medical x-rays, airport
security
Soft x-rays 10−9 m (1 nm) x-ray microscopy and Plasma-based x-ray
holography lasers
Ultraviolet (UV) 10−7 m (100 nm) Tanning booths, black lights Excimer
Visible 400–700 nm LEDs, light bulbs HeNe, dye, ion, SHG,
semiconductor, OPO
Near infrared (NIR) 10−6 m (1 μm) Fiber-optic communication Nd:YAG, Ti-sapphire,
semiconductor, fiber,
OPO, chemical
Mid-infrared (MIR) 10−5 m (10 μm) Clothes iron CO2, Raman scattering
Far-infrared (FIR) 10−4 m (100 μm) Long wave or dark heaters CO2-laser-pumped
methanol
Terahertz (THz) 10−3 m (1 mm) Airport security scanners FEL
Microwaves 10−2 m (1 cm) Radar, microwave ovens
Radio (UHF) 10 cm Cell phones, GPS, Wi-Fi,
Bluetooth
Radio (VHF) 1m TV and FM broadcasting,
radio-controlled devices,
satellite communication
Radio (LF) 103 m (1 km) AM broadcasting, navigation
Radio (ULF) 105 m (100 km) Magnetosphere, earth-mode
communication
Radio (ELF) 108 m Submarine communication,
60 Hz electric power grid

We can now review the tremendously wide range or spectrum of EM radiation.


Table 1.1 summarizes the EM spectrum as characterized by wavelength, the name of
the spectral range, example of sources or applications, and, where applicable, typical
lasers that emit light within that part of the spectrum. Some lasers emit light at only
one wavelength while others do so over a range of wavelengths. Therefore, table 1.1
is only meant to provide a general idea of where various lasers fit within the EM
spectrum.
The shortest EM waves in table 1.1 are gamma rays (γ-rays) with wavelengths less
than the size of an atom. As explained later, the energy in the EM wave is inversely
proportional to λ; hence, γ-rays are one of the most energetic forms of EM radiation.
Radioactive elements commonly emit γ-rays when they decay into smaller elements.
Since γ-rays more easily pass through objects than x-rays, detecting γ-rays is one way
of finding hidden clandestine nuclear bomb materials [1-1].

1-3
Electromagnetic Waves and Lasers (Second Edition)

Hard x-rays are commonly used in medicine and for security screening. A
relatively easy way to generate hard x-rays is by bombarding a target, such as
tungsten, with high-energy electrons to create x-rays via a process called brems-
strahlung [1-2]. Since electrons have an electric field, it is possible to have them emit
EM radiation. In the case of bremsstrahlung, this radiation is emitted when the
electrons are suddenly decelerated by collisions with the tungsten atoms. As we shall
see later, there are other ways electrons can be used to create EM radiation.
Soft x-rays have less energy than hard x-rays and consequently they do not
penetrate materials as easily. However, sometimes this can be a good trait because
unlike hard x-rays, which are energetic enough to disrupt molecules such as DNA,
soft x-rays can more gently probe these molecules without damaging them.
Although there are soft x-ray sources, these tend to have limited output power.
This situation is beginning to change as x-ray lasers are now being realized [1-3, 1-4].
However, as we shall explain later when discussing how lasers work, good resonator
mirrors are always needed, but these are more difficult to fabricate at x-ray
wavelengths because the x-rays want to pass through the mirrors rather than reflect
off them!
Ultraviolet (UV) light from the Sun is responsible for tanning of our skin, but UV
light is also important for photolithography where the short wavelength of UV light
enables creating tiny features on integrated circuits [1-5]. As we will show later, the
resolution of optical instruments, such as photolithography machines, is directly related
to the wavelength of the EM radiation. Excimer lasers [1-6, 1-7] are a powerful source
of UV light and are commonly used as the light source for photolithography.
As you probably remember from your early school days, the visible portion of the
EM spectrum is one of the narrowest, and yet it fills our lives with so many different
colors! And each of these colors are simply slightly different wavelengths of the light,
e.g., blue is around 475 nm, green is about 510 nm, yellow is around 570 nm, orange
is about 590 nm and red is around 650 nm. What this implies is, if we could ‘see’ the
colors of the rest of the EM spectrum, which are much broader than the visible
spectrum, we would be inundated with an unbelievable number of different colors.
This is why the rest of the EM spectrum is so useful because each one of those colors
can be used for different purposes, just as the colors mixed on a painter’s palette can
create beautiful paintings with many nuances of colors—all from just the short band
in the visible spectrum.
There are also a large number of different lasers that emit light in the visible.
Common ones are those emitting red light seen in bar code scanners, the red or green
light of laser pointers, and the ones used in laser light shows. More powerful visible
lasers include helium-neon (HeNe) (λ = 632.8 nm) [1-8] and argon ion (λ = 488.0 nm
and 514.5 nm) [1-8]. Figure 1.2 shows a photograph of a commercial HeNe laser. As
explained later, through a mechanism called second harmonic generation (SHG), it
is possible to take laser light emitted at a longer wavelength, such by a Nd:YAG
laser that emits near-infrared light at 1.06 μm, and convert it into visible laser light at
532 nm
Examples of near-infrared (NIR) sources include your TV remote control and the
semiconductor lasers used for fiber optic telecommunication [1-6, 1-9, 1-10]. Lasers

1-4
Electromagnetic Waves and Lasers (Second Edition)

Figure 1.2. Photograph of HeNe laser emitting light at 632.8 nm. Its red laser beam is exiting the hole in the
center of the cylinder and striking a white card. The red light on the card is so bright, the camera is saturated
and displays the image as a white spot.

that emit light in the NIR include Nd:YAG [1-6, 1-11], which can be used for cutting
materials; Ti:Sapphire [1-12] and optical parametric oscillators (OPO) [1-13, 1-14],
which are highly tunable over different NIR wavelengths making them useful for
scientific research; semiconductor lasers (e.g., AlGaAs, InGaAs, and InGaAsP) [1-6,
1-9, 1-10], which are used in CD and DVD players; and chemical lasers [HF/DF,
chemical oxygen iodine laser (COIL)] [1-6, 1-15, 1-16], which are powerful lasers,
but are powered by chemical reactions and not electricity.
While we cannot see mid-infrared (MIR) light, we can sense it as heat. Thus, a
clothes iron is a great source of MIR light as well as our bodies. Infrared-sensitive
surveillance video cameras are actually detecting the MIR radiation emitted from
our warm bodies. One of the most common MIR lasers used in industrial processing
is carbon dioxide (CO2) lasers (λ = 10.6 μm) [1-6, 1-8, 1-16]. For the same reason
CO2 is a greenhouse gas that absorbs MIR radiation, it is also a great emitter of
MIR radiation. These lasers are capable of emitting kilowatts of power for cutting
and welding materials. We mentioned it is possible to convert NIR laser light into
shorter visible light using SHG. Using a different process called Raman scattering
(see section 1.3.8), it is possible to convert laser light into longer wavelengths, such as
MIR [1-17].
So-called dark heaters are sources of far-infrared (FIR) and, thus, we can still
sense FIR as heat. One way to generate FIR is by using a CO2 laser to excite organic
molecules, such as methanol, which emits FIR radiation [1-16].
Terahertz (THz) waves, also called T-waves or mm-waves because the wavelength
is literally millimeters, are capable of penetrating certain nonmetal materials, much
like x-rays, but are much safer to use because they cannot disrupt molecules [1-18].
This is because the photon energy of THz light (see section 1.2.4) is so low that the
atoms cannot absorb it and, therefore, the THz light passes through the material.
This is why they are routinely used in airport scanners for screening passengers. A
free electron laser (FEL) can easily generate powerful THz light [1-6, 1-19]. An FEL
is a rather special type of laser because it is continuously tunable over a wide spectral

1-5
Electromagnetic Waves and Lasers (Second Edition)

range from x-rays [1-20] to THz. Therefore, it could have been listed under x-rays,
UV, visible, NIR, MIR, and FIR.
We are all familiar with microwaves because of microwave ovens, but micro-
waves are also commonly used in radars [1-21]. High-power microwaves are used to
accelerate electrons and protons to very high energies in linear accelerators and
storage rings [1-22]. As an aside, the predecessor to the laser was the maser, which
produced a beam of microwave radiation [1-23].
When the EM radiation has 10 cm and longer wavelengths, we are entering the
realm of radio waves [1-24]. Table 1.1 only gives a sampling of the many radio bands
that are available. We utilize radio waves during our daily lives in multiple ways,
from cell phones, GPS, and Wi-Fi connections, to watching TV and listening to the
radio. These are generally line-of-sight radio transmissions where an unobscured
pathway between the transmitter and receiver provides the strongest signal. (It is
also possible to receive signals via scattering of the radio wave off objects, but the
signal is much weaker.) This is why cell phone towers are so prolific in order to
ensure cell phone users have a line-of-sight pathway to at least one tower at all times
no matter where the user is located. However, when the wavelength becomes huge,
i.e., >100 km, then obstructions, such as buildings, are small compared to the
wavelength, and the EM wave is able to propagate literally around the obstacles.
These extra long wavelengths can also penetrate through the Earth and ocean
waters, which is handy for communicating with submerged submarines [1-25].
Unfortunately, these extra long wavelengths also correspond to very low frequen-
cies, i.e., a few Hertz, so that data transmission is very slow. Interestingly, one of the
largest sources of ELF waves is from our 60 Hz power grid. The national power grid
in the US is like a gigantic planetary EM wave source!

1.2 Characterization of EM waves


Besides wavelength and frequency, there are other ways to characterize EM waves
[1-2, 1-26, 1-27, 1-28]. The type of EM wave depicted in figure 1.1 is called a
transverse-electromagnetic (TEM) wave because the electric and magnetic field
components are oriented in the transverse direction relative to the direction of wave
propagation. This form of EM wave is one of the most common ones when working
with lasers.
The electric field in the EM wave is typically abbreviated as E-field and has units
of V m−1.1 The magnetic field is typically abbreviated as H-field with units of A m−1;
however, you will sometimes see it abbreviated as B-field. Unfortunately, this can be
confusing because, as explained later, in Maxwell’s equations, B represents magnetic
flux, i.e., B = μ0H, where μ0 = 4π × 10−7 H m−1 is the permeability of free space and
B has the units of Webers m−2 or, equivalently, Tesla (T). (1 Webers m−2 = 10 000
gauss = 1 T.) Thus, the magnetic flux from a magnet can be rightly referred to
as B-field; however, strictly speaking, it is incorrect to call the magnetic field
component in an EM wave the B-field.

1
SI units are generally used throughout this book.

1-6
Electromagnetic Waves and Lasers (Second Edition)

Assuming the EM wave is propagating in vacuum, we can mathematically model


the sinusoidal E-field and H-field in figure 1.1 as either sine or cosine functions, e.g.,
⎡ ⎛ z ⎞⎤ ⎡ ⎛ z ⎞⎤
Ex(z , t ) = E max cos ⎢2π ⎜ft − ⎟⎥ = E max cos ⎢ω⎜t − ⎟⎥ , (1.1a )
⎣ ⎝ ⎠
λ ⎦ ⎣ ⎝ c ⎠⎦

⎡ ⎛ z ⎞⎤ ⎡ ⎛ z ⎞⎤
Hy(z , t ) = Hmax cos ⎢2π ⎜ft − ⎟⎥ = Hmax cos ⎢ω⎜t − ⎟⎥, (1.1b)
⎣ ⎝ ⎠
λ ⎦ ⎣ ⎝ c ⎠⎦

1
E max = cHmax = Hmax, (1.1c )
μ 0 ε0

where ω = 2πf and ε0 = 8.854 × 10−12 farads m−1 is the permittivity of free space.
This shows explicitly how the fields vary both in time and position, and how the
magnitude of the H-field is directly linked with the magnitude of the E-field.
Moreover, in equation (1.1c) we utilize the identity that the speed of light
c = (μ0ε0)−1/2.
Working with trigonometric functions, such as in equations (1.1a) and (1.1b), can
be mathematically awkward at times. Hence, you will often see equations (1.1a) and
(1.1b) rewritten as
Ex(z , t ) = E max e jωte jkz ; Hx(z , t ) = Hmax e jωte jkz , (1.2)
where Euler’s identity, ejX = cos X + j sin X has been used and it is assumed the real
part is taken, i.e., Re[ejX] is implied, but generally not shown in the equation. In
addition, in equation (1.2), the wave number k = 2π/λ is introduced and has typical
units of cm−1. This k is the same as the k-vector noted earlier and, since it is
associated with the z-direction, we now understand why the k-vector relates to the
direction of propagation of the EM wave.

