Debashis Mukherji - Electron Microscopy - A Versatile Tool For Material Characterization-Bookboon Learning (2017)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 202

Debashis Mukherji

Electron Microscopy
A Versatile Tool for Material Characterization


DEBASHIS MUKHERJI

ELECTRON
MICROSCOPY
A VERSATILE TOOL FOR
MATERIAL CHARACTERIZATION

2
Electron Microscopy: A Versatile Tool for Material Characterization
1st edition
© 2017 Debashis Mukherji & bookboon.com
ISBN 978-87-403-1696-4
Peer reviewed by Prof. Dr. Gerhard Schumacher Department of Microstructure and Residual
Stress Analysis – Helmholtz-Zentrum Berlin & Dr Partha Ghosal, FEMSI

3
ELECTRON MICROSCOPY Contents

CONTENTS
Acknowledgements 6

Foreword 7

1 Introduction 9
References 20

2 Historical Timelines 22

3 Electron Beam Interaction with Materials 25


3.1 Electron Scattering 25
3.2 Electron Diffraction 37
3.3 X-ray Generation 40
3.4 References 43

Free eBook on
Learning & Development
By the Chief Learning Officer of McKinsey

Download Now

4
ELECTRON MICROSCOPY Contents

4 Instrumentation 44
4.1 Electron Sources 45
4.2 Electron Optics 52
4.3 Electron and X-ray Detectors 67
4.4 References 92

5 Microscope Types and Operating Modes 95


5.1 Transmission Electron Microscope 96
5.2 Scanning Electron Microscope 100
5.3 Operating Modes in TEM and SEM 104
5.4 References 107

6 Microscopy and Analysis 108


6.1 Imaging 109
6.2 Diffraction 161
6.3 Analysis 174
6.4 References 190

Endnotes 196

About the author 199

About the reviewer: Prof. Dr. Gerhard Schumacher 200

About the reviewer: Dr Partha Ghosal 201

Review by Dr Partha Ghosal 202

5
ELECTRON MICROSCOPY Acknowledgements

ACKNOWLEDGEMENTS
This book is a result of my long association with electron microscopes and microscopy
as a technique I used throughout my career as a material scientist. In this long journey, I
learned from many excellent microscopists from around the world, with whom I came in
contact at various times. The idea of the publisher Bookboon, to provide free access to the
ebook to students throughout the world appealed me very much and therefore I took up
the challenge of writing this book.

I must admit that although I have been using different kinds of electron microscopes and
electron microscopic methods in my scientific career, I found that writing the book was
not a simple task. In the process, I myself learned new things and got my knowledge on
many topics revised. I hope the end effect will be useful for students of material science,
physics and chemistry.

The names of two particular persons, both my old colleagues needs special mention for
their support and encouragements. I am truly indebt to them. I am grateful to Dr. Gerhard
Schumacher, from the Helmholz Zentrum, and a professor at the Technische Universität
Berlin, who is a friend and colleague from Germany. I thank Gerhard for taking his time,
despite his very busy schedule to proof read the book. We have worked together, on TEM
and Ni-base superalloys and Ti-alloys at Berlin. Some of his suggestions, I used in structuring
the book and to include topics like HAADF, in the book. His careful reading and useful
comments were invaluable.

The other person, Dr Partha Ghosal, is a friend and colleague from India. He is now the
head of the electron microscopy group at the Defence Metallurgical Research Laboratory,
Hyderabad, India, where I actually first learned to use a TEM. Partha provided me all the
experimental data on EELS that I have included in this book. I have used many different
electron microscopic techniques in my career and most of the experimental results in
the book is from my own work. However, I have never done EELS analysis on my own.
Therefore, I am indebted to Partha, for his contribution to the book, without which this
book would not have been complete.

I thank my family, particularly my wife Arundhati for her patience, support and encouragement.

6
ELECTRON MICROSCOPY Foreword

FOREWORD
The light microscope opened the first door to the microcosm, the electron microscope
opened the second door, what will we see when we open the third door?

– E. Ruska 1985

The two quotations in these page (see boxes) sum up the importance of electron microscopy
in modern science. One is from the Nobel Laureate Ernst Ruska (who got the Nobel Prize
in 1986 for the invention of Transmission Electron Microscope (TEM) in the early 1930s)
and the second one is from David Williams and Barry Carter (the authors of one of the
frequently referred book on TEM). Microscopes enhance our vision by allowing us to look
at things that are far too small to be visible with the naked eye. They do this by making
things appear bigger (magnifying them) and at the same time increasing the amount of
detail we can see (resolve them). Therefore, magnification and resolution are two important
concepts in microscopy and will be discussed in this book. Further, it is important to
realize that the fundamental principles of light optics are also the basis for understanding
the advantages and the limitations of electron microscopy. In electron microscope, an
electron beam is used as the illumination source, replacing light in the microscope. As the
wavelength of electron beam is orders of magnitude smaller than the wavelength of light,
it allows increasing the point-to-point resolution in the microscope. The wavelength-limited
resolution in electron microscope is ~0.1 nm, which is much smaller than the limits in light
microscopes (resolution limit ~250 nm).

This textbook companion will explain the main theoretical concepts of how an electron
microscope magnifies and resolves an object. It will further describe the main components
of the instrument: the electron source, the electron optics and the electron detectors, which
can be of various types. This in conjunction with how electrons interact with the sample
will let us understand the principles of electron microscopy. Electron microscopy is a most
valuable tool for the study of materials on a micro- and nano-scale. The versatility of electron
microscope is a result of the multitude of signals that are generated when a high-energy
electron beam interacts with mater. Electron beam focused on a sample surface may be
transmitted through the material if the sample is thin or will be scattered from the surface
in bulk samples, so two different types of microscopes have been developed: namely the
transmission electron microscope and the scanning electron microscope (SEM). These two
microscope types provide complimentary information for material characterization and they
have many similarities but also have major differences. This book will discuss the similarities
and the differences in SEM and TEM.

7
ELECTRON MICROSCOPY Foreword

“No other scientific instrument exists which can offer such a broad range of characterization
techniques with such high spatial and analytical resolution, coupled with a completely
quantitative understanding of the various techniques.”

– David B. Williams & C. Barry Carter

Transmission Electron Microscopy, Textbook for Materials Science, Springer 2009

The versatility of electron microscope also stems from the type of information it can
provide in the characterization of materials. In addition to imaging, which is the common
feature of any microscope, electron microscope also provides crystallographic and chemical
information of the sample. Examples about these are provided in this book, which will relate
mainly to metallic materials. However, the electron microscopes are used for characterizing
a variety of materials, e.g. metals, ceramic, polymer, biological material, etc. Soft matters
like biological materials or polymers are, however, sensitive to damage by the electron beam.
This book is mainly meant for the engineering students studying structural materials and
will not look into these aspects of damages. Furthermore, sample preparation for electron
microscopy is very important for good microscopic results, and that of biological material
is quite different from metallic material. Sample preparation will not be addressed in this
book. In fact, the book only covers crystalline materials and their interaction with the
electron beam and does not deal with soft matter, amorphous materials or quasi-crystals.
In crystalline samples, TEM can directly image the atomic structure in the sample. High-
resolution electron microscopy (HREM) will be discussed. Aberration in the electro-magnetic
lenses and the coherency of the electron beam play an important role in high-resolution
imaging. New development in electron microscopy focuses on how to improve the practical
resolution limits of the microscopes and focuses on the corrections of the lens aberrations.
However, new techniques of scanning transmission microscopy (STEM), convergent beam
electron diffraction (CBED) and high angle annular diffraction (HAADF) imaging are
equally important in contrast development and resolution and will be discussed. Finally,
diffraction and microanalysis are useful aspects of electron microscopy, which provide
valuable information on the crystallography and chemical compositions of materials. These
techniques and their underlying principles will be covered.

I hope this book will help the graduate and undergraduate students to get a comprehensive
understanding about the electron microscopy. Not only about how they function and
about the principles behind, but also how they can be used effectively in characterization
of materials.

8
ELECTRON MICROSCOPY Introduction

1 INTRODUCTION
In this chapter you will learn about

-- Magnification and resolution


-- Airy pattern and the limits of resolution
-- Köhler illumination, Abbe’s equation and numerical aperture
-- De Broglie’s equation and wavelength

What is a microscope?
A microscope (Greek: mikrós, “small” and skopeîn, “to look or see”) is an instrument used
for viewing objects that are too small for the naked eye. In other words, it is an instrument
to make things look bigger – that is, it magnifies an object that we are looking at.

In the ancient times (~2000 BC) the Chinese used water microscopes, made of a lens with
tube filled with water to visualize the unseen [1.1]. A water droplet on a leaf as seen in
figure 1.1 is a simple example of this kind of lens. In his publication “The Invention of
the Microscope”, Bardell reported that around 612 BC the Assyrians manufactured the
world’s oldest surviving lenses [1.1]. Nearly 2000 years later in the 14th century, the art of
grinding glass lenses was developed in Italy and spectacles were made to improve eyesight
[1.2]. However, just the simple lens can magnify only to a limited extent. In the medieval
times, the first compound lens was developed in Europe.

Fig. 1.1: A simple water lens.

9
ELECTRON MICROSCOPY Introduction

It is difficult to say who actually invented the compound microscope. Some credit the
Dutch spectacle-maker Zacharias Janssen to have invented it in 1590 [1.3]. Another claim
for the invention is from Janssen’s competitor, Hans Lippershey [1.3]. The famous Italian
scientist, Galileo Galilei soon improved the compound microscope design with a convex and
a concave lens in 1609 and called the device an Occhiolino, or “little eye” [1.4]. Like the
first microscopes from the mediaeval time, modern optical microscopes still use glass lenses.

Magnification and resolution


When we talk about how microscopes work, we often say that they make things look bigger –
however, making things bigger is only part of the story. Figure 1.2a shows an image of a
wood surface texture, the central portion of this image (marked by the red box) is simply
magnified 12 times in figure 1.2b. The resolution in the two images remains the same.
The magnified image is pixelated as it has no further information added on magnification
and therefore fails to show more detail in the image. If microscopes did nothing but make
what we can already see, only bigger, they probably would not be of much use! Fortunately,
microscopes make things appear bigger and at the same time it increase the amount of
detail we can see. Another word for the level of detail we can see is resolution (increasing
our ability to distinguish between two objects or ‘resolve’ them).

a. b.
Fig. 1.2: a) An image of a wood surface texture, b) the area marked by red box in 2a is magnified 12 times.

10
ELECTRON MICROSCOPY Introduction

Understanding the limits of resolution


Resolution can be defined as the ability to differentiate two closely placed objects as distinct
objects rather than just one. The naked eye can differentiate (resolve) two objects such as
the grains of sand or human hair that are about a tenth of a millimeter (~100 micrometer)
in size. We see two small objects placed any closer than 100 µm as a single shape and
will need a microscope to resolve them apart. Different kinds of microscopes can show us
different amounts of detail (i.e. they have different resolving power). The resolution limits
are well defined in physics and the knowledge of several of the fundamental principles of
light optics is needed in order to understand the limitations. The wavelength of the optical
microscope’s (or the eye’s) illumination source – i.e. the light, defines a fundamental limit
to the resolution. Understanding the limits of light microscopy led to the development of
the electron microscope. Electron microscopes have a far greater resolving power than light
optical microscopes.

Principles of light optics


The idea that light has a dual nature – wave and particle, is quite old. By the turn of the
20th century, physicists were convinced that most commonly observed phenomena of light
can be explained by the wave theory, but the photoelectric effect suggested a particle nature
for light. The current scientific theory accepts the concept of wave–particle duality and holds
that all particles (including light) also have a wave nature [1.5].

Diffraction
Diffraction refers to the phenomena that occur when a wave encounters an obstacle. In
the simplest term, diffraction can be considered as the slight bending of light when it
passes around the edge of an object (lens or an aperture constitutes an object). In classical
physics, the diffraction phenomenon is described as the interference of waves according to
the Huygens-Fresnel principle [1.6, 1.7]. This principle characterize the behavior when a
wave encounters an obstacle or a slit that is comparable in size to its wavelength – showing
that diffraction results when a wavefront is impeded by any object. Figure 1.3a shows an
example of how diffraction changes the wavefront in the presence of a small aperture. Notice
that this causes the parallel wavefront to emerge from the aperture as a spherical wavefront.
Diffraction processes are most prominent when the obstruction or gap (aperture) is about the
same size as the wavelength of the impinging wave and its effect gets less obvious as the size
of the aperture increases. On the other hand, interference is the phenomenon when waves
from two or more coherent sources superpose with one another producing a resultant wave.
Figure 1.3b shows a wavefront diagram of double-slit interference from a monochromatic
light source. This interference can be either constructive (when two wavefronts add up giving
more light) or destructive (when the wavefronts cancel each other leading to darkness).

11
ELECTRON MICROSCOPY Introduction

Light

Dark
S1

Light

Source
S2
Dark

Light

a b
Fig. 1.3: Diffraction and interference phenomenon.

One of the consequences of diffraction and interference is that it is impossible to achieve


absolute focus using any optical system that uses particles with wave-like properties.

Airy disc and resolution


The diffraction pattern resulting from a uniformly illuminated circular aperture has a bright
region in the center (Fig. 1.4a), known as the Airy disc. The diameter of this pattern is
related to the wavelength of the illuminating light and the size of the circular aperture. This
is an important factor since the Airy disc is the smallest point to which a beam of light can
be focused. Notice the primary, secondary and tertiary wavefronts generated in the Airy
pattern (Fig. 1.4b), which of course continue to emanate at higher orders, but their effect
on optical phenomena diminishes in importance with each higher order. George Biddell
Airy, wrote the first full theoretical treatment explaining the phenomenon in his article “On
the Diffraction of an Object-glass with Circular Aperture” published in 1835 [1.8]. Owing
to diffraction, the smallest point to which a lens or a mirror can focus a beam of light is
the size of the Airy disc (Fig. 1.4c). Even using a “perfect” optical system, a point of light
cannot be focused as a perfect dot. Instead, the image when viewed critically consists of
a disc composed of concentric circles with diminishing intensity (Fig. 1.4a). This is the
diffraction image of a point source in a clear aberration-free aperture. Its mathematical
description is the Point Spread Function (PSF), which expresses the normalized intensity
distribution of the point-source image (Fig. 1.4b).

12
ELECTRON MICROSCOPY Introduction

a.
D 

Intensity

Sin θ
b.

Airy’s
Disc

c.

Fig. 1.4: Airy’s disc.

The resolution can be perceived by the concept of Airy diffraction patterns. The size of
the Airy disc determines the limit of a microscope to resolve two close point sources. The
resolution of an optical instrument is dependent on the aperture of the lens or mirror and
the wavelength of the light observed. However, it is independent of the focal length or the
magnification. As the diameter of the aperture increases, the size of the Airy disc decreases
(Fig. 1.5).