1.2.1 Maxwell’s equations


All the physics associated with electromagnetism are eloquently expressed in
Maxwell’s equations [1-2, 1-26],
∂B
∇×E=− (Faraday’s law of induction), (1.3)
∂t
∂D
∇×H=J+ (Ampere’s law), (1.4)
∂t

∇ · D = ρ (Gauss’ law for electric field), (1.5)

∇ · B = 0 (Gauss’ law for magnetic field), (1.6)


where variables in bold represent vectors. You will also often find Maxwell’s
equations expressed in their integral form rather than the derivative (point) form

1-7
Electromagnetic Waves and Lasers (Second Edition)

of equations (1.3)—(1.6). For a Cartesian coordinate system with unit vectors ux, uy,
and uz in the x, y, and z directions, respectively, the curl is defined by
⎛ ∂E ∂Ey ⎞ ⎛ ∂Ex ∂Ez ⎞ ⎛ ∂Ey ∂Ex ⎞
∇×E=⎜ z − ⎟ux + ⎜ − ⎟uy + ⎜ − ⎟uz ; (1.7)
⎝ ∂y ∂z ⎠ ⎝ ∂z ∂x ⎠ ⎝ ∂x ∂y ⎠

and the divergence is defined by


∂Dx ∂Dy ∂Dz
∇·D= + + . (1.8)
∂x ∂y ∂z
Note, the curl yields a vector, which is why there are vector quantities on the
right-side of equations (1.3) and (1.4); while the divergence yields a scalar, which is
why there are scalar quantities on the right-side of equations (1.5) and (1.6). As
mentioned, E and H are the electric and magnetic field intensities measured in V m−1
and A m−1, respectively; D is the electric flux density or electric displacement
measured in coulomb m−2; B is the magnetic flux density or magnetic induction
measured in Weber m−2 or Tesla; and D = ε0E and B = μ0H in vacuum. J is electric
current density and ρ is volume charge density.
Equation (1.3) is Faraday’s law of induction, which shows that a time-varying
magnetic flux will create an electric field. An electric generator operates on this
principle where moving a magnetic field across wires induces current to flow through
the wire. Conversely, equation (1.4) is Ampère’s law as amended by Maxwell to
include time-varying displacement current, which indicates that either a current
density and/or a time-varying displacement current will create a magnetic field. A
classic example of this is sending current through a wire creates a magnetic field
around the wire.
This complementary role of equations (1.3) and (1.4) can be used to help
understand why EM waves propagate. For example, a radio antenna has a time-
varying current flowing through it created by a transmitter circuit. Through
equation (1.4) this generates a time-varying magnetic field that emanates from the
antenna. Through equation (1.3) this then creates a time-varying current, which
once again generates a time-varying magnetic field, and the process repeats itself.
Thus, the resultant EM wave always satisfies Maxwell’s equations.
Equations (1.5) and (1.6) are Gauss’s law for electric and magnetic fields,
respectively. The physical interpretation of equation (1.5) is that electric charge
can give rise to electric flux density. Gauss’ law comes in handy when analyzing the
fields from a distribution of stationary or moving electrons. In equation (1.6), we see
the right-hand side is zero. The implication of this is that magnetic charges do not
exist, i.e., there are no monopole magnetic sources. Put another way, all magnetic
field lines must be closed or terminate on north and south pole points; they cannot
simply emanate from a point out to infinity as the electric field does from a charged
particle, such as an electron.
Maxwell’s equations are very powerful and are the basis for understanding and
mathematically characterizing many aspects of EM wave theory. In keeping with the
primary aim of this book to provide the reader with a quick synopsis of EM wave

1-8
Electromagnetic Waves and Lasers (Second Edition)

theory, we will not derive these other relationships. Instead, we will simply state
them and encourage the reader to consult the references for more details.

1.2.2 Helmholtz wave equation


The first key equation that can be directly derived from Maxwell’s equations is the
Helmholtz wave equation [1-28], which describes the propagation of an EM wave
∂ 2E
∇2 E = με , (1.9)
∂t 2
where the Laplacian in Cartesian coordinates is defined as
∂ 2E ∂ 2E ∂ 2E
∇2 E = 2
+ 2
+ , (1.10)
∂x ∂y ∂z 2
and μ = μrμ0, where μr is the relative permeability, and ε = εrε0, where εr is the
relative permittivity. The relative permittivity and permeability account for the
dielectric and magnetic susceptibility properties, respectively, of the medium in
which the wave propagates. In that regards, equation (1.9) is valid for an EM wave
propagating through any medium. If the medium is vacuum, then the coefficient on
the right-hand side of equation (1.9) becomes μ0ε0 = c2.
Notice that equation (1.9) is a function of both space (x, y, z) and time (t). By
using the technique of separation of variables, it is assumed the solution of equation
(1.9) can be expressed as the product of two functions—one dependent on only space
and one on only time. This means the Helmholtz equation for only the spatially-
dependent function becomes
(∇2 + k 2 )E = 0, (1.11)
where k is the wave number introduced earlier. Equation (1.11) is important because
it describes how the EM wave can change while traveling through space, for
example, be focused to a spot or create interference patterns.
To illustrate how equation (1.11) can be used, when dealing with laser beams
propagating in, say, the z-direction, the transverse profile of the beam in the x and y
directions generally changes relatively slowly as a function of z. This allows one to
make the approximation of ignoring the ∂2/∂z2 term in the Laplacian in equation
(1.10), which considerably eases solving equation (1.11). This is referred to as the
paraxial approximation [1-29]. The various equations given later in this book
assume the paraxial approximation.
Extending this concept further, equation (1.11) assumes another useful approx-
imation in which the E-field components of x and y are only a function of time and z.
Hence, the EM wave is ‘flat’ in the transverse direction and is called a uniform plane
wave. Although as we shall see, a laser beam cannot stay perfectly collimated
forever, there are regions along the beam where it is quite collimated and within
these regions the light can be treated like a flat plane wave. This makes it easier to
analyze the characteristics of the beam and understand effects such as diffraction.

1-9
Electromagnetic Waves and Lasers (Second Edition)

1.2.3 Poynting vector


As another application of Maxwell’s equations, we can derive the amount of power
flow of an EM wave. The result is the Poynting vector [1-28, 1-29]
P = E × H, (1.12)
−2
which represents the instantaneous power density in W m that flows in the
direction of vector P. If we assume an EM wave as described by equation (1.1), then
the peak power density that arises from applying equation (1.12) is
2
1 E max
Ppeak = , (1.13)
2 η0
were η0 = [μ0/ε0]1/2 = 377 ohms is the intrinsic impedance of free space and Emax is
the peak of the EM electric field. If the wave is propagating through a medium, then
the generalized form of the intrinsic impedance, η = [μ/ε]1/2, would be used in
equation (1.13).
Equation (1.13) shows that the electric field scales as the square-root of the peak
power. In certain applications, it is the electric field of the laser beam that is
important, for example, it is the electric field that interacts with the electrons and, if
the field is high enough, it can ionize the atom by removing an electron from the
atom. In section 1.3.3, we will give an example of how equation (1.13) can be used to
estimate the conditions needed to breakdown air by simply focusing a laser beam
using a lens. The appendix goes into more details on how the electric field can be
used to accelerate electrons to high energies.

1.2.4 Wave–particle duality of EM radiation


We conclude this subsection on the basic characteristics of EM waves by discussing a
fundamental dichotomy. As mentioned, using the wave analogy to characterize EM
radiation is very convenient, but it does have limitations. One of the most important
limitations relates to the wave/particle duality nature of EM radiation [1-2, 1-29]. By
this, we mean that at times EM radiation behaves like waves and at other times it
behaves like particles of energy, which we call photons. As an example, atoms can
absorb or emit photons. To conceptualize this duality, it can be helpful to think of
the photon not as a solid particle, but as a small packet of waves with a frequency
corresponding to the wavelength of the light. Indeed, the amount of energy in a
photon is given by
E photon = hν = ℏω, (1.14)
where h is Planck’s constant (6.626 × 10−34 m2 kg s−1), ν is the frequency of the wave
packet, ℏ = h/2π, and ω is the radian frequency of the wave packet. This shows that
the shorter the wavelength (i.e., the higher the frequency), the higher the energy in
the photon. This is why γ-rays are so energetic.
The notion of light consisting of photons also provides some conceptual benefits.
For example, it is easier to visualize a laser beam as consisting of a stream of

1-10
Electromagnetic Waves and Lasers (Second Edition)

photons that has a well-defined boundary at the edge of the laser beam just like a jet
stream of water has a well-defined boundary and, thus, it is tempting to think the
photons might act like the water molecules. However, in reality the edge of a laser
beam is not as well defined as you might think, and, as we will show, the wave-like
nature of the light cannot be ignored.

1.3 Basic properties of EM waves


Often textbooks will assume EM waves with infinite transverse extents, i.e., infinitely
wide plane waves, in order to facilitate the mathematics. Here, we take a slightly
different approach where we discuss the basic properties of EM waves as manifested
by laser beams. The goal is to provide a more direct connection with the concepts
being discussed and how they are evident in the real world.

1.3.1 Phase and group velocity


The first very basic property of laser light is the velocity of the photons. In vacuum,
the velocity is c, the speed of light. If the photons are traveling through a dielectric,
then their velocity vphoton is less than the speed of light. This amount of reduction is
defined by the index of refraction of the dielectric n, where vphoton = c/n [1-2]. For
example, if the laser light is traveling through quartz (fused silica) that has an index
of refraction of 1.46, then the velocity of the photons is 1.46 times less than c, i.e.,
vphoton ≈ 2 × 108 m s−1. This also means the EM wave velocity is less by this amount.
If one looks at an EM wave propagating along the longitudinal direction z, as
depicted in figure 1.3(a), the velocity of any point on the wave is called the phase
velocity vphase, which is equal to the velocity of the photon, i.e., vphase = vphoton. The
phase velocity is often of importance for laser beams because it is the phase
relationship between the waves within beams that gives rise to phenomena such as
interference effects or mode-locking, which will be discussed in sections 1.3.5 and
2.5, respectively. More precisely, the phase velocity is given by [1-29]

Figure 1.3. (a) Pure EM wave where the wave travels at phase velocity v phase. (b) Modulated EM wave where
the wave still travels at phase velocity v phase, but the envelope travels at group velocity v group, where v group
depends on the dispersion characteristics of the medium that the wave is traveling through and, hence, v group
does not necessarily equal v phase.

1-11
Electromagnetic Waves and Lasers (Second Edition)

c λ ω
vphase = = = , (1.15)
n T k
where λ and ω are their values within the dielectric medium, and recall that k = 2π/λ.
There can be times when the envelope of the EM wave is modulated as illustrated
in figure 1.3(b). The velocity of the envelope is called the group velocity and is
defined as [1-29]
∂ω
vgroup = . (1.16)
∂k
To better understand group velocity, how this modulation can arise, and how this
envelope can move, figure 1.4(a) shows two waves (red and blue waveforms) where
the red wave has a slightly higher frequency than the blue one. If you sum these
waves together, the resultant wave is the modulated green one shown below the two
waves. Where the two waves are approximately in phase, the green wave is nearly
twice the amplitude of the two waves, and where they are approximately in anti-
phase, as indicated by the vertical black line, the green wave is close to zero
amplitude. Note, unlike interfering a laser beam with itself (see section 1.3.6) where
it is possible to have phase points where the combined wave is exactly twice the
amplitude of the original beam or exactly canceled out, because the red and blue
waves have slightly different frequencies, in general there will rarely be phase points
where they exactly double the amplitude or exactly cancel out.
Now suppose the phase of the blue wave shifts slightly with respect to the red
wave. This is illustrated in figure 1.4(b). The black vertical line is still at its original
position on the red wave and it indicates the amount of shift is about a quarter of a
wavelength. The sum of these two waves is again the green modulated wave and its
envelope, as indicated by the black dashed lines, has moved to the right with respect
to the red dashed envelope in figure 1.4(a). The red dashed envelope has been
reproduced in figure 1.4(b) to help see the movement of the black dashed envelope.
In figure 1.4(c), the phase of the blue wave has been shifted more so that, at the black

Figure 1.4. Example of group velocity. (a) Two waves of slightly different frequencies (red and blue
waveforms) summed together to create a modulated wave (green waveform) whose envelope is indicated by
red dashed lines. (b) Same as (a) except the blue waveform is slightly shifted forward in phase with respect to
the red wave resulting in the envelope of the green modulated wave moving to the right. (c) Further shifting in
phase of the blue wave with respect to the red one.

1-12
Electromagnetic Waves and Lasers (Second Edition)

vertical line, the two waves are approximately in phase resulting in the envelope of
the green wave being near its maximum. Thus, we see a shifting of the phase between
waves can cause the envelope of the combined waves to move with a velocity equal
to the group velocity, i.e., in proportion to the change of ω with respect to k.
The example in figure 1.4 begs two questions: (1) how is it possible to have two
slightly different wavelengths that are combined together, and (2) how is it possible
that their phases can slip with respect to each other? The answer to the second
question is the phenomenon called dispersion where, depending on the frequency of
the light, its velocity through a medium will change [1-2]. This is equivalent to saying
the index of refraction of the medium changes depending on the wavelength of the
light. The index of refraction of materials as a function of wavelength is readily
available and generally the refractive index changes slowly versus wavelength,
except when the wavelength nears a resonance of the material [1-30, 1-31]. At the
resonance, the material absorbs the light and the index of refraction no longer
behaves in a simple manner [1-2].
The answer to the first question is that if the light source is not monochromatic, i.e., it
is not a single wavelength, or the light source is extremely short in pulse length, then
there will be a range of different wavelengths propagating though the medium. These
different wavelengths or frequencies will travel at slightly different velocities through the
medium depending on the dispersion characteristics of the medium, thereby resulting in
a combined wave whose group velocity is given by equation (1.16).
As discussed in chapter 2, laser light emission is essentially monochromatic even
though there is a finite range of frequencies being emitted, called the linewidth or
bandwidth of the laser emission, which are centered at the laser emission frequency.
The linewidth is generally very small so that the index of refraction is basically
constant and there is negligible dispersion. Even if the laser emits multiple
longitudinal modes (see section 2.5) that each have slightly different frequencies,
the frequency range is again so small that dispersion is typically not an issue.
Nonetheless, dispersion can have practical ramifications, such as the exact focal
position of a lens depends on the wavelength of the light being focused by the lens.
This is why polychromatic, compound lenses are made in order to provide sharp,
i.e., well-focused, images over the entire visible spectrum.
Ultrashort laser pulses are a different matter. Some lasers are capable of emitting
pulses with extremely short pulse duration, e.g., femtoseconds (fs = 10−15 s) [1-32,
1-33] and now reaching down to attoseconds (at = 10−18 s) [1-34]. Although the laser
light may be centered at a particular wavelength, we know from Fourier analysis
that a pulse can be decomposed into a spectrum of different frequencies whose extent
or range of frequencies is inversely proportional to the time duration of the pulse.
Hence, these ultrashort laser pulses have very large frequency ranges and these
different frequency components can have appreciably different phase velocities
depending on the dispersion characteristics of the dielectric medium the laser pulse is
traveling through. One must avoid sending these ultrashort pulses through any
transmissive optic, such as a window or lens, because dispersion will cause the
different frequency components of the pulse to exit the optic at different times,
thereby increasing the time duration of the pulse.

1-13
Electromagnetic Waves and Lasers (Second Edition)

1.3.2 EM wave propagation


The next basic property of EM waves as manifested by laser beams is how the beam
propagates through space [1-2, 1-29]. Although laser beams are often thought of as
being ‘pencil-beams’ of light, in reality, the laser beam can never stay perfectly
collimated as it propagates. Before we explore this issue in more detail, we must first
examine the typical transverse intensity profile of a laser beam where intensity is
defined as power per unit area and has units such as W cm−2. If we assume the laser
beam is cylindrically symmetric in its intensity profile, then, as illustrated in
figure 1.5(a), a typical laser beam has a Gaussian intensity profile I(r) [1-28, 1-29],
i.e.,
2 /w 2
I (r ) = I0e−2r , (1.17)

where w is the radius of the Gaussian profile corresponding to the 1/e point and it is
also referred to as the spot size of the Gaussian beam [1-29]. The factor of 2 in the
exponent of equation (1.17) arises because it was assumed the amplitude of the
electric field of the Gaussian beam is given by E(r) = E0exp(−r2/w2). The intensity,
which is equivalent to power per unit area, is found by squaring the amplitude (see
equation (1.13)). This also implies that w represents the radius of the electric field
and not the radius of the laser beam intensity.
Some lasers, such as HeNe lasers emitting red light at 632.8 nm, generate laser
beams that fit a Gaussian profile very well. However, depending on other factors,
such as how the laser medium is excited, the beam output may be an elliptically-
shaped Gaussian profile where w in equation (1.17) may have very different values in
the x and y directions. Or, the output may have an almost flat-top profile, as
depicted in figure 1.5(b), and is referred to as a super-Gaussian beam [1-29]. Thus, in
the equations that follow where we characterize the various properties of the laser
beams, it should be kept in mind that these only precisely apply to pure Gaussian
beams. And, since pure Gaussian beams are actually uncommon, the equations are
an approximation of what you will encounter in real life.
This brings us to answering the question of how laser beams propagate through
space. Figure 1.6 shows schematically a Gaussian laser beam traveling in the
z-direction with a transverse intensity profile given by equation (1.17). At z = 0,

Figure 1.5. (a) Generic Gaussian beam profile. (b) Example of super-Gaussian beam profile.

1-14
Electromagnetic Waves and Lasers (Second Edition)

Figure 1.6. Defining propagation parameters for Gaussian laser beam.

the beam has its minimum radius, which is called the waist of the beam with a spot
size value of w = w0. This waist might be located at the output of the laser, i.e.,
immediately on the output side of its output resonator mirror (see figure 2.5), or it
might be at the focal point of a lens that is focusing the laser beam. This is also the
position along the length of the laser beam where it is a uniform plane wave. In
fact, it is the only point where it is a plane wave. At any other position along z in
both the positive and negative directions, the EM wave begins to acquire a
curvature and deviates from being a perfect plane wave. Fortunately, this can be a
gradual process where the laser beam can be considered essentially a plane wave
within a distance called the Rayleigh range [1-29] as indicated in figure 1.6.
By definition, the Rayleigh range is
πw02
zR = . (1.18)
λ
It specifically represents the distance where the area of the laser beam doubles. The
Rayleigh range also marks approximately the division point between the near-field
or Fresnel region (distances less than zR) and the far-field or Fraunhofer region
(distances larger than zR) [1-35]. Within the near-field, the laser beam behaves much
like a collimated beam consisting of wave fronts that can be approximated as plane
waves; whereas, in the far-field, the beam consists of spherical waves that diverge
with distance and grow larger in radius. This phenomenon is called self-diffraction
[1-29] and the divergence angle of the far-field beam (see figure 1.6) is given by
λ w
θ1/e = = 0, (1.19)
πw0 zR
where the angle θ1/e is defined at the 1/e point of the Gaussian beam (see equation
(1.17)).
Equation (1.18) implies a very important property of laser beams. If you want a
laser beam to stay well collimated over a long distance (i.e., large value for zR), you
want the waist to be as large as possible and/or you want to use the shortest possible
laser wavelength. Since you often do not have a choice on the laser wavelength, this
only gives you the waist size to control. But, the diameter of the laser beam being
emitted by the laser is also typically a fixed parameter. This is why expansion
telescopes are utilized at the output of the laser to expand the diameter of the laser
beam in order to achieve a longer Rayleigh range (see figure 3.29). For example, if
λ = 632.8 nm and w0 = 2 mm, then zR = 20 m; but if w0 = 20 cm, then zR = 200 km!

1-15
Electromagnetic Waves and Lasers (Second Edition)

Knowing the waist size of the laser beam, it is possible to calculate the radius of
the Gaussian beam as a function of z in the far-field through the approximate
expression [1-29]
w0z λz
w(z ) ≈ = . (1.20)
zr πw0

1.3.3 Focusing of Gaussian beams


We mentioned in conjunction with figure 1.6 that the waist can also be at the focal
plane of a lens. In fact, a positive-focal-length lens could be positioned on the right-
side of figure 1.6 with a beam of light coming towards the lens from the far right. The
beam outline shown in red in figure 1.6 would represent the focusing of this beam of
light to a focal radius equal to w0. This is more clearly illustrated in figure 1.7 where
a laser beam with incoming radius wi is being focused by a positive (plano-convex)
lens with focal length f. For the sake of this discussion, let us assume that this laser
beam consists of plane waves entering the lens. The lens causes these waves to exit
with the curvature of the exit surface of the lens. It turns out this curvature is the
same as the curvature in the far-field of the beam shown in figure 1.6. Hence, a lens
can be viewed as a means for transforming an EM wave from the far-field to the
near-field, i.e., focusing the beam down to a small spot at its focal plane. As the
beam travels past the focal plane it expands and returns to the far-field (shown in
blue in figure 1.7). Note, at the focal plane of the lens, the focused light has not only
its smallest radius w0, but at this point the wave is once again a plane wave.
With this realization, equation (1.20) can also represent approximately the focusing
properties of the lens if we replace z with f, the focal length of the lens. Thus, the spot
size of a laser beam focused by a lens with focal length f is given by [1-29]
λf
w0 ≈ , (1.21)
πwi
where wi is the radius of the incoming laser beam at the lens.

Figure 1.7. Focusing of laser beam by lens.

1-16
Electromagnetic Waves and Lasers (Second Edition)

In section 1.2.3 with regard to the Poynting vector, we mentioned how equation
(1.13) can be used to estimate the conditions needed to breakdown air by focusing
a laser beam. Suppose you have a Nd:YAG laser with an output pulse energy of
1 J, pulse length of 6 ns, and beam radius of wi = 5 mm. Its peak power is then
1 J/6 ns = 1.7 × 108 W = 170 MW. To cause laser-induced breakdown of air at this
wavelength and pulse duration requires roughly 3 × 1010 V m−1 of electric field. From
equation (1.13), the power density needed to achieve this field is 1.2 × 1018 W m−2.
This implies that our laser beam needs to be focused down to an area no greater than
(170 MW)/(1.2 × 1018 W m−2) = 1.4 × 10−10 m2. Hence, the waist spot size needs to
be equal to or smaller than w0 = [(1.4 × 10−10 m2)/π]1/2 = 6.7 × 10−6 m = 6.7 μm.
To achieve this spot size, equation (1.21) indicates for wi = 5 mm and λ = 1.06 μm,
that the focal length of the lens should be equal to or shorter than 10 cm. If you sent
this laser beam through this lens, then at 10 cm from the lens you will see bright
flashes of light suspended in mid-air and hear loud snapping sounds every time the
laser fires and breaks down the air!
Returning again to equation (1.21), as mentioned, laser beams are often not
precisely Gaussian so that expressions, such as equation (1.21), become less accurate
in their approximation as the beam profile deviates more from being purely
Gaussian. Moreover, the variable w in these equations are the 1/e values for the
Gaussian E-field; therefore, it is the 1/e2 value of the intensity profile that needs to be
measured, i.e., the radius where the intensity is 86% down from the peak. The 1/e2
intensity level for a Gaussian beam is shown in figure 1.8. It can be inconvenient or
awkward to measure the 86% intensity point. Instead the width of the Gaussian
profile is often measured at the point where the intensity is one-half its maximum
value. As illustrated in figure 1.8, this is called the full-width-at-half-maximum
(FWHM) and is equal to 2rFWHM, where rFWHM is the radius at FWHM.