13
ELECTRON MICROSCOPY Introduction

NA=0.3

NA=0.5

NA=0.85

Fig. 1.5: Numerical aperture and Airy’s disc size.

www.sylvania.com

We do not reinvent
the wheel we reinvent
light.
Fascinating lighting offers an infinite spectrum of
possibilities: Innovative technologies and new
markets provide both opportunities and challenges.
An environment in which your expertise is in high
demand. Enjoy the supportive working atmosphere
within our global group and benefit from international
career paths. Implement sustainable ideas in close
cooperation with other specialists and contribute to
influencing our future. Come and join us in reinventing
light every day.

Light is OSRAM

14
ELECTRON MICROSCOPY Introduction

Although there are other factors involved, if one considers that no other aberrations limit
the ability of the lens, then the lens is said to be diffraction limited; meaning its resolution
is limited only by the diffraction pattern it produces. In such case, the limit of resolution
is determined by the ability to distinguish between two closely spaced Airy discs. This is
schematically shown in figure 1.6. The Airy disc radius is measured as the distance between
the first order peak and the first order trough (i.e. from the midpoint of the central diffraction
disc to the minimum within the first dark ring). Resolution is empirically described as the
ability to discriminate between two airy discs placed side by side. If an object is just below
the level of resolution, the peaks generated by the two points will make the object appear
to be a single point (Fig. 1.6).

Fig. 1.6: Resolution in terms of overlapping Airy disks

Köhler illumination, Abbe’s equation and Numerical Aperture


One of the critical problems in transmitted and reflected light optical microscopy is to
create a homogeneous intense illumination on the specimen, at the same time ensuring
that the filament image from the bulb, which provides the light, is not visible in the final
image. This problem was effectively solved in 1893 when August Köhler designed a new
method of illumination with a perfectly defocused image of the light source to illuminate
the sample [1.9]. The Köhler method allows adjusting the size and the numerical aperture
of the object illumination in a microscope independent of each other. This illumination
source is the predominant technique for sample illumination in modern optical microscopes.

Even before Köhler, Ernst Abbe in 1873 devised a method for determining the resolving
power of lenses. In his calculations, Abbe took into account the refractive index (n) of the
medium between the illumination source and lens relative to free space, the natural sine of
half the angle (α) of the cone of light from specimen plane accepted by the objective lens
and the wavelength of light (λ). The first two terms he multiplied together to produce a
value he called numerical aperture (NA):

NA = n sin α

15
ELECTRON MICROSCOPY Introduction

NA is a measure of the ability of the objective lens to gather light and resolve it, at a fixed
distance from the specimen, through a specific media (e.g. air, water, immersion oil). In an
optical microscope, the value of n is between 1.0 (for air) up to 1.51 (for special immersion
oil). When imaging through air (usually with low magnifications), then n = 1 and so NA
depends solely on the sin of a which theoretically can be a maximum of 1, as sin (90°) = 1.
In practice, it is difficult to achieve NA > 0.95 with “dry” microscopy (i.e. through air).
Therefore, many optical microscope objectives (mostly for high magnifications) are designed
to be used with water (n = 1.33), glycerin (n = 1.47) or immersion oil (n = 1.51). The
resolving power (R) of the lens then can be expressed by Abbe’s formula as:

R = λ/2 ΝΑ

This applies in a microscope, when the NA of the condenser lens is equal to or greater
than the objective lens’s NA and when the illumination consists of nearly parallel rays
formed into a cone of light whose angle match the objective lenses angular aperture (i.e. at
a distance equal to f = focal length of the lens: Fig. 1.7). All of these criteria are fulfilled
in Köhler illumination.

D
Objective lens

α
f

NA = n Sin α

Fig. 1.7: Light from the specimen enters the objective


in an inverted cone in Köhler illumination.

16
ELECTRON MICROSCOPY Introduction

In 1898, Lord Rayleigh extended Abbe’s work [1.10] and showed that the interplay between
diffraction and aberration can be characterized by the point spread function. The narrower
the aperture of a lens the more likely the PSF is dominated by diffraction. In that case, the
angular resolution of an optical system can be estimated (from the diameter of the aperture
and the wavelength of the light) by the Rayleigh criterion. According to this criterion, two
point sources are regarded as just resolved when the principal diffraction maximum of one
image coincides with the first minimum of the other (Fig. 1.6). Rayleigh considered that for
two sources of equal strength, if the distance is greater, the two points are well resolved and
if it is smaller, they are regarded as not resolved [1.10]. The resolution R in a microscope
(here measured as a distance and not to be confused with the angular resolution) is given by:

R = 1.22 λ / (ΝΑ objective + ΝΑ condenser)

The factor 1.22 is derived from a calculation of the position of the first dark circular ring

360°
surrounding the central Airy disc of the diffraction pattern (the calculation involves a Bessel

.
function). It follows that the numerical apertures of both the objective and the condenser
lenses should be as high as possible for maximum resolution. In the case, that NA for both
lenses is the same, the equation may be reduced to:
thinking
R = 0.61 On sin D

360°
thinking . 360°
thinking .
Discover the truth at www.deloitte.ca/careers Dis

© Deloitte & Touche LLP and affiliated entities.

Discover the truth at www.deloitte.ca/careers © Deloitte & Touche LLP and affiliated entities.

Deloitte & Touche LLP and affiliated entities.

Discover the truth at www.deloitte.ca/careers


17
ELECTRON MICROSCOPY Introduction

This is the diffraction-limited resolution of an optical system. If all aberrations and distortions
are eliminated from the optical system, this will be the limit to resolution. If aberrations
and distortions are present in the lenses, they will determine the practical limit to the
resolution. Note that the numerical aperture value of the optical system is an important factor
in these equations. The effect of the wavelength of light is the other limiting factor, with
longer wavelengths producing lower resolution. At a fixed numerical aperture (= 0.95), the
wavelength versus the resolution as calculated from the above criterion is listed in Table 1.

Color and wavelength of light Wavelength used in calculation Resolution


(l nm) (l nm) (R µm)

Violet (380–450) 400 0.26

Blue (450–495) 475 0.31

Green (495–570) 510 0.33

Yellow (570–590) 570 0.37

Red (620–750) 650 0.42

Resolution in the microscope is limited to about half the wavelength of the illumination
source used to image the sample. As the wavelength of visible light ranges from about
400 to 700 nanometers, it became quite clear that the best compound optical microscopes
cannot resolve parts of a specimen that are closer together than about 200 nanometers. In
the early twentieth century, scientists has theorized ways of getting around the limitations
of the relatively large wavelength of visible light by using electrons and this eventually led
to the development of electron microscopes.

De Broglie equation and Wavelength


By combining some of the principles of classical physics with the quantum theory, Louis-
Victor de Broglie (in 1924) proposed that moving particles have wave-like properties and
that their wavelength can be calculated based on their mass and energy levels [1.11]. The
general form of the de Broglie equation is as follows:

O = h m Q

where:
l = wavelength, h = Planck’s constant (6.6 × 10-27), m = mass of the particle (for electron
it is 9.1 × 10-28 g) and n = velocity of the particle.

18
ELECTRON MICROSCOPY Introduction

The wavelength of electrons is related to their kinetic energy via the de Broglie equation
and their wave-like properties mean that a beam of electrons can be made to behave like a
beam of electromagnetic radiation. When an electron passes through a potential difference
(accelerating voltage field) V, its kinetic energy will be equal to the energy of the field, i.e.
eV (energy in electron volts) = V (the accelerating voltage). Additional correction must be
made to account for relativistic effects as electron’s velocity inside an electron microscope
approaches the speed of light (c). By using some assumptions about the velocity of the
particle and its mass, it is possible to express either wavelength (l) or velocity (n ) in terms
of the accelerating voltage (V ):

Some parameter values of the electrons as a function of the accelerating voltage in an electron
microscope are given in Table 2.

Accelerating Nonrelativistic Relativistic


Mass Velocity
Voltage Wavelength Wavelength
(x m0) (x 108 M/s)
(KV) (nm) (nm)

100 0.00386 0.00370 1.196 1.644

120 0.00352 0.00335 1.235 1.759

200 0.00273 0.00251 1.391 2.086

300 0.00223 0.00197 1.587 2.330

400 0.00193 0.00164 1.783 2.484

1000 0.00122 0.00087 2.957 2.823

Table 1.2: Electron parameter values as a function of accelerating voltage

Clearly, replacing light by an electron beam as the illuminating source enhances the resolution
in electron microscope.

19
ELECTRON MICROSCOPY Introduction

REFERENCES
1.1. D. Bardell, The Invention of the Microscope, BIOS, 75 (2004) 78–84.

1.2. T.F. Glick, S.J. Livesey, F. Wallis (editors), Medieval science, technology, and medicine:
an encyclopedia, Series: Routledge Encyclopedias of the Middle Ages, Taylor and Francis
Group, New York, (2005) 624 pages. ISBN 978-0-415-96930-7.

1.3. A. van Helden, S. Dupré, R. van Gent, H. Zuidervaart (editors), The Origins of the
Telescope, Knaw Press, Amsterdam, (2011) 368 pages. ISBN-13: 978-9069846156.

1.4. S.J. Gould, The Lying Stones of Marrakech: Penultimate Reflections in Natural History,
Chapter 2: The Sharp-Eyed Lynx, Outfoxed by Nature, Harmony, New York, (2000)
372 pages. ISBN 0-224-05044-3.

1.5. W. Greiner, Quantum Mechanics: An Introduction, Springer, Frankfurt am Main,


(2001) 486 pages. ISBN 3-540-67458-6.

We will turn your CV into


an opportunity of a lifetime

Do you like cars? Would you like to be a part of a successful brand? Send us your CV on
We will appreciate and reward both your enthusiasm and talent. www.employerforlife.com
Send us your CV. You will be surprised where it can take you.

20
ELECTRON MICROSCOPY Introduction

1.6. Chr. Huygens, Traitė de la Lumiere (Treatise on Light): completed in 1678 (published
in Leyden in 1690) – Referred in: O.S. Heavens and R.W. Ditchburn, Insight into
Optics, Wiley & Sons, Chichester, (1987). ISBN 0-471-92769-4.

1.7. A.J. Fresnel, Mémoire sur la Diffraction de la lumière, où l’on examine particulièrement
le phénomène des franges colorées que présentent les ombres des corps éclairés par un point
lumineux (Memory on the Diffraction of Light, in which one particularly examines the
phenomenon of the colored fringes presented by the shadows of bodies illuminated
by a luminous point), Annales de chimie et de physique, 1 (1816) 239–280.

1.8. G.B. Airy, On the Diffraction of an Object-glass with Circular Aperture, Transactions
of the Cambridge Philosophical Society, 5 (1835) 283–291.

1.9. A. Köhler, Ein neues Beleuchtungsverfahren für mikrophotographische Zwecke (A new
lighting process for microphotographic purposes), Zeitschrift für wissenschaftliche
Mikroskopie und für Mikroskopische Technik, 10 (1893) 433–440.

1.10. Lord Rayleigh, Investigations in optics, with special reference to the spectroscope,
Philosophical Magazine, 8 (1879)261–274. doi:10.1080/14786447908639684.

1.11. L. de Broglie, The reinterpretation of wave mechanics, Foundations of Physics 1 (1970).
5–15. doi:10.1007/BF00708650.

21
ELECTRON MICROSCOPY Historical Timelines

2 HISTORICAL TIMELINES
2000 BC – The Chinese use water microscopes made of a lens and a water-filled
tube to visualize the unseen

612 BC – The Assyrians manufacture the world’s oldest surviving lenses

1000 – The first vision aid was invented (inventor unknown) called a reading
stone, which was a glass sphere that magnified the image when laid
on top of the reading material

1284 – Salvino D’Armate (Italian) is credited with inventing the first wearable
eye glasses

1500 – The art of grinding lenses is developed in Italy and spectacles are made
to improve eyesight

1590 – Hans and Zacharias Janssen (Dutch) – make the first microscope by
placing two lenses in a tube

1609 – Galileo Galilei – develops a compound microscope with a convex and


a concave lens

1665 – Robert Hooke – publishes Micrographia, a collection of biological


micrographs. He coins the word cell for the structures he discovers in
cork bark

1674 – Anton van Leeuwenhoek – improves on a simple microscope for viewing


biological specimens to look at blood, insects and many other objects.
He was first to describe cells and bacteria, seen through microscopes

1860s – Ernst Abbe – discovers the Abbe sine condition, a breakthrough in


microscope design, which until then was largely based on trial and error

1863 – Henry Clifton Sorby – develops a metallurgical microscope to observe


structure of meteorites

22
ELECTRON MICROSCOPY Historical Timelines

1865 – The company of Carl Zeiss – becomes the dominant microscope


manufacturer of its era

1897 – Joseph John Thompson – Discovery of the Electron

1903 – Richard Zsigmondy – develops the ultramicroscope to study objects


below the wavelength of light

1926 – Richard Zsigmondy – Nobel Prize in Chemistry

1926 – H. Bush – Magnetic/Electric Fields as Lenses

1929 – E. Ruska – PhD Thesis Magnetic lenses

1931 – Max Knoll and Ernst Ruska starts to build the first Transmission
electron microscope (TEM)

1932 – Davisson and Calbrick – Electrostatic Lenses

23
ELECTRON MICROSCOPY Historical Timelines

1934 – Driest and Muller – EM surpasses resolution limit of LM

1935 – Max Knoll – first scanning microscope

1936 – Erwin Wilhelm Müller – invents the field emission microscope

1937 – Manfred Von Ardenne – first scanning transmission electron microscope


(STEM)

1938 – Ruska (at Siemens) – first commercial TEM ~10 nm resolution

1942 – Vladimir Kosmich Zworykin (Russian) – first scanning electron


microscope (SEM)

1944 – 2 nm resolution achieved in TEM

1951 – Erwin Wilhelm Müller – the first to see atoms

1953 – Frits Zernike – Nobel Prize in Physics for his invention of the phase
contrast microscope

1965 – ~0.2 nm resolution

1965 – Cambridge scientific instrument company – first commercial SEM

1968 – A. Crewe – STEM ~0.3 nm resolution probe size achieved with Field
Emission Gun

1981 – Gerd Binnig and Heinrich Rohrer – develop the scanning tunneling
microscope (STM)

1986 – Ruska, Binnig and Rohrer – Nobel Prize in Physics jointly for invention
of TEM and STM

1996 – National center for electron microscopy (NCEM) USA – sub-Armstrong


resolution microscope (< 0.1 nm) achieved by one-Armstrong microscope
(OAM) project

2009 – Lawrence Berkeley National Laboratory (USA) – 0.05 nm resolution


achieved by Transmission Electron Aberration-Corrected Microscope
(TEAM)

24
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

3 ELECTRON BEAM INTERACTION


WITH MATERIALS
In this chapter you will learn about

-- Electron-Matter Interaction
-- Scattering and Diffraction
-- Electron Scattering
-- Elastic and inelastic scattering
-- Coherent and incoherent scattering
-- Forward and back scattering
-- Interaction cross-section and mean free path
-- Differential cross-section and interaction volume
-- Electron Diffraction
-- The angle of scattering
-- Bragg’s law
-- Scattering of X-rays, neutrons and electrons
-- Diffraction of electrons
-- X-ray generation
-- Bremsstrahlung
-- Characteristic radiation

3.1 ELECTRON SCATTERING


3.1.1 ELECTRON-MATTER INTERACTION

The electron is a low mass elementary particle that carries negative charge. It was discussed
in Chapter 1 that accelerated electrons act not only as particles but also as waves. As such,
electrons can easily be deflected by passing close to other electrons or positively charged
nucleus of an atom. These electrostatic (Coulomb) interactions cause electron to deflect /
deviate from its path (Fig. 3.1). Particles are scattered (deflected) and therefore, when one
considers the electron as particle, the interaction is termed as scattering. However, as electron
beam can also be considered as a wave, and as waves are not scattered but diffracted, the
interaction is then termed as diffraction. Electron scattering/diffraction is fundamental
phenomenon in all electron microscopy. We cannot see anything in electron microscopes
unless the specimen interacts with and scatters electrons in some way. In other words, any
non-scattering object is invisible. One can say that if the electrons were not scattered, there
would be no mechanism to create electron microscopic images. Therefore, it is essential
to understand both the particle approach and the wave approach to electron scattering/
diffraction in order to be able to interpret all the information that comes from an electron
microscope. Electron scattering from matter is however complex.