Figure 1.8. Different conventions for measuring width of Gaussian profile. FWHM stands for full-width-at-
half-maximum. Sometimes you will see HWHM, which is half-width-at-half-maximum, and is the radius of the
Gaussian beam at the half-maximum point.

1-17
Electromagnetic Waves and Lasers (Second Edition)

An alternative width that is also straightforward to measure is at the 99% points


(see figure 1.8), which is essentially the same as measuring the total beam
diameter. This is equal to 2r99%, where r99% is the radius at the 99% point. It is
then possible to use these more easily measured quantities to calculate the w
radius for the Gaussian profile. Specifically, w = 1.70rFWHM = 0.659r99% [1-36].
As another way of calculating the spot size of a lens, we note the f-number or
speed of a lens is defined as [1-37]
f 1
f# = = , (1.22)
D 2NA
where D is the diameter of the lens and NA = D/2f is called the numerical aperture of
the lens [1-37]. Implied is that the lens is gathering all the light impinging on it and
focusing it to a spot. For the purposes of calculating the spot size of a laser beam
focused by the lens, let us assume the entire laser beam fills up the diameter of this
lens. (In reality, as explained in section 1.3.7, we would not want to entirely fill up
the lens aperture with the laser beam because of diffraction effects caused by the
edge of the aperture.) In that regards, D represents the diameter that contains, say,
99% of the laser beam power. This is a convenient parameter to work with because
generally one is interested in the beam diameter that contains essentially all the beam
power and, as mentioned, this 99% diameter is also easier to measure in the
laboratory.
Notice that the numerical aperture of the lens is inversely related to the f#. Thus,
a large f# where the focal length of the lens is long compared to the lens diameter
corresponds to small NA. This is generally the regime one would like to operate in
because the approximations for the Gaussian beam optics become progressively less
accurate as the f# becomes smaller and NA becomes larger, i.e., short focal length
lenses.
The diameter of the focused spot containing 99% of the beam power is then [1-36]
λ 2fλ
2r99% ≈ = 2f #λ = . (1.23)
NA D
We should emphasize that equation (1.23) is an approximation that becomes less
accurate for small f#. This means a f# of order 1 does not mean you can achieve a
spot size of order λ. In general, the spot size will be much larger than λ because
typical laser beams are not pure Gaussian ones. This brings us to another
characteristic of laser beams.

1.3.4 Non-ideal Gaussian beams


As mentioned, when a laser beam deviates from being a pure Gaussian (e.g., the
super-Gaussian in figure 1.5(b)), the approximations in the preceding equations not
only become less accurate, the actual results are also not as good as the ideal results.
For example, besides a non-Gaussian beam being unable to focus to as small a spot
size as a pure Gaussian beam, a non-Gaussian beam will also tend to spread out
faster than a pure Gaussian beam, i.e., its Rayleigh range will be shorter than

1-18
Electromagnetic Waves and Lasers (Second Edition)

predicted by equation (1.18). This is why pure Gaussian beams are also called
diffraction-limited beams because their amount of spreading as a function of
propagation distance is only limited by equation (1.18).
To help quantify the degree to which a non-Gaussian beam deviates from a pure
Gaussian one, equations (1.18), (1.19), and (1.20) can be modified as follows [1-38]
πw02 M 2λ M 2λz
zR = ; θ1/e = ; w(z ) ≈ . (1.24)
M 2λ πw0 πw0
where M2 is called the M2 factor or beam quality factor and is always greater than or
equal to unity. When M2 = 1, the beam is a diffraction-limited, pure Gaussian beam.
Real laser beams tend to have M2 > 1 and this is often empirically determined, for
example, by measuring the achieved spot size when the beam is focused by a lens and
comparing this with the spot size predicted by equation (1.20). The ratio of the two
spot sizes yields the value for M2. Good quality laser beams will have M2 values
close to 1; poor quality beams may be many times diffraction-limited in their
characteristics.

1.3.5 Beam coherence


The preceding characteristics related to Gaussian beams, e.g., the ability to focus to
very small spots size, are related to another important property of laser beams that
makes them unique in the family of EM waves. This is the coherence of the laser
beam [1-27, 1-29]. As explained later, one special attribute of lasers is their ability to
generate very pure colors of light, i.e., emit radiation at a single wavelength
(monochromatic), unlike, say, an incandescent light bulb that emits light over a
broad wavelength range. However, producing monochromatic light by itself does
not necessarily yield diffraction-limited Gaussian beams. For example, a light
emitting diode (LED), which generates light at one wavelength, emits incoherent
radiation. Coherence is achieved when all the EM waves within the laser beam are
synchronized in phase with each other. This coherence is illustrated in figure 1.9(a)
where the EM waves are all propagating perfectly in phase, that is, their peaks and
valleys are all aligned with each other. Incoherence is when the phase relationship

Figure 1.9. (a) EM waves coherently in phase. (b) EM waves incoherently out of phase.

1-19
Electromagnetic Waves and Lasers (Second Edition)

between the EM waves is completely random as depicted in figure 1.9(b). It is also


possible to have partial coherence where the waves are somewhat in phase [1-27].
Thus, a diffraction-limited laser beam has a high degree of coherence. A poor
quality laser beam with a large M2 value may have partial coherence. Indeed, until
the invention of lasers, it was difficult to create truly coherent light sources.
Ironically, once a coherent beam has been generated, it is actually difficult to
destroy the coherence. An example of this can be seen whenever one shines, say, the
red beam (λ = 632.8 nm) from a HeNe laser onto a wall (see figure 1.10). If you
examine the laser spot carefully, you will see a speckle pattern consisting of bright
and dark spots within the laser beam image on the wall [1-39]. These spots will seem
to move around especially if you look at the laser spot from different angles. What is
not obvious is that anyone else looking at the same laser spot will see a different
speckle pattern than the one you are observing. This is because the speckle pattern is
caused by the EM waves from the laser beam scattering off the atoms on the walls at
different angles. The waves are coherent before they strike the wall; however, when
they reflect off the wall, their phases with respect to each other shift because they are
scattering at different angles. This means the waves create an interference pattern,
where the peaks of one wave might line up with the valley of another wave, thereby,
canceling each other out and creating a ‘dark’ spot. Depending on where you are
standing, the angles of the scattered light are different. This is why everyone sees a
different speckle pattern.
Coherence not only gives rise to the basic properties of Gaussian beams, it is also
important for many applications. An example is holography [1-40] whose basic
concept is illustrated in figure 1.11. The first step is to imprint a hologram, which will
be the piece of photograph film depicted in figure 1.11(a). The output from a laser,
say a HeNe laser, is sent through a beam-splitter so that a portion of the beam
illuminates the object (probe beam) and the other portion is directed at the

Figure 1.10. Laser spot from HeNe laser (λ = 632.8 nm) shining on a wall. The mottled pattern is called
speckle and is due to the coherence of the beam, which causes the laser light scattering off the wall to interfere
with itself.

1-20
Electromagnetic Waves and Lasers (Second Edition)

Figure 1.11. Basic scheme for creating and viewing a hologram. (a) Light from a laser beam scatters off the
object and at the same time illuminates the photographic film, which creates an interference pattern on the film,
thereby creating the hologram. (b) The photographic film (hologram) is illuminated by the laser beam and, due
to the interference pattern on the hologram, the observer sees the same scattered light pattern that was created
by the object, thereby creating a virtual 3D image of the object.

photographic film (reference beam). Also shown in figure 1.11(a) are the wavefronts
of the beams that are oriented perpendicular to the direction of the beam
propagation. The wavefronts of the probe beam scatter off the object like water
waves reflecting off a pier. When these scattered wavefronts reach the position of the
photographic film, they coherently interfere with the wavefronts of the reference
beam. Due to the coherence between the two sets of wavefronts, there will be regions
of high light intensity where the wavefronts add together and regions of low light
intensity where the wavefronts cancel each other out. This produces an interference
pattern that is recorded by the photographic film and will look like random regions
of dark and light spots on the film. In fact, it is not necessary to use color film, black
and white film is fine.
The procedure for viewing the hologram is given in figure 1.11(b). A HeNe laser
beam is directed at the film. When the laser light passes through the film, the film
scatters the light to create the same scattered light distribution pattern that was
produced when the real object was scattering light towards the photographic film.
To an observer viewing this scattered light pattern, it looks exactly like what they
would see had they been present when the film was being exposed. Thus, the
observer will see a virtual image of the object. The 3D aspect of the hologram occurs
because when the film was exposed in figure 1.11(a), the scattered light from the
object came from all directions and angles with respect to viewing the object. When
the laser light scatters off the film in figure 1.11(b), it reproduces these same
directions and angles. Therefore, if the observer looks at the hologram from a
different direction, they will see the light that the object would have scattered in that
direction. To the observer it will appear they are looking at the object from a
different angle, thereby making the object appear 3D.

1-21
Electromagnetic Waves and Lasers (Second Edition)

1.3.6 Interferometry using laser beams


The concept of interference was introduced when explaining how a hologram works
and the importance of coherence. This concept will be explored further by discussing
interferometry [1-27]. Although there are many different types of laser interferom-
eters, the basic concept is to split the beam into two beams and then recombine the
beams at some point in space so that they create an interference pattern consisting of
bright and dark lines called fringes. Figure 1.12 shows the basic layout for a
Michelson interferometer. A laser beam (colored in blue) enters from the left. To
make a useful interferometric image, you generally need a laser beam with a finite
width, as depicted in figure 1.12. Achieving the desired width can be accomplished
by using an expansion telescope to enlarge the laser beam emitted by the laser (see
figure 3.29).
The laser beam is directed towards a 50/50 beam splitter. This is generally an
optic whose substrate material transmits the laser beam with low absorption, for
example, fused silica for visible and NIR light. It also has highly polished and very
flat surfaces to minimize scattering and distortion of the laser light from the
surfaces. One surface of the beam splitter has a multi-layer dielectric coating
[1-41, 1-42] deposited on it that has been designed to reflect 50% of the laser light
at 45° to the surface of the beam splitter and transmit the remainder of the light
through the optic. The other side of the beam splitter also has a multi-layer
dielectric coating, except this coating has been designed not to reflect any of the
laser light. It is referred to as an anti-reflection coating (AR-coating). The AR-
coating prevents the back surface of the beam splitter from reflecting the laser
beam, which would cause multiple images at the output of the interferometer.
This is important in order to achieve clear interference fringes from the
interferometer. (Section 2.8.1 explains more about optical coatings.)

Figure 1.12. Basic layout for a Michelson interferometer.

1-22
Electromagnetic Waves and Lasers (Second Edition)

In figure 1.12, the blue laser beam that passes through the beam splitter is
reflected by Mirror #1 directly upon itself so that when the reflected beam reaches
the beam splitter, it is reflected downward off the 50% reflective coating towards an
image detector, in this case, a video camera. We shall refer to this blue beam as the
reference beam. Note, both Mirrors 1 and 2 have their reflecting coating on the front
surface of the mirror so that the laser beam does not pass through the mirror
substrate.
The 50% of the laser beam that is reflected by the beam splitter is directed towards
Mirror #2. This beam has been colored yellow to differentiate it from the beam that
passes through the beam splitter. Mirror #2 reflects the yellow beam back towards
the beam splitter, but the mirror has been intentionally adjusted so that the yellow
beam is slightly skewed relative to the blue beam as the two beams overlap each
other at the video camera. The reason this is important is illustrated in figure 1.13,
which shows schematically the EM waves of the blue and yellow laser beams in
figure 1.12. For the purposes of this discussion, we can assume the colored bars
represent the positive peaks of the EM waves. If the two laser beams are exactly
parallel to each other, then there is no variation in the overlap of the EM waves
across the widths of the beams, as shown in figure 1.13(a). If the peaks of the yellow
beam happen to overlap in the negative valleys of the blue beam, it will cancel out the
electric field of the blue beam (so-called destructive interference). The video camera
would then detect no light. Conversely, if the peaks of the two beams overlap each
other (so-called constructive interference), then the camera will only see a uniform
bright image. This is referred to as operating on the zero order of the interferometer.
On the other hand, if the beams are slightly skewed, then as depicted in
figure 1.13(b), the EM waves cross each other creating regions in space where the
peaks sometimes overlap each other or sometimes overlap the valleys. The camera
will thus see a regular linear pattern of bright or dark lines. These are the fringes seen
by the camera. Notice in figure 1.13(b) that the horizontal distance over which, say,
the peaks overlap each other depends on the angle of the yellow beam relative to the
blue one. As the angle becomes larger, the horizontal distance gets smaller, which
means the fringes become narrower in width. Thus, skewing the beams permits
operating at higher orders on the interferometer where the higher the order, the
narrower the fringes.

Figure 1.13. Cartoon of interfering waves where the colored bars represent the peak of the wave. (a) If the two
laser beams are exactly parallel to each other, there are no fringes. (b) If the two laser beams are slightly
skewed, there are fringes.

1-23
Electromagnetic Waves and Lasers (Second Edition)

Figure 1.13(b) also implies that to achieve the sharpest contrast in the fringe
pattern requires utilizing a monochromatic light source consisting of perfect plane
waves. Lasers are the closest to being such a source, but they are not perfect. As
mentioned, there is generally a finite bandwidth of wavelengths over which the laser
emits light, so it is only partially monochromatic. Furthermore, even a collimated
laser beam still has some wavefront curvature, i.e., the fronts are not perfect plane
waves. In addition, if the optics in the interferometer (i.e., mirrors and beam splitter)
have any slight surface curvature, then this will further reduce the contrast. This is
why when specifying the optical quality of these optics, you generally want surface
flatnesses significantly less than the laser wavelength (e.g., <λ/10).
Figure 1.14(a) is a photograph of the fringe pattern produced by a laser
interferometer used to probe an electrical arc discharge [1-43]. In this case, the tilt
of the mirrors of the interferometer were adjusted to produce a horizontal fringe
pattern. The fringes appear wider in the middle of the photograph due to the laser
intensity being brighter in the middle since the laser beam had roughly a super-
Gaussian profile. The arc is located in one branch of the interferometer. When the
arc is present (see figure 1.14(b)), it ionizes the gas within the arc and creates a hot
plasma. This plasma has a different index of refraction compared to the surrounding
gas. Therefore, laser light passing through the arc experiences a phase shift relative
to the surrounding gas and results in the fringe pattern shifting within the region
where the arc is present.
These photographs illustrate how a laser interferometer is able to provide 2D and
even 3D information. For example, in figure 1.14(b), we see the fringe pattern within
the arc tends to shift up at the edges of the arc and shift down in the middle. This
behavior indicates that the plasma distribution within the arc is somewhat like a
hollow tube with higher plasma density at its walls. If you assume the arc is
circularly symmetric, which is a reasonable assumption, then it is possible to
transform the 2D information from this fringe pattern into a 3D plasma distribution
using a process called Abel transformation [1-44, 1-45].
Operating an interferometer on zero order with no fringe pattern can be useful
when using the interferometer as a means to precisely measure changes in distance.

Figure 1.14. Example of interference fringe pattern produced in a laser interferometer used to probe an
electrical arc discharge. (a) Fringe pattern with no arc present. (b) Fringe pattern with arc oriented vertically in
photograph.

1-24
Electromagnetic Waves and Lasers (Second Edition)

For example, suppose Mirror #2 is stationary, but Mirror #1 is allowed to move,


e.g., it might be sitting on a metal bar that contracts or expands in its length
depending on how much you heat the bar. This will cause Mirror #1 to move left or
right in figure 1.12 by a very small amount, e.g., microns. This movement can be
detected by setting the interferometer to zero-order and using a photodetector,
which can be a photodiode (see figure 3.23), to sense when the light intensity
increases and decreases. If you are using a HeNe laser (λ = 0.6328 μm) as the light
source for the interferometer, then this implies you will be able to detect movements
as small as 0.6328 μm/2 = 0.3164 μm. The factor of two arises because when the light
reflects off Mirror #1, it travels twice the distance that Mirror #1 has moved.
The AR-coating on the beam splitter in figure 1.12 provides an opportunity to
discuss another aspect of EM waves. This coating is created by depositing thin
multiple layers of two different dielectric materials on the beam splitter surface that
have different index of refraction values at the laser wavelength. Laser light reflects
off these layers; however, because of their different index of refraction, the phase
velocity of the wave within each material is different. It is possible by controlling the
thickness of the layers to have the reflected light from one dielectric layer be shifted
180° out of phase (π-phase shift) relative to light reflected by the other dielectric layer
(see section 2.8.1). Then, in a similar fashion, as shown in figure 1.13(a), the peaks of
one reflected wave overlap the valleys of the other reflected wave resulting in zero
light intensity. Thus, there is no reflected light.
In reality, because perfect laser beams and perfect AR-coatings do not exist,
there is some reflected light, but it is generally greatly reduced. By judicious design
of the multi-layer AR-coatings, it is now possible to make broadband coatings that
work well over a wide wavelength range. This has been applied, for example, on
AR-coated sunglasses where the coating needs to work over the visible sunlight
spectrum.

1.3.7 Diffraction
Thus, having coherence permits many useful applications. However, it can also be a
bane. Diffraction occurs when the EM wave interferes with itself [1-27]. Classic
examples are the single and double slit experiments where a symmetric diffraction
pattern is created when the laser beam passes through the slits. Another example is
knife-edge diffraction where the laser beam scrapes the sharp edge and creates an
asymmetric diffraction pattern. In each case, it is the wave nature of the EM wave
that best explains the creation of the diffraction pattern, i.e., the edges are a source of
scattered EM waves that constructively or destructively interfere with the non-
scattered EM wave.
Figure 1.15(a) shows an example of diffraction that occurs when a HeNe laser
beam is sent through a pin-hole whose diameter is much smaller than the laser beam
diameter. A series of concentric rings is formed called an Airy pattern [1-37]. The
light that forms these rings is being scattered at an angle from the edges of the pin-
hole and, hence, the light that forms these rings emanates as a cone of radiation from
the pin-hole. The intensity profile of an Airy pattern is depicted in figure 1.15(b) and

1-25
Electromagnetic Waves and Lasers (Second Edition)

Figure 1.15. (a) Diffraction of HeNe laser light when the laser beam passes through a hole whose diameter is
much smaller than the laser beam diameter. (b) Cross-section of Airy pattern intensity profile, similar to the
cross-section profile of the beam in (a).

Figure 1.16. Diffraction of HeNe laser beam when the beam passes through an iris whose diameter is slightly
smaller than the laser beam.

this would be similar to what would be measured along any diameter of the image
shown in figure 1.15(a). The rings are much fainter than the main central lobe.
Now suppose instead of a pin-hole we pass the HeNe laser beam through an
aperture (iris) whose diameter is only slightly smaller than the laser beam. What we
observe downstream of the iris is shown in figure 1.16. Notice that, unlike the fairly
uniform laser spot image shown in figure 1.10, we see a series of bright and dark
rings inside the laser spot. These are called Fresnel rings [1-29] and they are due to
diffraction caused by the laser light scraping the inside edge of the iris. What is
surprising is that these Fresnel rings can still occur even when the iris diameter is
considerably bigger than the apparent diameter of the laser beam. I say ‘apparent’
because the laser beam is a Gaussian one and, technically speaking, the tails of a

1-26
Electromagnetic Waves and Lasers (Second Edition)

Gaussian extend to infinity. Although the diameter of the laser beam certainly does
not extend to infinity, it can still have sufficient intensity near the edge of the beam,
which cannot be seen, to cause diffraction effects beyond r99%. As a practical matter,
to completely avoid creating Fresnel rings, the iris radius should be larger than
about 3r99%.
Fresnel rings can be a nuisance when they arise during constructing, say, an
interferometer because the rings will superimpose themselves on the fringe pattern.
There can be, however, an even more sinister problem. Under the right conditions,
through constructive interference, the Fresnel rings can concentrate themselves at
the very center of the laser beam, thereby creating an intense spot of laser power.
This spot is called the Spot of Arago [1-29] and an example is shown in figure 1.17.
Although not very bright in this example, the laser intensity in this spot can be high
enough to damage an optic that happens to be located where this spot has formed.
Diffraction either by the edges of, say, an aperture or self-diffraction of a
Gaussian beam ultimately affects the ability to distinguish features associated
with the EM wave, for example, discerning the profiles of two laser beams lying
close to each other. This is related to the resolution of the optical system and it can
be very important in many applications. As a simple example, the angular resolution
θ of the Airy pattern shown figure 1.15 is given by [1-37]
λ
θ = 1.22 , (1.25)
D
where θ is in radians, λ is the wavelength of the light, and D is the diameter of the
aperture or lens with λ and D measured in the same units, e.g., meters. This implies
for a fixed value of D that finer angular resolution occurs at shorter wavelengths.
Hence, if one wants to resolve tiny features using an optical probe, it is better to use

Figure 1.17. HeNe laser beam at a position downstream of an iris showing the Spot of Arago that has been
created in the center.

1-27
Electromagnetic Waves and Lasers (Second Edition)

as short a wavelength as possible. In an analogous manner, equation (1.21) shows


that the spot size of a focused Gaussian beam is directly proportional to the
wavelength of the light. Hence, for a fixed focal length and beam input size, a smaller
focal spot is possible by using a shorter wavelength. This is why Blu-ray discs are
able to contain much more digital data than regular DVD discs because they utilize
a laser diode emitting light at 405 nm rather than the laser diodes in DVD players
that emit light at 650 nm. This permits writing and resolving smaller pits in the disc
that represent the digital information.

1.3.8 Light scattering


The laser beam images in figures 1.10 and 1.15–1.17 also demonstrate the simple
phenomenon of light scattering, in this case, off a wall to enable obtaining a
photograph using a camera. In this situation it can be more convenient to refer to the
photons scattering off the molecules in the wall rather than the EM wave. Usually
the photons are elastically scattered, i.e., the energy of the photon does not change
and, thus, by equation (1.14), their frequency does not change. This is called
Rayleigh scattering [1-2].
Figure 1.18(a) illustrates what is happening during Rayleigh scattering from the
viewpoint of the energy of the photon and the molecule interacting with the photon.
A photon with a frequency ν0, corresponding to an energy hν0 (see equation (1.14)),
impinges on the molecule, which is at its lowest energy state E0. The photon causes
the molecule to be excited momentarily into a higher energy state corresponding to
the energy of the photon. This higher state is called a virtual state because it is
generally not associated with one of the allowed energy states of the molecule. What
usually happens is that the molecule relaxes back down to its original lowest energy
state and reemits the photon with the same energy as the incoming photon, i.e., hν0;
therefore, the reemitted light is at the same wavelength as the incoming light.
However, this remission is generally in a random new direction relative to the
incoming photon’s direction, which results in a scattered light distribution.

Figure 1.18. Rayleigh scattering versus Raman scattering. (a) Rayleigh scattering. (b) Raman scattering with
Stokes emission. (c) Raman scattering with anti-Stokes emission.