25
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

Incident electrons

BSE

nucleus

electron cloud
-

Scattered electrons
Fig. 3.1: Scattering of electron inside an atom.
BSE = backscattered electron.

AXA Global
Graduate Program
Find out more and apply

26
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

When an electron beam hits on to a material various interactions can occur. As the electrons
travel through the specimen, either they are scattered by a variety of processes or they may
remain unaffected. Scattering from a specimen is influenced by the kind of atoms the
specimen contain, its thickness, density and crystalline or amorphous nature. Scattering is
also governed by the acceleration of the electrons in the incident beam and the angle at
which the beam strikes the specimen. A thin specimen permits electrons to be transmitted
through it and scattering is mainly in the forward directions, although some electrons
may be also scattered back (Fig. 3.2a). As the specimen gets thicker, more electrons are
backscattered and in a bulk specimen, the electrons in the incident beam are scattered only
in the backward direction (Fig. 3.2b). Scattering in bulk specimen also generates secondary
electrons. The result, however, is that a non-uniform distribution of electrons emerge from
the exit surface of the specimen (Fig. 3.3). The spatial distribution of scattering can be
observed in electron microscopes as contrast in image of the specimen, and the angular
distribution of scattering can be viewed in the form of scattering patterns, for example the
diffraction patterns.

Incident electron beam


Incident electron beam
Backscattered
electrons

Thin specimen Backscattered Secondary


electrons electrons

Scattered/
Diffracted beam
Direct Bulk specimen
transmitted beam

a. b.
Fig. 3.2: Passage of incident electron beam through: (a) Thin and (b) Bulk specimen.

Incident electron beam


(coherent/uniform intensity)

Thin specimen

Scattered electron beam


(varying intensity)

27
Diffracted beam
Direct Bulk specimen
transmitted beam

a. b.
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

Incident electron beam


(coherent/uniform intensity)

Thin specimen

Scattered electron beam


(varying intensity)

Fig. 3.3: Scattering within the specimen changes both


the spatial and angular distributions of the emerging
electrons. The spatial distribution (intensity) creates
contrast in images of the specimen.

3.1.2 ELECTRON SCATTERING


For the rudimentary description of the elastic scattering of a single electron by an atom, it is
sufficient to regard it as a negatively charged particle and neglect its wave properties. When
an electron penetrates into the electron cloud of an atom in the specimen it is attracted
by the positive potential of the nucleus and repelled by the electron cloud (electrostatic or
Coulomb force), and as a result its path is either deflected towards the core or deflected
in other directions (Fig. 3.1). Furthermore, other kinds of interactions also occurs which
results in the generation of different signals schematically shown in Figure 3.4. These signals
can actually be detected in the microscope using various types of detectors and utilized for
forming an image and for analytical characterization of the specimen.

Incident beam
Backscattered electrons (BSE) Characteristic X-rays

Secondary electrons (SE)


Auger electrons
Cathodluminescence

Absorbed Electron-hole
specimen
electrons pairs

Elastically scattered electrons


Bremsstrahlung X-rays
Inelastically scattered electrons
Direct beam

Fig. 3.4: The various signals generated in a specimen when an incident electron beam hits a
specimen and interacts.

28
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

Types of scattering
The interactions of electrons with matter are classified according to the nature of the signal
it generates. The various types of electrons (signals also include x-rays and photons) that
are emitted from the sample on electron bombardment of the specimen are shown in
Fig. 3.4. However, the emitted electrons can also be classified in various other ways. For
example, they can be classified according to their scattered angle (forward or backscattering)
or according to the loss in their energy (namely elastic and inelastic interactions) or one
can separate scattered electrons into coherent and incoherent (which then refers to the wave
nature of electron).

The terminologies forward scattering or backscattering refer to the angle of scattering with
respect to the incident beam and the specimen that is normal to that beam. If an electron
is scattered through < 90°, then it is forward scattered and when the scattering angle is
> 90° it is backscattered.

�e Graduate Programme
I joined MITAS because for Engineers and Geoscientists
I wanted real responsibili� www.discovermitas.com
Maersk.com/Mitas �e G
I joined MITAS because for Engine
I wanted real responsibili� Ma

Month 16
I was a construction Mo
supervisor ina const
I was
the North Sea super
advising and the No
Real work he
helping foremen advis
International
al opportunities
Internationa
�ree wo
work
or placements ssolve problems
Real work he
helping fo
International
Internationaal opportunities
�ree wo
work
or placements ssolve pr

29
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

Elastic interaction results in no energy transfer from the incident electron to the sample and
as a result, the electron leaving the sample still has its original energy El = E0 (Fig. 3.5a).
Elastic scattering involves the interaction of an incident electron with an atomic nucleus.
As the nuclear mass greatly exceeds the mass of an electron (more than thousand times),
the energy exchange (loss) is very small and practically unmeasurable. While in the case
of inelastic interaction, due to Coulomb interaction between the incident electron and the
atomic electron, the similarity of mass between the projectile (incident electron) and the
target (atomic electron) allows the energy exchange (loss) to be appreciable. In inelastic
scattering energy is transferred from the incident electrons to the sample due to the interaction
and as a result the energy of the electron after the interaction is consequently reduced El
< E0 (Fig. 3.5b). Of course, no energy is also transferred if the electron passes through the
sample without any interaction at all. Such electrons contribute to the direct beam, which
contains the electrons that continue to move in the same direction of the incident beam. In
general, the elastic scattering events affect the trajectory of the beam of electrons inside the
specimen without changing the kinetic energy of the electron. In addition, elastic scattering
is responsible for the phenomenon of electron backscattering that forms an important
imaging signal in SEM. On the other hand, inelastic scattering events result in a transfer of
energy from the electrons in the beam to the atoms in the specimen. This event leads to
the generation of secondary electrons; Auger electrons; characteristic x-ray and bremsstrahlung
(continuum x-ray). It also generates electron-hole pairs in semiconductors and insulators (Fig.
3.4). Some electrons are absorbed in the specimen. Many of the different signals generated
in the specimen are utilized in the methods of analytical electron microscopy.

a. b.
Fig. 3.5: Schematic illustration of scattering process that occurs when an energetic electron of energy
E0 interacts with an atom: (a) Elastic scattering, in which the energy E1 after scattering equals E0.
(b) Inelastic scattering, in which the energy E1 after scattering is less than E0. Note that the scattering
angle in inelastic scattering is much less than in elastic scattering.

30
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

By considering electrons as wave scattering can be further separated into coherent and incoherent.
This distinction is related to the previous classification since elastically scattered electrons are
usually coherent and inelastic electrons are usually incoherent. Assuming that the incident
electron waves are coherent, that is, they are essentially in phase with one another and have
a fixed wavelength governed by the accelerating voltage (Fig. 3.3). Then, coherently scattered
electrons are those that remain in step and incoherently scattered electrons breakaway from
the step, i.e. phase relationship in the wave is changed after interaction with the specimen.
As long as the scattering angle remains low in elastic scattering, electrons usually remain
coherent. Whereas, elastic scattering to higher angles (> ~10°) are scattered incoherently.
Inelastic electrons are generally incoherent, even when scattered to low angles (< 1°).

The simplest scattering process is single scattering and we often approximate scattering
within a specimen to this process (i.e., an electron either undergoes a single-scattering event
or it suffers no scattering on passing through the specimen in a microscope). This can be
a reasonable assumption if the specimen is very thin (as in TEM), however, generally in a
bulk specimen (as in SEM) an electron is scattered more than once and then it is termed
as plural or multiple scattering. As TEM specimens are relatively thin, multiple scattering
(which is usually defined as scattering more than 20 times) is infrequent, but for the bulk
specimen used in SEM multiple scattering is very common. Elastic scattering usually occurs
at relatively low angles (1–10°), i.e., scattering is in the forward direction and this is true in
TEM. As the specimen gets thicker, fewer electrons are scattered only once, (i.e. multiple
scattering dominates) and more electrons are backscattered. Incoherent, backscattered electrons
are the only remnants of the incident beam that emerge from bulk specimen. In SEM, other
signals are also generated through inelastic interaction of the incident electrons, i.e. secondary
electrons and they emerge from the backside of the specimen as well. The greater the number
of scattering events, the more difficult it is to predict its path and what will happen to the
electron and therefore, it is more difficult to interpret the image, the diffraction pattern or
the spectra that emerge from the electron microscope. Generally, more the scattering events,
greater is the angle of scattering, although sometimes a second scattering event can redirect
the electron back into the direct beam and it will appear to have undergone no scattering.

Interaction cross-section and mean free path


If an electron passes through a specimen, it may be scattered not at all, once (single
scattering), several times (plural scattering), or many times (multiple scattering). The chance
that an electron is scattered, can be described either by the probability of a scattering event,
as determined by the interaction cross-section (s), or by the average distance an electron
travels between two interactions, i.e. the mean free path (L).

31
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

When an electron travels inside a material, it will have a certain probability to interact
with the nuclei or with the electron cloud of atoms in that material. In a very thin slice
of matter, this probability is proportional to the number of potential target atoms per
unit volume in the material and the thickness of the slice. A certain cross-section in the
matter determines the chance an electron will undergo any interaction. This also depends
on the electron beam energy and the type of the interaction. Thus, one can say that the
possible interaction of electron with a material has a certain cross-section. The concept of
the interaction cross-section is based on a simple model of an effective area within which
if an electron passes an interaction will certainly occur, as shown in the case of electrons
in incident beams 1 and 3 in Fig. 3.6a. For every interaction, the cross-section (s) can be
defined depending on an effective radius r, [3.1], where r has a different value for each
type of scattering process.

V = S r2

93%
OF MIM STUDENTS ARE
WORKING IN THEIR SECTOR 3 MONTHS
FOLLOWING GRADUATION

MASTER IN MANAGEMENT
• STUDY IN THE CENTER OF MADRID AND TAKE ADVANTAGE OF THE UNIQUE OPPORTUNITIES
Length: 1O MONTHS
THAT THE CAPITAL OF SPAIN OFFERS
Av. Experience: 1 YEAR
• PROPEL YOUR EDUCATION BY EARNING A DOUBLE DEGREE THAT BEST SUITS YOUR
Language: ENGLISH / SPANISH
PROFESSIONAL GOALS
Format: FULL-TIME
• STUDY A SEMESTER ABROAD AND BECOME A GLOBAL CITIZEN WITH THE BEYOND BORDERS
Intakes: SEPT / FEB
EXPERIENCE

5 Specializations #10 WORLDWIDE 55 Nationalities


MASTER IN MANAGEMENT
Personalize your program FINANCIAL TIMES
in class

www.ie.edu/master-management mim.admissions@ie.edu Follow us on IE MIM Experience

32
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

For the case of elastic scattering, this radius depends on the atomic number (Z   ) of the target
atom. Figure 3.6a schematically shows the distribution of the cross-section of atoms in a
material of total area A. The interaction cross-section is an effective area that quantifies the
intrinsic likelihood of a scattering event, when an incident electron beam strikes the target
object. If the cross-section of an atom is divided by the total area, then a probability for an
interaction event is obtained. Consequently, the chance of a definite interaction increases
with increasing cross-section. However, as each scattering event might occur as elastic or
as inelastic interaction, therefore, the total interaction cross section is the sum of all elastic
and inelastic events.

1 2 3

1 2 6

r 7 12
3 5
11
4 8
Total area A 10
9
a. b.

Fig. 3.6: (a) Interaction cross-section is the probability of an electron scattering event when an incident
electron beam strikes the target object. (b) Mean free path is the average distance that is traveled
by an electron between two scattering events.

Another important possibility to describe interactions is the mean free path (L). In figure 3.6b
one sees that an electron, which traverse the material, is scattered several times and the
distance between two consecutive scattering events are not the same. Instead of using an
area to describe the interaction one can use a length since the distance an electron travels
between interactions with atoms is clearly also important. The mean free path (L) is related
to the scattering cross-section. This is the average distance in the scattering process, which is
traveled by an electron between two scattering events (Fig. 3.6b). Monte Carlo simulations
can be used to predict the electron paths as a beam is scattered within a specimen. This is
a statistical method and use random numbers for calculating the paths of electrons. The
probability of scattering is taken into account in the calculation, as well as other important
parameters like acceleration voltage of the incident electron, the atomic number (Z   ) of
the target atoms, thickness and density of the material. Monte Carlo methods are more
frequently used in SEM imaging and X-ray spectroscopy calculations and less frequently in
TEM. Typical examples of calculated scattering by Monte Carlo methods in thin film and
bulk Si specimens are shown in Fig. 3.7.

33
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

a. b.
Fig. 3.7: Monte Carlo simulation of scattering events in Si: (a) thin specimen (10 nm thick), incident electron beam
(200 kV) and (b) bulk specimen, incident electron beam (25 kV). Incident electron beam hits the specimen at 0°
tilt angle and electrons penetrate the specimen and scatter. Simulation was done with Electron Flight Simulator
(version 3.1 E) software [3.2].

Differential cross-section and interaction volume


When an electron encounters a single isolated atom, it is scattered through an angle (q)
into some solid angle (W) and those scattered electrons are collected within the solid angle
W (Fig. 3.8). Differential cross-section is the angular distribution of the scattered electrons.