1-28
Electromagnetic Waves and Lasers (Second Edition)

Sometimes, though, the molecule relaxes down to an energy level that is higher
than the ground state, i.e., level E1 shown in figure 1.18(b). Now the reemitted
photon has a lower energy than the incoming photon, i.e., hνS < hν0, which implies
the reemitted light is at a lower frequency corresponding to a longer wavelength. As
depicted in figure 1.18(c), a complementary process is also possible where the
molecule happens to be in the excited state E1, the incoming photon excites the
molecule to a virtual level, and the molecule relaxes down to the ground state E0.
In this case, the reemitted photon has a higher energy than the incoming one,
i.e., hνAS > hν0, which means the reemitted light has a shorter wavelength than the
incoming one. These are examples of inelastic scattering because the energy of the
reemitted photon is different than the incoming one.
The two processes shown in figures 1.18(b) (c) are referred to as Raman scattering
[1-17], and when the reemitted light is at a longer wavelength (i.e., figure 1.18(b)) it is
called Stokes scattering and when it is at a shorter wavelength (i.e., figure 1.18(c)) it is
called anti-Stokes scattering. Stokes scattering tends to occur more often than anti-
Stokes scattering because most molecules are typically in the ground state and only a
few in excited states. Because the probability of having Stokes or anti-Stokes scattering
occurring is very small, it was not until the invention of lasers with their ability to emit
copious amounts of photons at one wavelength and maintain a relatively small beam
diameter over long distances (i.e., long Rayleigh range) did it become possible to
generate appreciable Stokes or anti-Stokes radiation. And, even then, special Raman
cells are generally used, which are typically long tubes (e.g., ~2 m) with windows at
their ends in which the laser beam enters on one end and the Stokes or anti-Stokes
radiation exits the opposite end. The tubes are filled with a gas (e.g., hydrogen) that
serves as the Raman medium. The gas is at very high pressures (e.g., 10 atm) in
order to maximize the number of molecules that can interact with the laser photons.
The concept of photons interacting with the energy levels of a medium is one of
the key aspects of how lasers work. This is a convenient segue into the next chapter
that will discuss lasers in more detail.

1.4 Summary
(1) The propagation of EM waves can be mathematically characterized by
using the Helmholtz wave equation as derived from Maxwell’s equations.
(2) The average power flow in an EM wave is given by the Poynting vector
and is equal to Pave (W) = Emax2 (V/m)/377 Ω.
(3) Photons can be conceptualized as compact packets of waves with wave-
length λ and frequency ν, and an energy equal to hν.
(4) The phase velocity of an EM wave is vphase = c/n = ω/k, where c is the
speed of light, n is the index of refraction, ω is the radian frequency of the
light, and k = 2π/λ. The phase velocity corresponds to how fast the wave is
moving in any medium.
(5) The group velocity of EM waves is vgroup = ∂ω/∂k and corresponds to how
fast the envelope moves consisting of combined waves with different
frequencies.

1-29
Another random document with
no related content on Scribd:
'Splitting hairs don't you mean, Chris?' asked Mrs. Winthrop with a half-
smile; 'but I see the difference, Snap. There is no disgrace about this, is
there?'

'No, I didn't think so,' replied Snap, 'but my uncle says I am a disgrace to
my family and always shall be.'

'He always did say that,' muttered the Admiral. 'Never mind what your
uncle says; I mean,' added the old gentleman, correcting himself, 'don't take
it too much to heart. You see he has very strict ideas of what young lads
should be.'

'What is it that you have been doing, Snap? Is it too bad to tell me?' asked
Mrs. Winthrop after a while.

For a moment the boy hung his head, thinking, and then raised it with a
proud look in his eyes.

'No, dear,' he said, dropping unconsciously into an old habit, 'it isn't, and
so it can't be very bad!' And with that he told the whole foolish story of his
share in the smoking orgy, of his reprieve, of the mop incident and the
bolster fight, and, last of all, of that Fernhall ghost.

At this part of the recital of his wrongdoings the Admiral's face, which
had been growing redder and redder all the time, got fairly beyond control,
and the old gentleman nearly went into convulsions of laughter. 'Shameful,
sir; gross breach of discipline, sir; ha! ha! ha! "Don't like me in the spirit,
had better take me in the flesh." Capital—cap—infamous, I mean, infamous.
Your uncle never did anything like that, sir, not he,' spluttered the veteran;
'couldn't have done if he had tried,' he added sotto voce.

'But,' said Mrs. Winthrop, after a pause, 'what are you going to do, Snap?'

'My uncle wants me to go into the Church or Mr. Mathieson's office,'


replied the boy.

'The Church or Mr. Mathieson's office—that is a strange choice, isn't it?'


asked his friend. 'Which do you mean to do?'
'Neither,' answered Snap stoutly; 'I'm not fit for one, and I should do no
good in the other. I shall do what some other fellows I know have done. I'll
emigrate and turn cow-boy. I like hard work and could do it,' and half
consciously he held out one of his sinewy brown hands, and looked at it as if
it was a witness for him in this matter.

'What does your uncle say to that, Snap?' asked the Admiral.

'Not much, sir, bad or good. He says I am an ungrateful young wretch for
refusing to go into Mr. Mathieson's office, and that I shall never come to any
good. But, then, I've heard that from him often enough before,' said Snap
grimly, 'and I think he will let me go, and that is the main point.'

'And when do you mean to start?' asked the Admiral.

'Oh, as soon as he will let me, sir. You see, my father left me a few
hundred pounds, so that I dare say when Mr. Hales sees that my mind is
made up he will let me go. You don't think much worse of me, I hope, sir, do
you?'

'Worse of you?' said the old sailor stoutly, 'no! You are a young fool, I
dare say, but so was I at your time of life. Come up to lunch!' And, planting
his rod by the side of the stream, he turned towards the house, Mrs.
Winthrop and Snap following him.

At lunch Snap had to tell the whole story again to Billy and Frank, but
when he came to the point at which he had decided to 'go west,' instead of
eliciting the sympathy of his audience, he only seemed to rouse their envy.

'By Jove,' said Frank, 'if it wasn't for this jolly old place I should wish that
I had got your character and your punishment, Snap!'

For a week or more both the Admiral and his sister had been very unlike
their old selves, so quiet were they and distrait, except when by an effort one
or the other seemed to rouse to a mood whose merriness had something false
and strained in it, even to the unobservant young eyes of the boys. Why was
it that at this speech of Frank's Mrs. Winthrop's sweet eyes filled with
sudden tears, and that piece of pickle went the wrong way and almost
choked the Admiral? Perhaps, if you follow the story further, you may be
able to guess.

After lunch they all wandered down again to the trout-stream, where
'Uncle's Ogden,' as they called the Admiral's rod, stood planted in the
ground, like the spear of some knight-err ant of old days. It was a lovely
spot, this home of the Winthrops—such a home as exists only in England;
beautiful by nature, beautiful by art, mellowed by age, and endeared to the
owners by centuries of happy memories. The sunlight loved it and lingered
about it in one moss-grown corner or another from the first glimpse of dawn
to the last red ray of sunset. The house had been built in a hollow, after the
unsanitary fashion of our forefathers; round it closed a rampart of low
wooded hills, which sheltered its grey gables from the winter winds; and in
front of it a close-cropped lawn ran from the open French windows of the
morning-room to the sunlit ripples of the little river Tane as it raced away to
the mill on the home-farm.

For five centuries the Winthrops had lived at Fairbury, not brilliantly,
perhaps, but happily and honestly, as squires who knew that their tenants'
interests and their own were identical; sometimes as soldiers who went away
to fight for the land they loved, only to come back to enjoy in it the honours
they had won. It was a fair home and a fair name, and so far, in five
centuries, none of the race had done anything to bring either into disrepute.
No wonder the Winthrops loved Fairbury.
THE ADMIRAL FISHING

But I am digressing, and must hark back to the Admiral, who has stolen
on in front of his followers and is now crouching, like an old tiger, a couple
of yards from the bank of the brook. Above him, waving to and fro almost
like that tiger's tail, is the graceful, gleaming fly-rod, with its long light line,
which looks in the summer air no thicker than gossamer threads. In front of
the old gentleman's position, and on the other side of the stream, is a
crumbling stone wall, and for a foot or two from it, between it and the
Admiral, the water glides by in shadow. Had you watched it very carefully,
you might, if you were a fisherman, have detected a still, small rise, so small
that it hardly looked like a rise at all. Surely none but the most experienced
would have guessed that it was the rise of the largest fish in that stream. But
the Admiral was 'very experienced,' and knew almost how many spots there
were on the deep, broad sides of the four-pounder whose luncheon of tiny
half-drowned duns was disturbing the waters opposite. At last the fly was
dry enough to please him, and Admiral Chris let it go. A score of times
before, in the last few days, he had had just as good a chance of beguiling his
victim, and each time his cast had been light and true, so that the harshest of
critics or most jealous of rivals (the same thing, you know) could have found
no fault in it. Each time the fly, dry as a bone and light as thistle-down, had
lit upon the stream just the right distance above the feeding fish, and had
sailed over him with jaunty wings well cocked, so close an imitation of
nature that the man who made it could hardly have picked it out from among
the dozen live flies which sailed by with it. But a man's eyes are no match
for a fish's, and the old 'sock-dollager' had noticed something wrong—a
shade of colour, a minute mistake in form, or something too delicate even for
Ogden's fingers to set right—and had forthwith declined to be tempted. But
this time fate was against the gallant fish. The Admiral had miscalculated his
cast, and the little dun hit hard against the crumbling wall and tumbled back
from it into the water 'anyhow.'

Though a mistake, it was the most deadly cast the Admiral could have
made. A score of flies had fallen in the same helpless fashion from that wall
in the last half-hour, and as each fell the great fish had risen and sucked them
down. This fell right into his mouth. He saw no gleam of gut in the
treacherous shadow, he had seen no upright figure on the bank for an hour
and a half; he had no time to scrutinise the fly as it sailed down to him, so he
turned like a thought in a quick brain, caught the fly, and knew that he too
was caught, almost before the Admiral had had time to realise that he had for
once made a bad cast. And then the struggle began; and such is the injustice
of man's nature that even gentle Mrs. Winthrop did not feel a touch of
compassion for that gallant little trout, battling for his life against a man who
weighed fifteen stone to his four pounds, and had had as many years to learn
wisdom in, almost, as the fish had lived weeks. No doubt she would have
felt sorry for the fish if she had thought of these things, but then you see she
didn't think of them.

'By George! I'm into him,' shouted the Admiral.

Anyone only slightly acquainted with our sporting idioms might have
taken this speech literally, and wondered how such a very small whale could
have held such a very large Jonah. But the Admiral never stopped to pick his
words when excited, as poor Billy soon discovered. An evil fate had
prompted Billy to snatch up the net as soon as his uncle struck his fish, and
now, as the four-pounder darted down stream, the boy made a dab at him
with it.

'Ah, you young owl! You lubberly young sea-cook,' roared the infuriated
old gentleman. 'What are you doing? Do you think you're going to take a
trout like a spoonful of porridge? Get below him, and wait till I steer him
into the net.'

Frightened by Towzer's futile 'dab,' the trout had made a desperate dash
for the further side of the stream, making the Admiral's reel screech as the
line ran out. Skilfully the old man humoured his victim, now giving him line,
now just balking him in his efforts to reach a weed-bed or a dangerous-
looking root. People talk of salmon which have taken a day to kill; it is a
good trout which gives the angler ten minutes' 'play.' The Admiral's trout
was tired even in less time than that, and came slowly swimming down past
a small island of water weeds, beyond the deep water on the house-side of
the stream, submissive now to his captor's guiding hand. Gently the Admiral
drew him towards the shallows, and in another moment he would have been
in the net, when suddenly, without warning, he gave his head one vicious
shake, and, leaping clear out of the water, fell back upon the little island,
where he lay high and dry, the red spots on his side gleaming in the sun. It
was his last effort for freedom, and now, as he lay gasping within a few
inches of the clear stream, of home and safety, the treacherous steel thing
dropped out of his mouth, the current caught the belly of the loose line and
floated it down stream, and the Admiral stood on the further bank dumb with
disgust, the last link broken which bound his fish to him. In a moment more
the fish would recover from his fall, and then one kick, however feeble,
would be enough to roll him back into the Tane, and so good-bye to all the
fruits of several weeks' patience and cunning, and good-bye, too, to all
chance of catching 'the best trout, by George, sir, in the brook!' It was hard!

But there was another chance in the Admiral's favour which he had not
counted upon. Even as the fish fell back upon the dry weeds Snap slid
quietly as an otter into the stream. A few strong, silent strokes, and he was
alongside the weeds, and as the fish's gaping gills opened before he made
what would have been to the Admiral a fatal effort Snap's fingers were
inserted, and the great trout carried off through his own element as
unceremoniously as if it really was an otter which had got him.
'I'm not a bad retriever, sir, am I?' asked the hoy as he laid his prize down
at old Winthrop's feet. That worthy sportsman was delighted.

'No, my boy,' he replied, 'you are first-rate, though perhaps Mr. Hales
would call you a sad dog if he saw you in those dripping garments. Be off
and change into some of Frank's toggery.'

'All right, sir; come on, Frank,' replied Snap, and together the three boys
raced off to their own domain in one of the wings of Fairbury Court, given
over long ago to boys, dogs, and disorder.

Meanwhile the Admiral retired to weigh his fish, which he did most
carefully, allowing three ounces for its loss of weight since landing—an
altogether unnecessary concession, as it had not been out of the water then
more than five minutes. However, he entered it in his fishing journal as 3 lbs.
11 ozs., caught August 2, and retrieved by Snap Hales. As he closed the
book he sighed and muttered, 'That is about the last trout I shall take on the
Tane.'

CHAPTER VI

THE BLOW FALLS

The day after the Admiral's triumph over his fishy tenant he and his sister
called a meeting in the morning-room after breakfast. It was an informal
meeting, but, as he said that the business to be done was important, the
young squire restrained his impatience to go and see the men about rolling
the cricket pitch in the park, and waited to hear what his uncle had to say.

'I'm sorry, Frank,' the old man said, 'that you will have to put off "the
Magpies" for next week, but I am afraid we can't have any cricket here this
August.'
'Why, uncle,' expostulated Frank, 'it is the very best fun we have, and the
Magpies are capital good fellows as well as good cricketers.'

'Yes, I know,' replied his uncle gravely, 'but even cricket must give way
sometimes, and now it happens that your mother and I are suddenly called
away on business, on very important business,' and here he looked sternly at
his sister-in-law, who turned her face from the light, and appeared to busy
herself with the arrangement of a vase of flowers on the old oak over-mantel.

'But, uncle,' put in Towzer, 'couldn't Frank take care of the Magpies even
if you and mother were not here? Of course it would not be half such fun as
if you were here to score and Mother to look on, but Humphreys (the butler)
would see that the dinners were all right. I'm sure he could,' added the boy
more confidently, catching at a sign of approval in his brother's face.

'It wouldn't do, my boy,' asserted Admiral Chris, 'it would not do at all; it
would be rude to your guests, you wouldn't be able to manage, and besides,'
he added, as if in despair for a convincing argument, 'we might be able to get
back, and then neither your mother nor I need miss the match.'

This was quite another story, and so the boys consented, albeit with a
very bad grace, to postpone their cricket.

'What I propose now instead of the match,' continued the Admiral, 'is a
little travel for you two, and I've asked Snap Hales's uncle to let him go with
you. I want you to go off and try a fishing tour in Wales, whilst your mother
and I finish our business in London, and then we'll all meet again in a
fortnight's time.'

'Bravo, uncle!' cried Frank, 'but what am I to do for a rod?'

'Oh, if yours is broken you had better take mine,' replied the Admiral.

'What, your big Castle Connel? Thank you, sir; it would be as much good
to me in such cramped places as you used to tell us about as a clothes-prop!'
replied Frank.

'No, not the Castle Connel, the Ogden; I shan't want it, and you will take
care of it, I know,' was the unexpected reply.
'Your Ogden, sir!' said Frank; 'why, I thought no one might look at it from
less than ten paces.'

'You're an impertinent young monkey, Frank,' laughed the Admiral, 'but


still you may have it.'

And so it was settled that the Magpie match should be given up, and
Frank and Billy be packed off on a fishing tour in Wales, whilst their mother
and the Admiral went up to town and transacted the troublesome business
which had had the bad taste to demand their attention during the Midsummer
holidays.

A little later in the day a man came up from the village with a note from
Mr. Hales for the Admiral. The boys did not see it, but it was understood to
contain his consent in writing 'to the proposal that Snap should join the
expedition.' For the rest of that day all was excitement and bustling
preparations for a start. It seemed almost as if they were preparing for
something much more important than a fortnight's trip into Wales. Snap was
up at the house all day. That with him was common enough. His own
packing had not taken him long. The boy was keener-eyed than his young
companions, and, in spite of an apparent roughness, was more sensitive to
external influences than either of them. Hence it was perhaps that he noticed
what they overlooked; noticed that Mrs. Winthrop's eyes followed her sons
about from room to room, that she seemed to dread to lose them from her
sight, that the dinner that night was what the boys called a birthday dinner,
that is, consisted of all the little dishes of which Mrs. Winthrop knew each
boy was specially fond, and what struck him more than anything was that
two or three times he was sure her eyes filled with tears at some chance
remark of Frank's or Willy's which to him had no sad meaning in it. He was
puzzled, and, worse than that, 'depressed.'

The start next morning was even less auspicious than 'packing-day' had
been. The midsummer weather seemed to have gone, and the gables of the
old house showed through a grey and rainy sky; rain knocked the leaves off
the roses, and battered angrily at the window-panes. The pretty Tane was
swollen and mud-coloured, and, altogether, leaving home on a fishing trip to
Wales felt worse than leaving home the first time for school.
The Admiral had determined on seeing them on their way as far as the
county town, and drove to the station with them in the morning. If it had not
been so absolutely absurd, Snap would have fancied once or twice that the
old gentleman did not like any of the boys to be alone with his neighbours,
or even with the servants. It would have been very unlike him if it had been
the case, so of course Snap was mistaken.

'Towzer,' asked Frank in a whisper as they drove away, 'what was the
Mater crying about?'

The Admiral overheard him, and replied:

'Crying, what nonsense, Frank; your mother was waving good-bye and
good riddance to you with that foolish scrap of lace of hers; that's all.
Crying, indeed!' and the old seaman snorted indignantly at the idea.

It was all very well for the Admiral to deny the fact, and to go very near
to getting angry about it, but Snap at any rate knew that it was a fact, and
that Admiral Chris knew it too. It was the first untruth he had ever heard
from the upright old gentleman by his side, and Snap's wonder and dislike to
this journey grew. As Snap looked back a turn in the road gave him another
rain-blurred glimpse of Fairbury, with a little drooping figure which still
watched from the Hall steps, and a conviction that something was wrong
somewhere forced itself insensibly upon him, though as yet he was not wise
enough to guess where the evil threatened.

The rain had an angry sound in it, unlike the merry splash of heavy
summer showers: there seemed a sorrow in the sigh of the wind, unlike the
scent-laden sigh of summer breezes after rain. Nature looked ugly and
unhappy, and the boys were soon glad to curl themselves up in their
respective corners of the railway carriage, with their backs resolutely turned
upon the rain-blurred panes of the carriage window.

At the station the Admiral had met his favourite aversion, Mr. Crombie.
What Mr. Crombie had originally done to offend the Admiral no one knew,
but he had done it effectually. Crombie gave Admiral Chris the gout even
worse than '47 port or the east wind.
Crombie was on the point of addressing Frank when the Admiral
intervened and carried off the boys to get tickets. A little to Frank's surprise,
his uncle took third-class tickets, for, although on long journeys the old
gentleman invariably practised this wise economy, Frank had been
accustomed to hear him say, 'Always take "firsts" on our own line, to support
a local institution.'

As the Admiral took his tickets the voice of his persecutor sounded
behind him. Crombie had followed his foe.

'What!' he said—and the sneering tone was so marked that it made the
boys wince—'an Admiral travelling third!'

'Yes, Sir,' retorted the Admiral fiercely. 'God bless me, you don't mean to
say there is a "fourth" on? Only persons who are afraid of being mistaken for
their butlers travel first nowadays,' and with an indignant snort the old
gentleman squared his shoulders, poked out his chin, and walked down the
platform with a regular quarter-deck roll, leaving Mr. Crombie to meditate
on what he was pleased to call 'the "side" of them beggarly aristocrats.'

At Glowsbury, the county town for Fairbury, Admiral Chris left the boys,
hurrying away with an old crony of his, who, in spite of nods and winks,
would blurt out, 'I'm so sorry, Winthrop.' But the Admiral let him get no
further. 'Good-bye, lads,' he sang out, and then away he trotted, holding on to
his astonished friend, whom he rapidly hustled out of earshot, so that the
boys never knew the cause of that old gentleman's sorrow.

It didn't trouble them much either, for, once in Wales, the weather grew
fine again—provokingly fine, the boys thought. If ever you go to dear little
Wales, O Transatlantic cousin, to see the view, you may bet your bottom
dollar that you won't see it. You will be like that other tourist who 'viewed
the mist, but missed the view.' If, however, you can jockey the Welsh climate
into a belief that you are going there solely for fishing, you may rely on such
weather as the Winthrops got, that is to say, clear skies, broiling suns, and
tiny silver streams calling out for rain-storms to swell their diminished
waters, and crying out in vain. The waters will be clearer than crystal, the
fish more shy than a boy of fourteen amongst ladies, and the views perfect.
Unfortunately, it is extremely difficult to jockey anything Welsh: Wales is
very unbelieving, and especially does it disbelieve anglers.

The boys opened their campaign on the Welsh borders, fished


successfully for samlets—bright, silvery little fellows, which had to be put
back—and with a miserable want of success for the brown trout, which they
were allowed to keep if they could catch them. Sometimes they walked from
point to point, but then they found that their expenses in gingerbeer were
almost as great as if they had spent the money in a third-class ticket; once
they tried a long run by rail on the—well, I dare not tell you its real name—
so I'll say the Grand Old Dawdler's line. They bought third-class tickets, but
travelled first, because the line had only three coaches in at that time, and
they were all first. Two rustics travelled with them; it was rather a busy day
with the Grand Old Dawdler's line. The station-master at the starting-point,
who sold them their tickets, went with them as engine-driver and guard, and
at each of the little stations which they passed he acted as station-master.
This system of centralising all the service in one person had its advantages:
there is only one person to tip, and if he is sober the travelling, if slow (say
seven miles an hour), is very fairly safe.

Once, and once only, they tried tricycles. Wales is not as level as a
billiard-table. Towzer, careless of the picturesque, wished that it was. On
tricycles, he explained, if you were not used to them, you could travel on the
flat rather faster than you could walk; uphill you had to get off and shove,
and downhill you were either run away with, or, if you put on the brake, the
tricycle stopped, you didn't—on the contrary, you proceeded upon your
journey by a series of gyrations through the air, until suddenly planted on
your head in the next county but two. Besides all this it cost more to send
back your tricycle by rail than a first-class ticket would have cost, whereas if
you didn't send it back you were liable to be tried at the next assizes.

A letter which I insert here, and which Mrs. Winthrop still keeps, for the
sake, not of its melodious metre, but for the sake of auld lang syne, will give
the reader some idea of the Winthrops' fishing adventures. I am inclined to
think that Frank wrote it. Big, strong fellow as he was, he had a habit of
constantly writing to the Mater, and I happen to know that Snap was too bad-
tempered at that time to write anything. He had passed all that morning in
trying to cast on a certain wooded reach. He had caught the grass; he had
cracked his line like a coach-whip, and lost a score of flies by so doing; and
had at last settled solemnly down to dig up with his penknife a great furze-
bush on the bank which appeared to his angry imagination to rise from
behind at every fly which he tried to throw.