Incident electrons

Scattering object

Scattered electrons


dΩ

Direct beam

Fig. 3.8: When incident electrons are scattered through


a solid angle the differential cross-section is the angular
distribution of the scattered electrons.

34
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

There is a simple geometrical relationship between q and W.

: S(1 -cosT

With the increase in the scattering angle q the probability of scattering (i.e. the cross-
section s) decreases, as scattering into high angles is rather unlikely. The angular distribution
of electrons scattered by an atom is described by the differential cross-section ds/dW. It is
a measure of the probability for scattering within a solid angle dΩ. The ds/dW value is
important since it is possible to measure it experimentally.

35
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

The combined effect of elastic and inelastic scattering is to limit the penetration of the incident
electron beam into the solid. The resulting region over which the incident electrons interact
with the solid, depositing energy and producing different forms of secondary radiation is
known as the interaction volume. Figure 3.9 schematically depict the interaction volume and
shows which type of signal is generated from within this volume and from what depth. For
example, Auger electrons are generated at the surface of the specimen and the characteristic
x-rays from deep inside.

Fig. 3.9: Interaction volume

An understanding of the size and shape of the interaction volume as a function of the specimen
and the beam parameters is vital for proper interpretation of features of image formation
and the spatial resolution of x-ray microanalysis in electron microscope. Often Monte Carlo
simulation is used to get an idea about how the scattering takes part in the solid. Although
such calculations are rather rough estimates of the real physical processes, the results describe
the interaction and, in particular, the shape and size of the interaction volume reasonably well.
Simulations of interaction volume in bulk Ni for different accelerating voltage of the incident
electron beam is shown in Fig. 3.10. It can be seen that the depth of penetration and the
interaction volume strongly depends on the acceleration voltage of the incident electron beam.

Fig. 3.10: Interaction volume in Nickel as a function of the acceleration voltage of the incident
electron beam calculated by Monte Carlo simulation.

36
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

3.2 ELECTRON DIFFRACTION


The terminology diffraction and scattering can be confusing as diffraction is a very special
form of elastic scattering. This confusion is more in the case of electrons because electron
in a beam can be conceived both, as particle and wave. In the simplest term diffraction is
a deviation which occur when a wave encounters an obstacle or a slit (as the bending of
light in a lens or around the corners of an aperture), as was discussed briefly in Chapter 1.
Whereas scattering is the process, in which particles are deflected because of collision (as
discussed above in this Chapter).

The diffraction phenomenon is described as the interference of waves according to the


Huygens-Fresnel principle [3.3–3.4] and it occurs with all waves (such as visible light or
electromagnetic waves and include x-rays, radio waves, sound waves and water waves). While
diffraction occurs whenever propagating waves encounter obstruction, its effects are most
pronounced for waves whose wavelength is roughly comparable to the dimensions of the
diffracting object. When the obstructing objects are multiple and closely spaced, a complex
pattern of varying intensity can result from diffraction. This is due to the interference of
different parts of the wave that travel by different paths. Thus, electrons after transmitting
through the matter may give rise to interference, either constructive or destructive. Two
coherent waves, which are scattered at different points, can meet and due to the difference
in the length of the path they travel, they are in different phases when they meet. The
interference will be constructive when the two waves have the same phase and destructive
when there is phase shifted. This is very important for electron waves transmitting through
matter, as there are multiple atoms in the matter and their spacing is comparable to the
wavelength of electron beam.

Solely incoherent scattering of the incident coherent electrons takes place if the scattering
centers are arranged in an irregular way. This is the distribution of atoms in amorphous
compounds. Although the scattering is predominantly in forward direction, the scattered
waves are no more in phase with respect to each other after the scattering events. Therefore,
an enhancement of the wave intensity through constructive interference cannot happen.
On the other hand, if a crystalline specimen is transmitted by electrons with a certain
wavelength l, then coherent scattering takes place. All the regularly arranged atoms in the
crystal act as scattering centers. The scattering at each point occurs in accordance with the
electrostatic interaction of the nucleus with the negatively charged electron as discussed.
Since the spacing between the scattering centers is now regular, constructive interference
of the scattered electrons in certain specific directions happen and thereby diffracted beams
are generated (Fig. 3.11).

37
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

incident beam

lattice
planes

diffracted beam direct beam



Fig. 3.11: Diffraction of an electron beam by a crystal.
The crystal is schematically represented by a set of
parallel equidistant lattice planes.

In the past 5 years we have drilled around

95,000 km
—that’s more than twice around the world.

Who are we?


We are the world’s leading provider of reservoir characterization,
drilling, production, and processing technologies to the oil and
gas industry.

Who are we looking for?


We offer countless opportunities in the following domains:
n Operations
n Research, Engineering, and Manufacturing
n Geoscience and Petrotechnical
n Commercial and Business

We’re looking for high-energy, self-motivated graduates


with vision and integrity to join our team. What will you be?

careers.slb.com

38
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

Bragg diffraction was first proposed by William Lawrence Bragg and his father William
Henry Bragg in 1913 [3.5]. They studied crystalline solids with x-rays and found that x-rays
reflected by crystals produced certain patterns. The x-ray wave of specific wavelength when
incident on a crystal produced intense peaks of the reflected radiation at certain specific
angles. This phenomenon is known as Bragg diffraction. The concept of Bragg diffraction
applies equally to neutron and electron diffraction processes as well [3.6].

3.2.1 BRAGG’S LAW

The conditions for constructive interference for two electron waves diffracted at two parallel
lattice planes packed with atoms are depicted in Figure 3.12. The two incident electron
waves are in phase with each other before hitting the crystal. For constructive interference,
the two waves have to be in phase again, after diffraction at the lattice planes. For this to
happen the path difference between the waves must be an integer multiple of the electron
wavelength. This path depends on the incident angle q, the distance between the lattice
planes d and the electron wavelength l. The path length is twice d sinq. Since this value
must be a multiple of l, it follows:

2 d sinT = nO.

This is the Bragg equation for diffraction.

incident wave scattered wave

θ θ

d Sinθ

Fig. 3.12: Bragg diffraction: for constructive interference, the path difference of the waves
(2d sinθ) must equal the integer multiple of the wavelength (nl).

39
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

X-ray, neutron and electron diffraction


X-ray diffraction (XRD), as well as neutron and electron diffraction (ND and ED) are all
caused by constructive interference of scattered waves, and the same fundamental Bragg
law apply. However, ED, ND and XRD show some distinct differences. The first important
difference is their wavelength. Typically, the wavelength of electrons (e.g., 0.0197 Å for
300 keV electrons) is much shorter than that of x-rays and neutron (which are in the order
of 1 Å). Therefore, the radius of the Ewald sphere1 is much larger and more reflections are
observed by ED than by XRD and ND. Secondly, the diffraction angles are very small in
ED: 0 < Θ < 2° compared to XRD: 0 < Θ < 180º. Moreover, electrons are scattered by
both the negatively charged electron cloud and the positively charged nuclei in a material
(Coulomb interaction), while x-rays interact only with the electron-cloud and neutrons
interact with atomic nuclei and the magnetic field of unpaired electrons. Since neutrons
are electrically neutral, they penetrate matter more deeply than electrically charged electrons
of comparable kinetic energy. However, the interaction of electrons with matter is much
stronger (106–107 times more) than that of x-rays. Further, in the case of both x-rays and
electrons, the scattering power falls off with increasing scattering angle θ and absorption
by the matter is strong. Whereas, the scattering power of neutron is independent of θ and
lower number of neutrons are absorbed by the matter.

As electrons are scattered by both the electrons and the nuclei in a material, the incoming
negatively charged electrons interact with the local electromagnetic fields of the specimen.
The incoming electrons are therefore directly scattered by the specimen; it is not a field-to-
field exchange as occurs for the case of x-rays.

3.3 X-RAY GENERATION


A high-speed electron passing an atom within its electron cloud may be decelerated or
completely stopped by the Coulomb force of the nucleus and this inelastic interaction
generates x-rays. This braking radiation is termed Bremsstrahlung (a German term for braking
radiation). The moving particle loses kinetic energy, which is converted into a photon and
Bremsstrahlung has a continuous spectrum as shown in Fig. 3.13.

40
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

characteristic x-ray

X-ray intensity (a.u)

Bremsstrahlung

Photon energy (kev)


Fig. 3.13: Bremsstrahlung and Characteristic X-rays schematically presented.

Excellent Economics and Business programmes at:

“The perfect start


of a successful,
international career.”

CLICK HERE
to discover why both socially
and academically the University
of Groningen is one of the best
places for a student to be
www.rug.nl/feb/education

41
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

Further, if because of scattering a part of the energy of the incident electron beam is
transferred to the specimen several other processes also may occur and one of them is inner-
shell ionization. The incident electron travelling through the electron cloud of an atom might
displace an electron from an inner-shell of the target atom (Fig. 3.14a). The target atom is
then ionized because it carries a surplus positive charge and the resulting electronic state of
the generated ion is energetically unstable. When this happens, another electron from an
outer-shell of the atom is quickly attracted into the void in the deficient inner-shell. By this
process, the atom can relax but it must also give away the excess energy. This excess energy
of the electron is emitted as photon and leads to the generation of a characteristic X-ray.
The energies of characteristic photons are a function of the energy levels of various electron
orbital levels (Fig. 3.14b) and hence are characteristic of the target atoms. Characteristic
radiation has a higher intensity than Bremsstrahlung at the specific energy of the spectra as
shown in Fig. 3.13.

K electron removed
K


ejected K shell electron

K excitation
Increasing Energy

Kα x-ray emitted L shell electron fill vacancy

L electron removed
L
L excitation

+
M electron removed
M
K Mα
M

L N electron removed
N
M Valency electron
N

removed
shell orbits
Normal
a. b.

Fig. 3.14: X-ray generation: a) inner-shell ionization, b) excess energies of characteristic photons emitted
from the atom are a function of the energy of the various electron orbital levels in the atom.

42
ELECTRON MICROSCOPY Electron Beam Interaction with Materials

3.4 REFERENCES
3.1. R.D. Heidenreich, Fundamentals of Transmission Electron Microscope (Interscience
monographs and texts in physics and astronomy), Interscience, New York, (1964) 414
pages. ISBN-13: 978-0470368152

3.2. 
Electron Flight Simulator software (a simulation software based on the Monte
Carlo model by D. Joy, University of Tennessee), Small World LLC, (2004).
www.small-world.net

3.3. Chr. Huygens, Traitė de la Lumiere (Treatise on Light): completed in 1678, (published
in Leyden in 1690) – Referred in: O.S. Heavens and R.W. Ditchburn, Insight into
Optics, Wiley & Sons, Chichester, (1987), ISBN 0-471-92769-4.

3.4. A. J. Fresnel, Mémoire sur la Diffraction de la lumière, où l’on examine particulièrement
le phénomène des franges colorées que présentent les ombres des corps éclairés par un point
lumineux (Memory on the Diffraction of Light, in which one particularly examines the
phenomenon of the colored fringes presented by the shadows of bodies illuminated
by a luminous point), Annales de chimie et de physique, 1 (1816) 239–280.

3.5. 
W.H. Bragg, W.L. Bragg, The Reflection of X-rays by Crystals, Proceedings Royal
Society London A, 88 (1913) 428–438. doi: 10.1098/rspa.1913.0040.

3.6. J.M. Cowley, Diffraction physics, North-Holland, Amsterdam (1975) 481 pages.
ISBN-13: 978-0444822185.

43
ELECTRON MICROSCOPY Instrumentation

4 INSTRUMENTATION
In this chapter you will learn about

-- Electron Source
-- Thermionic and field emission source
-- Tungsten filament, LaB6 filament, cold cathode and Schottky emitter
-- Beam current, beam current density and source brightness
-- Spot size and resolution
-- Electron optics
-- Electrostatic and magnetic lens
-- Aberration and astigmatism
-- Detection of electrons and other signals
-- Electron detection and image recording in TEM
-- Electron detection and imaging in SEM
-- EELS detectors in TEM and STEM
-- Detection of x-ray signal
-- Recording Kikuchi patterns

American online
LIGS University
is currently enrolling in the
Interactive Online BBA, MBA, MSc,
DBA and PhD programs:

▶▶ enroll by September 30th, 2014 and


▶▶ save up to 16% on the tuition!
▶▶ pay in 10 installments / 2 years
▶▶ Interactive Online education
▶▶ visit www.ligsuniversity.com to
find out more!

Note: LIGS University is not accredited by any


nationally recognized accrediting agency listed
by the US Secretary of Education.
More info here.

44
ELECTRON MICROSCOPY Instrumentation

An electron microscope may be conveniently divided in three components according to


their function:

1. Source of illumination – consisting of the electron gun and condenser lens,


2. Objective lens – the heart of any microscope and used for the image formation
of the specimen. It is followed by a magnification system (which is conceptually
different in TEM and SEM),
3. Detectors and recording systems – used to display/store image (also other signals).

The differences in the construction and functioning of scanning and transmission electron
microscopes will be discussed in Chapter 5. Here the three main components of the electron
microscope are broadly discussed.

4.1 ELECTRON SOURCES


The electron microscopes use a beam of accelerated electrons for illumination. This electron
beam is generated within the microscope using the so-called electron gun. An electron gun
(also called electron emitter) is a device with an electrical component in a vacuum tube that
produces a narrow, collimated electron beam with precise kinetic energy. Previously, the
largest use of electron gun was in cathode ray tubes (CRT) which were used in television
sets, computer displays and oscilloscopes. However, in modern times CRTs are mostly
replaced by flat-panel displays. However, electron gun is still the major source of electrons
in the electron microscope. The function of the electron gun is to provide an intense
coherent beam of high-energy electrons. This is subsequently regulated and guided by a
set of condenser lens in the electron microscope. The electron gun actually determines the
electron beam parameters like the beam current (ib), the current density (Jb) and the spot size
(dp). The beam current controls the image contrast and influences intensity of x-ray signals
that are generated in the specimen. The spot size determines the image resolution (also
spatial resolution of analytical microscopy, e.g. elemental maps from x-ray spectroscopy).

45
ELECTRON MICROSCOPY Instrumentation

Electron emission
Electrons revolve around the nucleus of an atom in different orbital shells. Those at the
outermost shell are called valence electrons and they are loosely attached to the nucleus.
The valence electrons can gain enough energy to break the bonding with the parent atom
to move freely in metals, and they are called free electrons. However, these free electrons
cannot escape from the surface of the metal but are responsible for the metal’s electrical
conductivity. The process of electron emission is analogous to that of ionization of a free
atom, in which one or more electrons are eventually ejected from the atom. The energy of
electrons in an atom is lower than that of an electron at rest in vacuum and consequently
in order to ionize an atom, energy must be supplied to the electrons in some way or the
other. In an analogous situation, a substance does not emit electrons spontaneously, but only
if some of the electrons have energies greater than that of the electrons at rest in vacuum.
In other words, if one provides additional external energy to the metal it is possible to eject
the electrons from the surface of the metal. This process is called electron emission. There are
four methods of producing electron emission: 1. thermionic emission, 2. photoelectric emission,
3. secondary emission and 4. field emission. In thermionic emission, electron emission is achieved
by heating the electrode, while photoelectric emission is achieved through application of light
on the surface of the electrode. Electron emission from a metallic surface by bombardment
of high-speed electrons is known as secondary emission and in field emission; electrons are
extracted by the application of electrostatic field.

Electron gun
Amongst the four emission processes, only thermionic emission and field emission are actually
used in electron guns to produce illumination [4.1]. In both processes, they are generated
from a filament, usually made of tungsten metal but also sometimes from other crystals like
Lanthanum Hexaboride, LaB6. The electrons are then accelerated by an electric potential
and focused into a beam by electrostatic or electromagnetic lenses. The electron guns in
microscopes are therefore of two types: Thermionic gun and Field emission gun (FEG). Further,
the FEG sources are also of two kinds – Schottky field emission and Cold field emission. The
main components of the thermionic and field emission electron guns are schematically
shown in Fig. 4.1.

46
ELECTRON MICROSCOPY Instrumentation

<2 kV
W-hairpin
Filament
Field Emitter Tip

Grid
10-400kV
Wehnelt
cylinder –
10-1000 kV
+ Anode

crossover
crossover
Anode

Thermionic Electron Gun Field Emission Electron Gun


a. b.
Fig. 4.1: Thermionic and Field emission guns.

47
ELECTRON MICROSCOPY Instrumentation

The electron gun consists of a cathode, a Wehnelt cylinder (in thermionic gun) or a grid
(in field emission gun) and an Anode, which is positively biased (Fig. 4.1). The different
types of cathodes are shown in Fig. 4.2. The cathode in thermionic gun, is a “V” shaped
tungsten (W) wire hairpin or a LaB6 crystal and in the field emission gun, a tungsten rod
with a very sharp pointed tip at the end, typically < 50 nm tip radius. Both the W-hairpin
and the LaB6 cathodes emit electrons when heated to high temperature. The heating is
generally achieved by passing an electric current (emission current) through the filament.
On the other hand, in the field emission gun the cathode is held at a negative potential
relative to the grid (biased usually between 0 and 2000 V) such that the high electric field
at the cathode forces electrons to emit from the tip. A large potential difference between
the cathode and the anode is used to accelerate the electrons in the beam. Emitters in FEG
are either of cold cathode type or of Schottky-type. The cold cathode, as the name suggests,
is not heated but the Schottky electrodes are heated so that the thermionic emission is
enhanced by lowering of energy barrier in the presence of the high electric field. The cold
cathodes are made of solid W rod with pointed tip but the Schottky emitters are tungsten
tips coated with zirconium oxide, which has the unusual property of increasing electrical
conductivity at high temperatures. Although the Schottky filaments are heated, they are
heated to a much lower temperature than the thermionic filaments.

W filament LaB6 filament W FEG tip

Fig. 4.2: Different types of filaments used in electron guns.

The beam current density (Jb) is relatively constant for a given type of electron gun. More
precisely, the electron source brightness (ß) is fixed for a given type of electron source. The
brightness of the different cathodes follows the order – W-hairpin, LaB6 and FEG, listed in
the order of their increasing current density. Current density of the beam is however related
to both the beam current (ib) and the spot size (dp) and larger values of ib also means a larger
value of dp, particularly in a given electron source.

48
ELECTRON MICROSCOPY Instrumentation

The emitted electrons are accelerated towards the anode by the potential difference between
the cathode and anode. This acceleration voltage in a TEM is above 80 kV and can reach
as high as a million volt in some microscopes (but usually in the range of 200 to 400 kV).
Whereas in SEM the electrons are accelerated to, a much lower voltage (20–30 kV) and
therefore the electron beams in SEM have a lower kinetic energy than in TEM. Surrounding
the filament is a grid cup or Wehnelt cylinder with a circular aperture centered at the filament
apex. The effect of such a gun configuration is that the electric field causes the emitted
electrons to converge to a crossover. The electrons are accelerated through this voltage field
towards the anode. The current density (Jb) in the electron beam at the crossover represents
the current that could be concentrated into a focused spot and the maximum value is
determined by the type of filament source. The current density per unit solid angle is called
the electron beam brightness (ß ). The FEG is a much brighter source and has a smaller spot
size than thermionic guns. Amongst the thermionic filaments, LaB6 is a brighter source than
W-hairpin. The tungsten filaments also have much shorter filament life compared to LaB6
filaments and field emission cathodes, but are much cheaper.

The thermionic filaments are heated electrically. With increasing filament current (If  ) there
is an increase of the current in the electron beam (Ib) due to higher emission of electrons
from the cathode as the temperature increases. However, beyond a certain temperature the
rapid rise in the beam current (Ib) reaches a flat maximum, where after Ib is independent of
If  . When this condition is reached, the filament is saturated. In practice, this is achieved
by gradually increasing the filament current and observing the beam current meter until
there is no change. The LaB6 filament is sensitive to thermal shock and must be heated
slowly. They are therefore heated stepwise, increasing the filament current with a waiting
time between steps. During heating the unmagnified image of the source changes as shown
in Fig. 4.3. At the saturation point, the maximum current density is obtained in the beam.

Increasing If and Ib I b = If Increase in If but not in Ib


Fig. 4.3: Schematic representation of emission pattern from a thermionic electron
source as filament current (If ) is increased to saturation.

49
ELECTRON MICROSCOPY Instrumentation

The effective source size or the crossover diameter (do) of a field emitter is less than 10 nm
as compared with around 10 µm for LaB6 and around 50 µm for the tungsten filament.
No further demagnification is therefore necessary to produce a high-resolution image in
the FEG system. However, the source design must take care of electron optical aberrations
otherwise the benefits of the high brightness of the source is lost. Table 4.1 compares several
characteristics of the different types of electron sources.

Characteristics W-hairpin LaB6 Cold Field Emission

Filament Temperature (K) 2700 2000 300

Current density Jb (A/cm2) 1.75 100 104

Spot size dp (µm) 30 5 0.1

Brightness ß (A/cm2 x sr) 2.5 x 105 1 x 107 2 x 108

Filament Life (hr) 30 500 1000

Vacuum required (Torr) 10-5 10-7 10-10

Table 4.1 Characteristics of different electron sources compared at 100 KeV

50
ELECTRON MICROSCOPY Instrumentation

Condenser lens system


The purpose of the condenser lens system is to deliver electrons from the gun crossover to
the specimen under the specific condition needed for imaging in the microscope. It works in
conjunction with the condenser aperture to eliminate the high-angle electrons from the beam.
The condenser system is mainly used to form the beam and limit the amount of current in
the beam. It is also used to control the diameter of the electron beam and focus the electrons
onto the specimen. In practice, the electron stream is condensed in the condenser lens by
controlling the probe current. On increasing the probe-current (condenser lens current),
the beam focuses well above the aperture. A large part of the beam is intercepted by the
aperture and hence the size of the spot and the current is reduced. Two or more lenses
can act together and their ray diagrams can be constructed using thin-lens approximation
for each of them. The condenser illumination system consisting of two or more lenses and
an aperture is used in both SEM and TEM. Its function is to control spot size and beam
convergence. Fig. 4.4 shows the ray diagram for the double condenser system to produce
nearly parallel beam. The black dots represent the focal point of each lens.

In a conventional TEM a focused high intensity reduced image of the gun crossover at the
specimen is required for some mode of the microscope operation and a nearly parallel beam
at reduced intensity at other modes of operation. In SEM on the other hand, the singular
requirement is for a small diameter, high intensity probe on the specimen. The condenser
lens systems are designed accordingly to provide these diverse needs. The spot size of the
beam in SEM also determines the image resolution.

Gun crossover

Condenser
Lens C1

Condenser
Lens C2
Condenser aperture

Specimen

Fig. 4.4: Double condenser lens system in electron microscope.

51
ELECTRON MICROSCOPY Instrumentation

4.2 ELECTRON OPTICS


In a sense, an electron microscope works in the same way as an optical microscope. After
a beam of electrons is generated by the high voltage electron emitter (the electron gun) at
the top of the column, they travel down the column to interact with the specimen. These
electrons pass through and scattered within the thin specimen in a TEM or are scattered by
the bulk specimen in the backward direction in a SEM. Because of the scattering, an image
of the specimen can be formed, like in an optical microscope. An objective and a series of
magnifying lenses are needed in the optical microscope to form an image of the specimen
and magnify it. In the conventional optical microscope, this is achieved by means of two
or more glass lenses. The basic characteristic of a lens that enables it to form a magnified
image is its ability to bring to a focus a broad beam of light falling on it. Since many of
the optical principles of image formation are also applicable to an electron microscope (in
which electron beams are focused by magnetic or electrostatic lenses), it is appropriate that
we examine these principles in some detail.

The optical paths and the magnification methods are somewhat different in SEM and TEM.
This will be addressed in Chapter 5, but like in the light optical microscope (LOM), the
electron path in the electron microscope is defined by the use of lens. The lenses are used
to change magnification and the focal point and the various apertures along the column
are used to change the contrast and resolution of the image. The path of ray in optical and
electron microscopes are shown in Fig. 4.5.

Clearly, an optical or an electron image is closely related to the corresponding object. In


1858 the Scottish physicist James Clark Maxwell, who also developed the equations relating
electric and magnetic fields that underlie all electrostatic and magnetic phenomena in
electromagnetic lens published three requirements for a perfect image [4.2]: 1. For each
point in the object, there is an equivalent point in the image, 2. the object and the image
are geometrically similar and 3. if the object is planar and perpendicular to the optic axis,
so is the image. Although the light image formed by a glass lens may appear similar to
the object, close inspection often reveals the presence of distortion, i.e. aberrations always
occur, as lenses are not perfect. Aberrations also occur in magnetic or electrostatic lens, so
the resulting electron images are not perfect either and needs to be corrected.

52
ELECTRON MICROSCOPY Instrumentation

Light optical microscope Transmission electron microscope Scanning electron microscope


Cathode Cathode
light source electron source electron source
Anode Anode

condenser lens condenser lens condenser lens

specimen specimen

objective lens objective lens objective lens

intermediate lens
scanning coil
intermediate image
intermediate image
projector lens(ocular) projector lens SE detector
specimen
scanning of electron probe

final image (on monitor)


final image (on eye or photo film) final image (on CCD)

Fig. 4.5: Path of ray in LOM, TEM and SEM compared.

Join the best at Top master’s programmes


• 3
 3rd place Financial Times worldwide ranking: MSc
the Maastricht University International Business
• 1st place: MSc International Business
School of Business and • 1st place: MSc Financial Economics
• 2nd place: MSc Management of Learning

Economics! • 2nd place: MSc Economics


• 2nd place: MSc Econometrics and Operations Research
• 2nd place: MSc Global Supply Chain Management and
Change
Sources: Keuzegids Master ranking 2013; Elsevier ‘Beste Studies’ ranking 2012;
Financial Times Global Masters in Management ranking 2012

Maastricht
University is
the best specialist
university in the
Visit us and find out why we are the best! Netherlands
(Elsevier)
Master’s Open Day: 22 February 2014

www.mastersopenday.nl

53
ELECTRON MICROSCOPY Instrumentation

Deflection of electrons in electrostatic and magnetic fields


To obtain the equivalent of a convex lens for electrons, the amount of deflection must be
increased with increasing deviation of the electron ray from the optic axis. For such focusing,
it is not possible to use refraction in the case of electrons (like glass is used in optical lenses),
as electrons are strongly scattered and absorbed by a solid. However, electrons, which have
electrostatic charge, are deflected by an electric field. Alternatively, one can use the fact
that the electrons in a beam are moving and the beam is therefore equivalent to an electric
current in a wire, which can be deflected by an applied magnetic field.

In a uniform electrostatic field, a beam of electron moves in a path analogous to a particle


in a gravitational field. This is illustrated in Fig. 4.6 in which a uniform potential (V ) is
applied between two plate electrodes A and B. The negatively charged electrons passing
through this field are subjected to a traverse displacement towards the positively charged
plate B and the electrons follow a parabolic path (Fig. 4.6). This deviation in the electron
path can be used to construct an electrostatic lens with appropriately shaped set of electrodes
such that the path of the beam is symmetrical about a common axis.

-V

A
electrons (- e)
equipotential field lines

B parabolic deflection

+V

Fig. 4.6: Motion of electron in an electrostatic field.

A ray diagram is schematically shown in the ideal case of focusing action corresponding to
thin-lens optics (Fig. 4.7). Here the thin-lens L focuses an incident beam from the object
O forming an image at O’. In drawing ray diagrams using geometric optics to depict the
image formation, the light or electron beam is represented by rays rather than waves and
so one ignores any diffraction effects. In the thin-lens ray diagram (Fig. 4.7), the bending
of rays is imagined to occur at the mid-plane of the lens. These special rays define the focal
length f of the lens and the location of the back-focal plane (dotted vertical line). One
special ray passes through the center of the lens, where the prism angle is zero and this ray
does not deviate from a straight line.

54
ELECTRON MICROSCOPY Instrumentation

Thin-lens(L)

back focal plane

F O’
principal axis
O

Fig. 4.7: Focusing action of a thin lens, where f = focal length.

Magnetic field with axial symmetry can also be used as electron lens, but in this case,
there is no analogy between the motion of electrons in the magnetic field and the paths
of a light ray. In a magnetic field, the force acting on the electron is always at right angles
to the motion of the electron (Fleming’s left hand rule [4.3]) and so the resulting path
of the electron is circular. If the electron enters the magnetic field at an angle to the field
direction, the resulting path of the electron (or indeed any charged particle) will be helical
(Fig. 4.8). Thus, an electron moving in a uniform magnetic field (H ) follows a helical path
and experiences a force normal to both the field and the original direction of motion of
the electrons. The electrons are then deflected in a plane normal to the field, intersecting
the lines of forces at right angles.

solenoid

electrons (- e)
2r H
V

P
Fig. 4.8: Helical path of an electron in a uniform magnetic field (H ).

Lenses for electron beam


The straightforward example of the uniform field between two parallel conducting plates
(Fig. 4.6) is good for deflecting an electron beam and used in cathode ray tubes, but are
not suitable for focusing. Different electrostatic lens designs are available and although some
electron microscopes use them, most modern microscopes use electromagnetic lenses. The
magnetic lenses do not require high voltage insulation and have somewhat lower aberrations
than electrostatic lenses.

55
ELECTRON MICROSCOPY Instrumentation

To focus an electron beam, the electromagnetic lens uses a magnetic field produced by a coil
carrying a direct current. As in the electrostatic case, a uniform field (applied perpendicular
to the beam) would produce overall deflection but no focusing action. So, to obtain focusing
a field with axial symmetry is needed, which can be generated by a short coil. Fig. 4.9 shows
the focusing action of a short magnetic lens. This is different from the solenoid case (shown
in Fig. 4.7) because here both the object and the image points lie outside the magnetic field.