'Aug. 12, 1874.

'DEAR MATER,—-

'Snap Hales arose, from his night's repose,


In the midst of the Cambrian mountains,
Where from cliff and from crag, over peat-moss and hag,
The Tanat shepherds her fountains.

(Observe here the resemblance to Shelley.)

'He rolled in his tub, and tackled his grub,


He booted and hatted in haste,
Then said, "If you're wishing, boy Bill, to go fishing,
There isn't one moment to waste."

'He strode to the brook, and with lordly look


Quoth, "Now, little fish, if you're in,
Let some grayling or trout just put up his snout
And swallow this minnow of tin."

'As if at his wish, up bounded a fish,


Gave one dubious sniffle or snuff,
Thought 'It's covered with paint, I'll be hooked if it ain't,
And the fellow who made it's a muff."

'Then Harold had tries with all sorts of flies,


Which were brilliant, gigantic, and rare,
But among them were none which resembled a "dun,"
So the fish were content with a stare.
'To a tree by that brook many flies took their hook,
Many more were whipped off in the wind;
One fixed in the nose, several more in the clothes
Of that angler before and behind.

'Then his cast-line broke, and Harold spoke,


Right wrathful words spoke he,
"Very well! you may grin, but I'll just wade in
Where there's neither briar nor tree."

'With naked foot, without stocking or boot,


Right into the stream he strode—
With a splash and a splutter, with a murmur and a mutter,
And he frequently "Ah'd" and "Oh'd."

'Alas, as he tripped his bare feet slipped,


They slipped on those slimy stones,
And down he came (I forget the name
Of the very identical bones

'Upon which he sat); but he'd flies in his hat,


And as he went down the stream
The fish arose, and tugged at his clothes,
Until he began to scream.

'Round his hat's broad brim they began to swim,


And into his face did stare.
His mouth they eyed, they peeped inside,
Much wondering who lived there.

'Their victim cried, "In vain I've tried


To snare these fishes free.
Alas, for my sin, as they've got me in,
I fear they'll swallow me."

'But, "Alack, this Jonah's a fourteen-stoner,"


'Twas thus that the fishes cried.
"If we gape till we split, there will still be a bit
Of the monster left outside."

'So Will landed him safe, our fisherman waif,


In safety he landed him;
With gobble and munch he chawed up his lunch,
He was hungry after his swim.

'He has sworn he will never again endeavour


Those innocent fish to hurt,
For all he can get is thundering wet,
And any amount of dirt.

'Your truthful
'FRANK.'

After this, perhaps, it is not surprising that the boys voted fishing very
poor fun, and took to mountaineering instead. They had climbed Cader Idris
(a very pretty climb from its more difficult side) and Snowdon, and were
resting at a first-rate hotel not far from Snowdon's foot, when they found the
following letter on their breakfast-table from the Admiral:—

'DEAR FRANK,—As your mother is not very well, I intend to bring her
down to Dolgelly for a few days. Take some nice quiet rooms where we can
all be lodged together at less expense than at an hotel.

'Your affect. uncle,


'CHRISTOPHER WINTHROP.

'P.S.—I have some important news to give you, and should like you all to
be at home when we arrive by the 12.50 train to-morrow.'
Frank read the letter out to the rest at breakfast, and then laid it quietly
down by his plate.

'Snap,' he said, 'there is something wrong at home. I can't make out what
the Admiral is always harping on economy for. Surely our mother' (and
unconsciously there was a tone of pride in that 'our mother') 'can afford to go
to any of these wretched little hotels if she likes. I shan't take rooms. It's all
nonsense; I'm not going to have her murdered by Welsh cooks, especially if
she is ill.'

No one having any explanation of the Admiral's letter to offer, or any


objection to staying where they were, the conversation dropped, but the boys
were restless and unhappy until the 12.50 train was clue in.

When that train pulled up with a jerk at the platform the three had already
been waiting for it half an hour, for their impatience had made them early,
and long habit had made the train late. As soon as they could find their
mother and Admiral Chris the boys pounced upon them, and in the first burst
of eager welcome the cloud vanished. But it reappeared again before the
party reached the hotel, and the Admiral was as nearly angry as he knew
how to be on finding that the rooms taken for himself and sister were, as
usual, just the best in the hotel.

The dinner was a poor and spiritless affair, and Snap noticed that the old
gentleman, instead of lighting a cigar after leaving the table, took at once to
a pipe.

'Why, sir,' remonstrated Snap, 'you are false to your principles for the first
time in my experience of you; I thought that you always told us that the cigar
was a necessary appetiser, to be taken before the solid comfort of the
evening pipe.'

'Nonsense, my boy, nonsense, I never said that. A cigar is a poor thing at


best. Nothing like a pipe for a sailor,' blurted out the Admiral, looking
annoyed at Snap's innocent speech, and glancing nervously in Mrs.
Winthrop's direction, while over her sweet face a cloud passed as she too
noticed for the first time this little change in her brother-in-law's habits.
Coming up to her eldest boy's chair, and leaning caressingly against him,
the little mother turned Frank's head towards her, so that she could look
down into his honest blue eyes.

'What is it, little mother; do you want a kiss in public? For shame, dear!'
laughed Frank.

'Tell me, Frank,' she said, taking no notice of his chaff, 'do you want very
much to go to Oxford?'

'Right away, mother? No, thank you. I am doing very well here.'

'But when you leave Fernhall, Frank?'

'Well, yes, mother! You wouldn't have me go to Cambridge, because, you


see, all my own friends are at the Nose,' replied Frank.

'The Nose?' asked Mrs. Winthrop, looking puzzled.

'Brazen Nose, dear, Brazen Nose!—the college, you know, at which Dick
and the Rector's son now are.'

'But what should you say, Frank, if you could not go either to Oxford or
Cambridge?' persisted his mother.

'Conundrum, mother. I give it up,' answered the boy lazily; 'call me early,
dear, to-morrow, and ask me an easier one.'

Poor little lady, the tears came into her eyes as the smile grew in his, and
at last Frank saw it. Jumping up and putting his arm round her, he asked:

'Why, mother dear, what is it? I was joking. I'll go anywhere you want.'

'Yes, my boy, I know,' sobbed the little woman, 'but you can't go either to
Oxford or Cambridge. There, Chris, tell them the rest,' and, slipping out of
Frank's arms, she left the room.

After this beginning the whole story was soon told. The Admiral's pipe
had gone out, his collar seemed to be choking him, but, now that he was
fairly cornered, he didn't flinch any longer.

'Yes!' he said, 'that is about the truth of it. We are all ruined. Fairbury was
sold three days after you left it. That is why we sent you down here. We
wanted to spare Frank the wrench, and we didn't want any of you punching
the auctioneer's head, or any nonsense of that sort. We have all got to work
now, lads, for our living.'

Here the old man rose and put his strong hands on Frank's shoulders, and
looked him full in the face.

'With God with them, my boys aren't afraid, are they?'

Frank gripped the old man's hand, and Billy crept up close to him, while
Snap, watching from a distance, felt hurt to the heart that he had not lost and
was not privileged to suffer with them.

And yet 'Fairbury sold' seemed too much for any of them to realise all at
once. Fairbury seemed part and parcel of themselves. It was to them as its
shell to an oyster. The Winthrops (the whole race) had been born in it, and it
had grown as they grew. After a while Towzer broke the silence.

'Then, uncle, where are we going home to?' he asked.

'Home, my lad! Well, I suppose we must make a new home somewhere. It


should not be difficult at our age, should it, Frank?' added the gallant old
man, as if he were the youngest of the young as well as the bravest of the
brave.

'But, uncle, won't mother's tenants pay their rent?' asked Frank.

'My boy, your mother has no tenants,' said Mrs. Winthrop, who had re-
entered the room, 'and you'll never be Squire of Fairbury, as you should have
been. It does seem hard.'

And so it did, and one young heart, of no kin to hers, felt it almost as
much as she did, and Snap swore then, though it seemed a ludicrous thing
even to himself, that, if ever he could, he would put back that sweet woman
and her boy in their own old home.
But I must hurry over this part of my story. Sorrow and tears are only
valuable for the effects they leave behind. Without the rain there would be
no corn; without misfortune and poverty there would be very little effort and
achievement in the world. But it is more pleasant to dwell on the happy
results than on the causes.

When Frank had insisted on seeing his mother to her bedroom, with a
quaint assumption of authority which she never resisted, the Admiral
explained how all their troubles had arisen. A friend to whom Mrs. Winthrop
had lent 500l. had repaid that sum to her agent in Scotland. The agent (a
lawyer), acting on the Admiral's instructions with regard to small sums paid
in the absence of Mrs. Winthrop on the Continent, had invested the 500l. in
some bank shares. The shares were bought, he believed, much under their
value. Alas, the public knew better than that lawyer. The bank was an
unlimited affair, and broke soon after he had bought its shares, and Mrs.
Winthrop became responsible for the payment of its debts to the last penny
which she possessed. Without any fault of theirs, without warning, the
Winthrops had to give up their all. This is one of the dangers of civilised life,
and, unfortunately, company promoters, swindling bankers, and such like are
not yet allowed to hang for their sins.

Luckily, the Admiral was not involved in the general ruin, and was as
staunch and true as his kind generally are in the time of trouble. 'My dear,' he
had said to his sister, when he had finished abusing the bank, the bankers,
the Government, and every person or thing directly or indirectly connected
with banking, 'it was my fault for not looking after the money myself.
Nonsense! of course it was. What should a poor devil of a lawyer know
about banking, or law, or anything except bills? However,' he added more
calmly, 'there is my little property and pension for you and the boys, and, as
for me, I dare say that I can get a secretaryship to a club or something of that
sort in town.'

The Admiral had a hazy idea that the letters E.N. behind his name were
sufficient qualification and testimonial for any public office, from the
directorship of a guinea-pig company to the secretaryship of the Royal
Geographical Society.
'And now, lads,' he was saying an hour after Mrs. Winthrop had retired
for the night, 'think it all well over. There is a stool in an office for one of
you, if you like. No place like the City for making money in; or, if you don't
like that, Frank, we can find money enough somehow to send you to the Bar.
We have employed attornies enough in our time, and of course some of them
would send you briefs enough to give you a start' (would they? poor
Admiral!); 'or there is young Sumner's craze—cattle-ranching or farming in
the far West—a rough life, no doubt, but—— Ah, well, it's not for me to
choose. I'm not beginning life. I wish I was—as a cowboy,' and the old man
picked up his candle and trotted off to bed with almost enough fire in voice
and eye to persuade you that he was still young enough to begin another
round with Fate.

That night the boys sat up on the edges of their beds until long after
midnight, talking things over. Frank was very grave, and inclined to
persuade his younger brother to take to the office-stool.

'And you, Major?' asked Towzer; 'are you going to the Bar?'

'Well, no,' replied his brother, 'I don't think that I could stand being buried
alive in those dim, musty chambers yet, and I've no ambition to conquer
Fortune with the jawbone of an ass.'

'Very well, then, if you won't set me an example, let's drop London and
talk cattle-ranching,' said Towzer. 'Snap, you've got an old "Field" in your
bag, haven't you?'

'Yes, here you are,' replied the person addressed, producing an old copy
of that one good paper from his portmanteau.

'Look in the advertisement-sheet,' suggested Frank, 'there is always


something about ranching there.'

'"Expedition to Spitzbergen,"' read Snap; 'that won't do. "Wanted another


gun to join a party going to the Zambesi." Ah, here you are, "Employment
for gentlemen's sons. The advertiser, who has been settled at Oxloops, on the
north fork of the Stinking Water, for the last ten years, is prepared to receive
two or more sons of gentlemen upon his ranche, and instruct them in the

You might also like