Vr
X
Hr Hz
Vz O’
θ Vф Z
O

Fig. 4.9: Electron path and focusing action in a short magnetic lens.

56
ELECTRON MICROSCOPY Instrumentation

Shielding the lens coils with soft iron pole piece concentrates the magnetic field over a
very small region so that the radial component of the field (Hr) becomes very large at the
center and it is this component and not the axial component (Hz) which mainly causes the
deviation of the moving electron from its flight path. However, the field in this type of lens
is constant only over a small region close to the axis. An electron traveling along the coil axis
(z) is not deviated from the axis, as it feels no effect of the magnetic field. The symmetry
axis of the magnetic field is therefore the optical axis of the cylindrical lens. For non-axial
trajectories, the motion of the electron is more complicated. These electrons are not only
deviated from their path by the magnetic field but their trajectory also gets a rotational
component (f) as the electron passes through the axial-symmetric magnetic lens. Moreover,
as the direction of H changes continuously (so also the magnetic force), it causes focusing
of the electron beam. The focal length (f  ) of the lens is proportional to the magnetic field
and the electron velocity and thus can be changed simply by varying the lens current. All
magnetic lenses are convergent whether the object is in the magnetic field or not.

Aberration and defects in electron lenses


In an ideal electron optical system, all rays from a point in the object plane would converge
to the same point in the image plane forming a clear image. In real systems, however, errors
in the image occur because of imperfections in the optical system. The influences, which
cause different rays to converge to different points, are called aberrations. Alternatively, it
can be said that aberration result when the optical system misdirects some of the rays from
the object. Aberration occur both in light optical and electron optical systems. They are of
different kinds and are primarily caused by defects in the lenses but can also originate due
to other imperfections in the optical system. The main lens aberrations are: 1. Chromatic
aberration, 2. Spherical aberration, 3. Astigmatism, 4. Curvature of field, 5. Distortion and
6. Coma.

Spherical and chromatic aberrations are on-axis aberrations, i.e. these aberrations happen to
rays that are close to the optical axis. While off-axis aberrations include: Curvature of field,
Coma, Astigmatism and Distortion.

57
ELECTRON MICROSCOPY Instrumentation

Chromatic aberration
The term chromatic aberration is adopted from light optics. Owing to the dispersion2 in
glass lenses, light rays with different wavelengths are focused at different focal points. The
electron beam generated at the electron source in the microscope usually have a slight spread
in its energy (i.e. consists of electrons with different wavelengths, just like in the beam of
nonmonochromatic light). It is then easy to imagine that faster and slower electrons are
influenced differently by the electromagnetic field of the magnetic lens in the microscope.
For an electron, the de Broglie wavelength depends on the particle momentum and therefore
on its kinetic energy E and the focusing power of a magnetic lens depend inversely on the
kinetic energy. Therefore, if electrons are present with different kinetic energies in the beam,
they will be focused at different distances from the lens. Any fluctuation in the acceleration
voltage in the electron microscope therefore, leads to chromatic aberration. Thus, for any
image plane, there will be a chromatic circle of confusion rather than a point focus. Ray
diagram in Fig. 4.10 illustrates the change in focus and the circle of confusion resulting from
chromatic aberration. The radius of the circle of confusion is proportional to Cc .a.DE/E,
where Cc is the coefficient of axial chromatic aberration, DE/E the energy spread in the
electron beam and a the angle-limiting aperture subtended at the lens.

circle of least high-energy e¯


confusion

electrons from F F low-energy e¯


principal axis
Source
E (E-∆E)
(Energy spread ∆E)
circle of least
confusion

Thin-lens

Fig. 4.10: Chromatic aberration.

The chromatic aberration of entirely magnetic systems is unavoidable but they may be
decreased by using a smaller energy spread and a high acceleration voltage of the electron
beam. The chromatic aberration coefficient Cc is about the same magnitude as the focal length
f for weak lenses and about 0.6 times the focal length for strong lenses. In strong lenses,
the specimen will be immersed in the lens field.

58
ELECTRON MICROSCOPY Instrumentation

Spherical aberration
The imperfections in the shape of glass lenses cause spherical aberration in optical systems
but they can be corrected for example by using pairs of lenses (or doublets as they are
called) with different shapes and different kinds of glass. Electron lenses are inherently poorer
compared to the optical lenses and symmetrical round electromagnetic lenses unavoidably
suffer from spherical aberrations, having always a positive spherical aberration coefficient Cs.
The effect of spherical aberration can be defined by means of a diagram that shows electrons
arriving at a thin lens after traveling parallel to the optic axis but not necessarily along it
(Fig. 4.11). Those rays that arrive very close to the optic axis (known as paraxial rays) are
brought to a focus at the point F on the Gaussian image plane, i.e. at the focal length.
When spherical aberration is present in the electron optic system, electrons arriving at an
appreciably large radial distance from the optic axis are focused to a different point F1 located
at a shorter distance from the center of the lens (Fig. 4.11). At the Gaussian image plane,
the circle of confusion has radius proportional to Cs.a3, where a is the maximum angle
of the focused electrons. The aberration coefficient is in the same order of the magnitude
of the focal length. If the object is immersed in the lens field, the minimum value of the
aberration coefficient is half of the focal length. It diminishes when stronger lenses are used.
Spherical aberration arises because the focus depends on the axial distance of the electron
path. The outermost rays are focused more strongly than those close to the principal axis
and therefore, they can be eliminated by using an aperture in front of the objective lens
(shown in grey in Fig. 4.11).

aperture
smallest circle
of confusion

F1 F
paraxial
principal axis
rays Gaussian
focus

Thin-lens

Fig. 4.11: Spherical aberration

Fig. 4.11 directly represents the situation in a SEM where the objective lens focuses a near-
parallel beam into an electron probe of very small diameter at the specimen. Because the
spatial resolution of the electron image cannot be better than the probe diameter, spherical
aberration limits the spatial resolution in the SEM image. However, if the specimen is
advanced toward the lens (i.e. reducing the working distance in SEM), the disc of confusion
gets smaller and at a certain location (represented by the thin vertical line in Fig. 4.11) it
has the minimum diameter.

59
ELECTRON MICROSCOPY Instrumentation

In the case of TEM, a relatively broad beam of electrons arrives at the specimen and an
objective lens simultaneously images each object point. As in SEM, the electromagnetic lens
behave the same way and the further off axis the electron is, the more strongly it is bent
towards the axis due to spherical aberration. As a result, a point in the object is imaged as a
disk at the Gaussian image plane and this limits resolution in TEM. Spherical aberration is
important in the objective lens because it degrades the detail that the TEM can resolve; all
the other lenses (i.e. the projector lenses) magnify any error introduced by the objective lens.

Need help with your


dissertation?
Get in-depth feedback & advice from experts in your
topic area. Find out what you can do to improve
the quality of your dissertation!

Get Help Now

Go to www.helpmyassignment.co.uk for more info

60
ELECTRON MICROSCOPY Instrumentation

Astigmatism
All discussions on aberration, so far assumed complete axial symmetry of the magnetic field
that focuses the electrons in an electromagnetic lens. In practice, however, lens pole-pieces
cannot be machined with perfect accuracy and the pole-piece material may be inhomogeneous,
resulting in local variations in the relative magnetic permeability. The result is that the image
plane for object lying in one direction will be different from the image plane for object
lying in another direction. This difference in magnetic field will give rise to a difference
in focusing power of the lens hence there will be no sharp image plane but only a plane
of least confusion between two sharply focused images. The image will then suffer from
so-called astigmatism. The term astigmatism comes from the Greek α- (a-) meaning without
and στ ί γμα (stigma) a mark or spot. In an optical system with astigmatism, the rays that
propagate in two perpendicular planes have different foci. Therefore, if the optical system
with astigmatism is used to form an image of a cross, the vertical and horizontal lines will
be in sharp focus but at two different distances (Fig. 4.12).

Meridonal M
plane
S

Sagittal
plane
P

Fig. 4.12: Astigmatism

The aberration is manifested by the off-axis image of a specimen point appearing as a line or
ellipse instead of a point. Depending on the angle of the off-axis rays entering the lens, the
line image may be oriented in either of two different directions, meridional3 (tangential) or
sagittal (equatorial). Astigmatism can be compensated for by a so-called stigmator. In electron
optics, two quadrupoles mounted in an angle of 45° serve as a stigmator. In practice, one
cannot predict the azimuthal direction corresponding to the smallest or the largest focusing
power of an astigmatic lens. Therefore, the direction, as well as the strength of the stigmator
correction must be adjustable.

61
ELECTRON MICROSCOPY Instrumentation

Coma aberration
The name coma originates from comet. Coma (or comatic aberration) is a modification of
spherical aberration and generally occurs if the electron beam is off axial and/or oblique
to the optical axis (Fig. 4.13). Electron beams exiting from a point that is not located at
the optical axis on the object plane, do not come to one point on the image plane after
passing through the objective lens. This produces a comet-shaped (cone-shaped) image. This
phenomenon is called (off-axial) coma aberration. Coma is an aberration, which is inherent
to the electron microscope. The axial coma aberration in electron microscope arises from
asymmetry of the pole piece bore of the lens, magnetic non-uniformity of the pole piece
material, electron charging on the aperture, etc.

While in light optics, coma can be described by one real aberration coefficient (Bc), this is
not possible in electron optics. In the presence of a magnetic field, which causes an electron
to follow a helical trajectory, coma comprises of two contributions, radial (or isotropic)
and azimuthal (or anisotropic). Hence, in general two coefficients are necessary to describe
coma of a round magnetic lens. However, coefficient of coma (Bc) in electron optics can be
expressed by a single complex quantity, whose real part represents the radial coma and the
imaginary part the azimuthal coma.

P1
P2 Coma
P3

no Coma

P Thin-lens

Fig. 4.13: Coma

Comparing spherical aberration Cs with coma Bc reveals a fundamental difference. While the
impact of Cs depends solely on the geometrical ray parameters, the effect of coma Bc depends
on the location of the object point as well. This simply means that coma causes each object
points to be imaged differently. As Bc scales with the distance of the object point from the
optic axis and it vanishes when the object point is on the optic axis, it is different than Cs
which affects all object points equally.

62
ELECTRON MICROSCOPY Instrumentation

Electron micrographs are affected by coma when the direction of illumination does not
coincide with the direction of the true optical axis of the electron microscope. Hence,
elimination of coma is of great importance in high-resolution imaging work in TEM.
Coma aberration can be avoided by proper adjustment of the beam. The blur due to the
aberration moves to one side when the object is off axis, so that the image seems to have
a tail. To reduce the effect, methods for coma-free alignment have been developed, which
was originally proposed by Zemlin et al. [4.4]. By employing digital cameras and image-
processing systems, the process of coma-free alignment can be performed automatically in
modern TEM.

Brain power By 2020, wind could provide one-tenth of our planet’s


electricity needs. Already today, SKF’s innovative know-
how is crucial to running a large proportion of the
world’s wind turbines.
Up to 25 % of the generating costs relate to mainte-
nance. These can be reduced dramatically thanks to our
systems for on-line condition monitoring and automatic
lubrication. We help make it more economical to create
cleaner, cheaper energy out of thin air.
By sharing our experience, expertise, and creativity,
industries can boost performance beyond expectations.
Therefore we need the best employees who can
meet this challenge!

The Power of Knowledge Engineering

Plug into The Power of Knowledge Engineering.


Visit us at www.skf.com/knowledge

63
ELECTRON MICROSCOPY Instrumentation

Curvature of field and Distortion


A simple glass lens focuses image points from an extended flat object onto a spherical surface.
The nominal curvature of this surface is the reciprocal of the lens radius and is referred to
the Petzval Curvature of the lens. Therefore, a flat object normal to the optical axis cannot
be brought properly into focus on a flat image plane (Fig. 4.14). This is known as curvature
of field aberration. Due to the curvature of field the image of a flat object does not lie in
a plane, therefore when the center is in focus, the extremes are blurred. This aberration
can be compensated for by curving the viewing screen (as in large TV screens or drive-in
theater screens). Curvature of field is not a serious problem in the TEM or SEM, because
the angular deviation of electrons from the optic axis is very small. This result in a large
depth of focus and the image remains acceptably sharp as the plane of viewing is moved
along the optic axis.

Extended
Object

Thin-lens Image plane


Object plane

Fig. 4.14: Field curvature.

Further, when spherical aberration is present in a lens (i.e. Cs.a3), the image magnification
will vary in proportion to the cube of the distance of the image point from the axis.
The result is a distorted image. There are two kinds of distortions: barrel and pincushion
(Fig. 4.15) that are rather common in TEM, especially at low magnifications. The barrel
distortion occur when the magnification decreases and the pincushion distortion when the
magnification increases with distance.

Fig. 4.15: Pincushion and barrel distortions.

64
ELECTRON MICROSCOPY Instrumentation

Correcting chromatic and spherical aberrations


In electron microscope, the most important focusing defects are caused by lens aberrations
and they reduce spatial resolution of the image. In 1936, Otto Scherzer showed that every
rotational symmetric electromagnetic field acting as a lens on an electron beam has intrinsic
lens aberrations [4.5]. This fundamental theorem has come to be known as Scherzer’s
Theorem. Two kinds of axial aberrations that leads to image blurring even for object points
that lie on the optic axis are spherical and chromatic aberrations. Both of these aberrations
seriously degrade the performance of the electron microscope and limit its resolution. These
aberrations in electron microscopes cannot be corrected by combining rotational symmetric
lenses (as in light optics). Scherzer was the first to suggest that the most promising way to
get rid of lens aberrations in electron optics is to introduce non-rotational symmetry in the
electron path. By doing so, one of the preconditions (see Scherzer Theorem) that render
aberrations unavoidable would be eliminated. Around 1948 he designed and constructed
the first electrostatic corrector using three octupole elements to compensate for the spherical
aberration of the objective lens. Although spherical aberration of the lens was corrected
the microscope’s resolution could not be improved owing to instabilities in the mechanical
alignment of the correctors.

In modern TEM, the chromatic aberration of the electron optical system is compensated
by incorporating multipole correctors (e.g. with crossed electric and magnetic quadrupoles
acting as first-order Wien filters [4.6]) in the system. However, owing to the limit of the
maximum electric field strength that can be achieved in such quadrupole pairs, it is difficult
to correct chromatic aberrations in systems with acceleration voltage above 300 kV.

65
ELECTRON MICROSCOPY Instrumentation

Transmission electron microscopy is a useful real space analysis method. Considering the
wavelength of the electrons at an acceleration voltage of 200 kV is roughly 25 pm and is
sufficiently small to resolve individual atoms in periodic objects, electron microscopes can
be used for high-resolution lattice resolved images of crystalline solids. However, because
of the imperfect lenses used, for a long period the point-to-point resolution was limited to
0.2 nm in TEM operating at 200 kV [4.7]. The point-to-point resolution of TEM images
of a sample is expressed in terms of the phase-contrast transfer function (PCTF) when the
sample is assumed as a weak-phase object [4.8]. Digital Fourier transform (DFT) patterns
of the object give the Scherzer limit corresponding to the point-to-point resolution and the
information limits determined by the convergence of incident electrons and the chromatic
aberration of the instrument. Resolution limits in modern TEM will be discussed in more
details in Chapter 6.

66
ELECTRON MICROSCOPY Instrumentation

4.3 ELECTRON AND X-RAY DETECTORS


Detection of electrons and other signals
In a light microscope, images can be viewed directly by the human eye. However, electrons
cannot be observed directly and therefore, electron microscopes (both TEM and SEM) need to
use an interface, such as a fluorescent screen, photographic plate or electronic display to make the
image visible to the human eye. Moreover, as interaction of electrons with the matter produces
various other signals (as discussed in Chapter 3) which are used in electron microscopes for
analytical purposes, therefore, many different kind of detectors are employed in the microscopes.

Electron detection and image recording in TEM


Older transmission electron microscopes used a fluorescent screen containing phosphorous for
directly viewing the electron image. The fluorescent screen, which emits light when impacted
by the transmitted electrons was used for real-time imaging and adjustments and a film camera
to record permanent, high resolution images in TEM (electrons have the same influence on
photographic material as light). The fluorescent screen was under vacuum in the projection
chamber of the microscope and the live image could be observed through a glass window, using
a binocular magnifier if needed. Photo film coated with photographic emulsion has been used
as an image recording material in the TEM since the invention of the electron microscope.
The specimen image could be projected on the photo film placed below the fluorescent screen
when the screen was lifted up, allowing the use of the photo camera for recording. However,
in modern TEM many different systems are possible for viewing and recoding image and
diffraction patterns. These are photographic film, real time video charge-coupled device (CCD)
camera, slow scan CCD camera and imaging plate (IP). The TV or the CCD cameras are
image detectors different from photo film and they have high sensitivity. However, their image
quality was not comparable to photo films when they were first introduced in TEM in the
late 1970s. The imaging plate, which was developed as an image recording material for x-ray
radiography (as they had higher sensitivity than conventional x-ray film) had higher image
quality than the other electrical detectors. The IP were applied to TEM image recording in
the early 1980s, when they had better resolution than the CCD. Since both the CCD and IP
produce image data as electrical signal it is convenient to digitally store and process the image.
CCD have the added advantage that the image can be viewed live on a computer display,
while the image recorded on IP must be first scanned and digitized before they are visible. For
many decades, film has been the recording medium of choice for electron microscopy because
of its better resolution and the electron microscopists used specialized electron micrograph
films to record the TEM image for analysis, until new developments (which include electron
tomography and cryogenic electron microscopy) pushed for digital imaging and recording.
The modern instruments rely primarily on solid-state imaging devices (such as CCD) for
image capture and now they mostly have replaced the photo films. Recent years has seen a
widespread use of the CCD in all fields of TEM.

67
ELECTRON MICROSCOPY Instrumentation

Scintillator

(convert electron to photon)

Fiber optics
(light transfer)

CCD chip CMOS chip


(electronically transfer signal) (electronically transfer signal)
Read-out Read-out
a. b.

Fig. 4.16: Detectors in TEM: a) slow scan CCD, b) direct detection with DDD.

Currently available phosphor/fiber-optic CCD cameras (Fig. 4.16a) do provide immediate


readout and much lower noise than photo films and thus allow for adequate performance in
many applications, including electron diffraction. However, due to intrinsic light scattering
within the phosphor the spatial resolution of the CCD camera is limited. This limitation
is fundamental as the CCD use indirect detection method based on scintillator screen. A
scintillator material over the image detector element is used to convert high-energy incident
electrons to light, which creates the charge in the underlying CCD element (Fig. 4.16a). The
scintillator not only introduces some loss of resolution, the conversion process also decreases
the efficiency with which electrons contribute to image contrast. This can be critical in
applications that are sensitive to damage by the electron beam, such as cryogenically prepared
samples of delicate biological materials, where it is essential to extract the maximum amount
of information from a faint, noisy signal before the sample is destroyed.

To realize the full imaging potential in the TEM, a new imaging detector called the Direct
Detection Device (DDD) is recently developed (Fig. 4.16b). The DDD is based on direct
detection of electrons, thus avoiding the intermediate light conversion step required in
CCD. It can detect high-energy electrons with a high signal-to-noise ratio ensuring single
electron sensitivity. The direct electron detectors promise significant improvement in image
resolution and contrast, particularly in signal-limited applications. Eliminating the scintillator
with a direct electron detector improves image resolution and increases detector efficiency
by up to three times. Two types of direct detectors have been developed [4.9]: one based
on hybrid technology employing separate pixelated sensor and readout electronics connected
with hybrid pixel detectors (HPD) and the other a monolithic active pixel sensor (MAP),
where sensing and readout are all in one plane. Due to the small pixel size (5μm), the
spatial resolution of the DDD exceeds any current imaging detectors in TEM.

68
ELECTRON MICROSCOPY Instrumentation

Electron detection and imaging in SEM


In the TEM, image is formed with electrons which are transmitted through the thin sample
in the forward direction: either using the direct beam (which produces the so-called bright
field image), or with the diffracted beam (dark field image). However, in a SEM, the bulk
sample mostly scatters electrons in the backward direction and the images are formed using
the secondary or the backscatter electrons, which are produced by the interaction of the
incident electron beam with the sample. Most common detector for detecting secondary
electron is the Everhart-Thornley detector (ETD), named after its designers, Thomas E.
Everhart and Richard F.M. Thornley [4.10]. ETD can also detect backscattered electrons.

69
ELECTRON MICROSCOPY Instrumentation

Everhart-Thornley Detector
In the SEM, a small electron probe is scanned in a raster across the surface of the specimen.
The most common imaging mode collects low-energy (< 50 eV) secondary electrons that
are ejected from the outer shell of the specimen atoms by inelastic scattering interactions
with the incident electron beam. Due to their low energy, these electrons originate within
a few nanometers of the sample surface [4.11]. Secondary electrons leaving the surface are
collected by the ETD, amplified and used to form the image of the area scanned by the
electron probe. However, owing to their low energies, the secondary electrons must be
attracted towards the SE detector using a positive bias on the grid in front of the detector.
As a result, the electron can travel in curved paths to the detector (Fig. 4.17). The ETD
primarily consists of a scintillator inside a Faraday Cage and a low positive voltage (typically
200 V) is applied to the Grid (Faraday Cage) to attract the relatively low energy secondary
electrons. Other electrons within the specimen chamber are not attracted by this low
voltage. They will only reach the detector if their direction of travel takes them to it. The
scintillator has a high positive voltage (10 to 15 kV) to accelerate the incoming electrons.
The accelerated secondary electrons have sufficient energy to cause the scintillator to emit
flashes of light/photons (Cathodoluminescence), which are conducted to a photomultiplier
placed outside the SEM vacuum chamber via a fiber-optic tube. The amplified electrical
signal output by the photomultiplier is displayed as a two-dimensional intensity distribution
that can be viewed, photographed and saved as a digital image. The ET secondary electron
detector can also be used in the SEM in backscattered electron mode by either turning off
the Grid bias or by applying a negative voltage. If a negative voltage is applied to the grid,
SE is rejected and only BSE (owing to their much higher energy) are detected by the ETD.
However, efficiency in this mode is low and better-backscattered electron images come from
dedicated BSE detectors rather than from using the ETD as a BSE detector.

Photomultiplier Incident beam


Output signal

Fiber optics e¯
Scintillator
(10-15kV) Grid
SE
(200V)

Bulk specimen

Fig. 4.17: Components of Everhart-Thornley detector and showing


how secondary electrons are collected by the ETD.

70
ELECTRON MICROSCOPY Instrumentation

It can be seen in Fig. 4.17 that the secondary electrons emitted toward the detector are only
collected by the ETD. The position of the specimen surface relative to the detector will
also affect how many electrons are collected by the detector and typically the ETD collect
much less than 50% of the secondary electrons. If the specimen has surface topography as
shown in Fig. 4.18, then SE coming from behind elevated regions (hills) are shadowed and
do not reach the detector. Therefore, the detector efficiency combined with topographic
features of the sample surface creates contrast in the image produced by the ETD. As the
ETD views the sample from one side, the sample faces looking away from the detector are
always shadowed (Fig. 4.18).

Fig. 4.18: Topographic contrast generation in ETD.

BSE Detector
The ETD, which is normally positioned at an angle with respect to the specimen and to one
side in the SEM vacuum chamber (Fig. 4.17 and 4.18), is quite inefficient for the detection
of backscattered electrons because few such electrons are emitted in the solid angle subtended
by the detector. Moreover, the positively biased detection grid in ETD has little ability to
attract the high-energy backscatter electrons. Dedicated backscatter detectors are therefore
used for BSE image formation in SEM. The high-energy backscattered electrons reflected
out of the bulk specimen travel in straight lines and are more likely to be scattered at high
angles. Therefore, the detector for BSE is placed directly above the sample (Fig. 4.19a).
The detectors are doughnut shaped (i.e. disc with a central hole) and are concentric with
the incident electron beam and therefore maximize the solid angle for the collection of
backscattered electrons. In a SEM, generally the BSE detector can be retracted out of
the beam when not in use. They are generally solid-state semiconductor diodes (although
scintillator type detectors similar to those used in the ETD are also available). Modern
detectors are usually segmented (Fig. 4.19b) either as annular rings or as radial segments
to allow contrast enhancement by mixing signal from different segments.

71
ELECTRON MICROSCOPY Instrumentation

The heavy atoms in the specimen (high atomic number elements) backscatter electrons more
strongly than light atoms (low atomic number elements) and therefore, the BSE image display
atomic number (Z) contrast with the area containing heavy elements appearing brighter.
The resulting BSE image contains less topographic information but they are useful to detect
contrast due to chemical composition difference, i.e. in different phases. Moreover, BSE
are less affected by the electric charge than SE and therefore, are more suitable for imaging
samples with insulating materials. Oxide particles, for example are imaged by BSE with less
noise, whereas, charging of oxide particles interfere in SE images.

When all parts of the segmented detectors are used to collect the backscattered electrons
symmetrically about the beam, Z-contrast is produced. However, strong topographic contrast
can be produced in BSE detectors by collecting backscattered electrons from one side above
the specimen using only some segments, which are asymmetrical. The resulting contrast
appears as illumination of the topography from that side. Semiconductor detectors can be
easily switched in or out to control the type of contrast produced and its directionality.

72
ELECTRON MICROSCOPY Instrumentation

B C
Incident beam A D
M
L E
e¯ K F
J G
I H

Magnetic Lens
A
pole piece C
B

BSE detector
BSE
A D B

Bulk specimen
C
a. b.
Fig. 4.19: BSE detector

In-lens or Through-the-lens Detectors


Traditionally, for imaging nano-scale inhomogeneities in specimen the transmission electron
microscope is used, as it is not limited by the interaction volume that is experienced in
electron-beam/specimen interaction by the scanning electron microscope. Recent development
in the detection systems in SEM, specifically in secondary and backscattered in-lens detectors,
makes it possible to use low energy electron beams in SEM to image nanoparticles. The
in-lens detectors (also sometimes called through-the-lens detector, TTL or TLD) are located
inside the objective lens in the electron column of the SEM (Fig. 4.20). They are either
looking-in from one side in the lens (TTL or TLD) or are arranged rotationally symmetric
around the optical axis (in-lens). Figure 4.20 schematically shows the principle of in-lens
or TTL detectors for secondary electron (SE) detection [4.12]. To image the sample surface
at high resolution, SE1 electrons that have experienced one scattering event only should
be detected. The SE1 electrons are generated in the upper range of the interaction volume,
and therefore, contain direct information of the sample surface. These electrons can be
detected very efficiently by the in-lens SE detector, whose detection efficiency results from its
geometric positioning in the beam path and the special lens design. Due to a sophisticated
magnetic field at the pole piece, the secondary electrons are collected with high efficiency.
In particular, at low voltages and small working distances, images with high contrast can
be obtained.

73
ELECTRON MICROSCOPY Instrumentation

Incident beam Incident beam


e¯ In-lens detector

Magnetic Lens Magnetic Lens


pole piece pole piece
TLD

Bulk specimen Bulk specimen


a. b.
Fig. 4.20: a) through-the-lens and b) in-lens detectors

Challenge the way we run

EXPERIENCE THE POWER OF


FULL ENGAGEMENT…

RUN FASTER.
RUN LONGER.. READ MORE & PRE-ORDER TODAY
RUN EASIER… WWW.GAITEYE.COM

1349906_A6_4+0.indd 1 22-08-2014 12:56:57

74
ELECTRON MICROSCOPY Instrumentation

EELS Detectors in TEM and STEM


Electron energy-loss spectrometry (EELS) involves measurement of the energy distribution of
electrons that have interacted with a specimen and lost energy due to inelastic scattering.
Relatively simple instrumentation can provide spectra with an energy resolution down to
a few milli-electron volts, sufficient to resolve vibrational as well as electronic modes of
energy loss. The detection and discrimination of electrons of different energy emerging from
the bottom surface of a thin-foil specimen in TEM is the basis for EELS. The electron
spectrometers can be magnetic, electrostatic or a combination of both [4.13]. The detector
mainly detects electrons that have lost energy due to ionization of individual atoms or by
plasmon excitation in the specimen and the typical energy loss in this process are in the range
of 102–103 eV. To detect the ionization edge, a detector resolution of ~1 eV is sufficient.
For this reason and the low cost of construction, the EELS detectors in AEM are mainly
the magnetic prism type spectrometer. The simplest form of the energy-loss system consists
of a magnetic prism below the viewing chamber of a conventional TEM/STEM. Lifting
the viewing screen allows electrons to enter the spectrometer, where they are dispersed
according to their kinetic energy by the prism. It is the energy in the incident beam (E0)
minus any energy loss (DE) occurring in the specimen. Typical EELS spectrometers are
shown in Fig. 4.21. The magnetic prism spreads the electrons of different energies and the
electrons of specific energies are selected by a slit, and they then impinge on a scintillator-
photomultiplier electronic (Fig. 4.21a). Newer parallel collections of electrons of different
energies has replaced old serial collection (Fig. 4.21b) in modern microscopes and made
possible the so-called Energy-filtered transmission electron microscopy (EFTEM). Provided
the image aberrations are corrected by quadrupole (or even sextupole) lenses, the post-
column magnetic prism can also produce EFTEM images. In effect, EFTEM is a family
of technique, which include imaging and utilizes the property of the energy loss spectrum
to enhance image contrast through removal of chromatic aberration effects. Essentially an
energy-filtered electron beam is used for the image formation. However, it is a post filtering
process and should be distinguished from the online omega filtering in the TEM column.
The energy filter process also allows microanalysis and mapping – i.e. record and quantify
the electron-energy-loss spectra and produce maps to provide chemical analysis of thin-foil
samples at nanometer resolution.

75
ELECTRON MICROSCOPY Instrumentation

EELS (serial collection) EELS (parallel collection)


projector crossover projector crossover

view screen
view screen entry aperuture
entry aperuture

quadrupole lens
transverse
deflector

prism E - ∆E energy selection slit prism


spectrometer spectrometer parallel
E photo-multimplier deflector
beam trap aperture
scintillator

a. b.

Fig. 4.21: EELS spectrometers: a) serial collection of electrons and b) parallel collection and mapping

As electrons from the incident beam pass through a thin specimen, they interact with atoms
of the solid (see section 3.1). Many of the electrons pass through the thin sample without
losing energy (elastically scattered). A fraction undergoes inelastic scattering and loses
energy as they interact with the specimen and this leaves the sample in an excited state. The
material (specimen) then can de-excite by giving up energy, typically in the form of visible
photons, x-rays or Auger electrons. Each type of interaction between the electron beam and
the specimen produces a characteristic change in the energy and angular distribution of the
scattered electrons, as was discussed in chapter 3. The energy loss process is the primary
interaction event and all other sources of analytical information (x-rays, Auger electrons,
etc.) are secondary and a product of the initial inelastic event. Thus, EELS have the highest
potential yield of information per inelastic events. It also has a very high energy-resolution.

76
ELECTRON MICROSCOPY Instrumentation

A typical electron energy-loss spectrum is schematically presented in Fig. 4.22. The first
peak, which is the most intense for a very thin specimen, occurs at 0 eV and is called
the zero-loss peak. It represents electrons which have not been scattered in the specimen
(transmitted electrons) and which have been elastically scattered via interaction with the
atomic nuclei in the specimen. A prominent form of inelastic scattering in solids involves
plasmon excitation. In this low-loss region of the spectrum (< 50 eV) the peaks are formed
due the excitation of the outer-shell electrons (conduction electrons in metals and valence
electrons in semiconductors and insulators). It provides similar information to that provided
by optical spectroscopy, containing valuable information about the band structure and in
particular about the dielectric properties of a material (e.g., band gap, surface plasmons).
Signal intensities in the low-loss region are larger than in the high-loss region of the
spectrum. At higher energy losses (> 50 eV), where the number of inelastically scattered
electrons is much lower, the spectrum shows characteristic features called ionization edges
(due to their typical shape, a rapid rise followed by a more gradual fall). These edges are
the exact equivalent of an absorption edge in x-ray absorption spectrometry (XAS) and
arise from the same process. Very generally, when an x-ray (or an electron) strikes an atom
it excites a core electron that is either promoted to an unoccupied level, or ejected from
the atom. Both of these processes will create a core hole. If the electron dissociates, this
produces an excited ion as well as photoelectron and is studied by x-ray photoelectron
spectroscopy (XPS). In XPS, where electrons are excited from atomic core levels of a solid
into the adjacent vacuum, a chemical shift of the associated spectral peak represents a
change in core-level energy arising from the change in the effective charge on the atom.
In EELS, changes in the chemical environment of an atom can cause a similar shift in the
threshold energy of an ionization edge (up to a few electron volts). The edges are formed
when an inner-shell electron absorb enough energy from a beam electron to be excited to
a state above the Fermi level. However, at the edge threshold, the final state of the core
electron lies just above the Fermi level and below the vacuum level, so in this case any
chemical shift includes not only the effect of electron transfer to the atom but also change
in valence-band width, due to redistribution of the coordinates of neighboring atoms. This
second factor is predominant in the case of metals and it implies that, while the core-loss
fine structure reflects the density of unoccupied states above the Fermi level, the chemical
shift carries information about the occupied states [4.14].

77
ELECTRON MICROSCOPY Instrumentation

plasmon peaks

zero loss
peak

Intensity

ionization edge

0 50 100 150 200


Energy Loss (eV)

Fig. 4.22: The schematically drawn EELS spectrum: spectra from a thin region (blue line) and
a thicker area (orange line) of the specimen are shown. The zero-loss peak is matched in
height. The plasmon peaks occur at multiples of the plasmon energy and the broad feature
starting around 100 eV shows an ionization edge.

This e-book
is made with SETASIGN
SetaPDF

PDF components for PHP developers

www.setasign.com

78
ELECTRON MICROSCOPY Instrumentation

Detection of x-ray signal


Characteristic x-rays emitted from the sample due to the interaction of the incident electron
beam with the specimen are used in electron microscopes to analyze the chemical composition
of the various phases present in the sample. The x-rays are collected by specially designed
detectors and the spectra analyzed according to their energy or wavelength. The energy
dispersive spectroscopic analysis (EDS) is more common, and used both in SEM and in TEM,
while wavelength dispersive spectroscopic analysis (WDS) is available only in SEM. WDS
was initially developed as an independent standalone instrument known as electron micro
probe analyzer (EPMA), which operates similar to a scanning electron microscope [4.15].
Recently, WDS is also available in SEM with a much more compact detector system and
used as an analytical tool particularly useful for light element analysis and for separating of
overlapping x-ray peaks. The x-rays having characteristic energies/wavelengths can be either
detected using a solid-state energy dispersive spectrometer detector or a diffracting crystal in
tandem with a gas-filled proportional counter wavelength dispersive spectrometer detector.
Each detector type and the analysis method have its merits and drawbacks.

Energy Dispersive Spectroscopy


In energy dispersive analysis, detection and dispersion are a single operation and a proportional
counter or different types of solid-state detectors (Si(Li), Ge(Li), SDD) are used to detect and
collect the incoming x-ray signal. The most common energy dispersive spectrometer detector
till recently was the Si(Li) detector, which require cryogenic temperatures and are cooled
with liquid nitrogen. However, newer microscopes are now often equipped with silicon drift
detectors (SDD) with Peltier cooling systems. They all share the same detection principle:
An incoming x-ray photon ionizes a large number of detector atoms with the amount of
charge produced being proportional to the energy of the incoming photon. Semiconductor
detectors are solid-state devices that operate essentially like ionization chambers, where the
charge carriers are not electrons and ions, as in the gas counters, but electrons and holes.
X-rays interact with the semiconductor to produce electron-hole pairs and such pairs
are swept away under the influence of an electric field (Fig. 4.23). Since the number of
electron-hole pairs generated is proportional to the energy of the interacting radiation, by
converting the resulting charge into a voltage pulse whose amplitude is proportional to the
x-ray energy, a sequence of detected x-rays can be transformed into a spectrum of counts
vs. energy, thereby forming an energy-dispersive spectrum.

79
ELECTRON MICROSCOPY Instrumentation

X-ray
holes

Si (Li)
(Li doped Si)

electrons

Fig. 4.23: Semiconductor x-ray detector principle

Silicon Lithium detector


Si(Li) detectors are Li doped Si semiconductor, also known as Silicon (Lithium drifted)
detectors. They convert incoming x-rays into electron-hole pairs at a rate of one pair for each
3.8 eV of initial energy [4.16]. The crystal is thick enough (3–5 mm) to capture virtually
100% of the x-rays in the energy range of interest.

Incident beam

LN2
Dewar

Si (Li)
Window Detector
(Be or polymer)
X-ray
Cold finger
FET preamplifier
Bulk specimen

Fig. 4.24: Si(Li) detector for EDS measurement

80
ELECTRON MICROSCOPY Instrumentation

The construction details and the main components of the detector system are presented in
Fig. 4.24. The window (beryllium or polymer-based) in front provides a barrier to maintain
vacuum within the detector whilst being as transparent as possible to low energy x-rays. The
polymer-based thin windows can be made much thinner than the Be windows and therefore
are transparent to much lower energy x-rays, and can allow detection of x-rays down to
100eV. The Si(Li) crystal (which is a silicon junction type p-i-n diode) is placed just behind
the window and convert the x-rays to electron-hole pairs and the FET (field effect transistor)
convert current into a voltage. The lithium-drifted center part forms the non-conducting
i-layer, where Li compensates the residual acceptors, which would otherwise make the layer
p-type. To obtain sufficiently low conductivity, the detector must be maintained at low
temperature and liquid-nitrogen cooling must be used for the best resolution.

Free eBook on
Learning & Development
By the Chief Learning Officer of McKinsey

Download Now

81
ELECTRON MICROSCOPY Instrumentation

Silicon drift detector


Silicon drift detector is a recent advancement in solid-state x-ray detection and use high
purity Si instead of Si(Li) as the semiconductor component. The innovative design and
construction of the SDD not only offer advantage of being liquid nitrogen free and excellent
energy resolution at high count rates, but large active area of the detector also allows
collection of large amounts of data in shorter time periods, at lower excitation voltages,
and under normal SEM imaging conditions. Like other solid-state x-ray detectors, silicon
drift detectors measure the energy of an incoming photon by the amount of ionization it
produces in the detector material. The high purity of Si allows for the use of Peltier cooling
instead of the traditional liquid nitrogen used in the Si(Li) detectors. An SDD has less
electronic noise than a comparable planar detector and this gives the SDD better energy
resolution at moderate count rates and much better energy resolution at high-count rates.

The SDD components are same as in the Si(Li) detector (shown in Fig. 4.24), consisting of
the window, the detector chip, the FET and the cooling system and differ mainly in the type
of the cooling device and the chip construction. As the SDD operate at slightly below 0°C
they are cooled by Peltier (thermoelectric) devices bound to the SDD sensor. The details of
the SDD chip, is shown in Fig. 4.25. In the slightly older detector designs, the collection
electrode is centrally located with an external FET convertor. Newer designs integrate the
FET directly into the chip, which greatly improves energy resolution and throughput.

Fig. 4.25: Cut section through an SDD detector chip used for EDS measurement

82
ELECTRON MICROSCOPY Instrumentation

The SDD sensor has a large contact area on the entrance side facing the incoming x-rays.
On the opposite side, a small anode contact is centrally located and surrounded by a
number of concentric drift electrodes (Fig. 4.25). When a bias is applied to the SDD and
the detector is exposed to x-rays, it converts each x-ray detected into an electron-hole cloud
with a charge that is proportional to the characteristic energy of that x-ray. When several
x-ray quanta hit the detector at the same time but at different positions, they are separated
under the electric field. These electron-hole pairs are raised into the conduction band of the
silicon semiconductor leaving behind the holes that behave like free positive charges within
the sensor. The electrons are then ‘drifted’ down a field gradient applied between the drift
rings to be collected at the anode. The electrons from the different hits of the x-ray drift
different path lengths due to the diffusion of the charge cloud on their way to the anode
thereby each hit is separated and recorded as a different x-ray count.

Wavelength Dispersive Spectroscopy


Wavelength dispersive x-ray spectroscopy was originally used in the EPMA, developed
for accurate determination of chemical composition in micro volumes of bulk specimens.
Although the instrument operates in the same principle as SEM, EPMA is a standalone
analytical tool. The detector in the EPMA is relatively large but recent advancement made
it possible to design new type of WDS compact enough to fit in the SEM.

X-rays are produced in the interaction volume immediately below the impact zone of the
finely focused electron beam in a SEM. However, only a very small fraction of all x-rays
is at the proper “take-off” angle to head up into the spectrometer (a much smaller fraction
compared to an EDS detector mounted close the sample). The key feature of the WDS
is a crystal-focusing spectrometer, which is used to isolate x-ray signals according to their
wavelength for the quantitative analysis. Once x-rays are generated in the specimen by the
electron beam, they are selected using an analytical crystal and directed into a gas-filled
proportional counting tube. The x-rays absorbed by gas molecules (a special gas mixture
P10: 90% Ar, 10% CH4) in the tube eject photoelectrons and produce a secondary cascade
of interactions, yielding an amplification of the signal (typically103–105 times) which is
handled by the electronics.

83
ELECTRON MICROSCOPY Instrumentation

electron beam
Johansson diffractor
θ θ

X-ray
Detector
Bulk specimen

Rowland circle

Fig. 4.26: Geometry of WDS detector showing the principle of x-ray


wavelength selection by a crystal (Johansson diffractor – a layered crystal
bend to radius 2R with the inner surface ground to R, the radius of the
Rowland circle).

www.sylvania.com

We do not reinvent
the wheel we reinvent
light.
Fascinating lighting offers an infinite spectrum of
possibilities: Innovative technologies and new
markets provide both opportunities and challenges.
An environment in which your expertise is in high
demand. Enjoy the supportive working atmosphere
within our global group and benefit from international
career paths. Implement sustainable ideas in close
cooperation with other specialists and contribute to
influencing our future. Come and join us in reinventing
light every day.

Light is OSRAM

84
ELECTRON MICROSCOPY Instrumentation

The geometry between the specimen and the analytical crystal is maintained such that the
take-off angle remains constant. When x-rays encounter the analytical crystal at a specific
angle q, only those x-rays that satisfy Bragg’s Law are reflected and a single wavelength is
passed on to the detector. The principle of wavelength selection is schematically shown
in Fig. 4.26. The wavelength of the x-rays reflected into the detector may be varied by
changing the position of the analyzing crystal relative to the sample (Fig. 4.27), then
the x-ray source-to-crystal distance is a linear function of the wavelength. Consequently,
x-rays from only one element present in the specimen can be measured at a time on the
WDS and the position of a given analytical crystal must be changed in order to adjust to
a wavelength characteristic of another element in the specimen. The analytical geometry
shown in Fig. 4.26 is constructed by bending the analytical crystal to a radius of 2R, where
R is the radius of the focusing circle, known as the Rowland circle. Typically, in EPMA
there are 4–5 spectrometers mounted in sequence around the specimen chamber. In each
spectrometer, a crystal with different lattice spacing is used so that the spectrometer can
reach all elemental wavelengths of interest for measurement and will optimize performance
in different wavelength ranges.

L1 < L2 < L3

e¯ beam e¯ beam e¯ beam


Crystal Cry
al stal
yst
Cr
Detector

X-ray X-ray
X-ray
Detector
Sample Sample Sample

Detector

Rowland circle Rowland circle Rowland circle

Sample of Crystal distance L1 Sample of Crystal distance L2 Sample of Crystal distance L3

Fig. 4.27: Different positions of the analytical crystal and the detector on the Rowland circle. The geometry
is such that the distances from the crystal to the detector and from the source to the crystal are similar (L) at
each position and where L1 < L2 < L3.

More typically, in SEM there is only a single WD spectrometer. The original WDS
spectrometers were designed by applying the Rowland circle technique. In advanced WDS
spectrometers in SEM, a parallel beam spectrometer is used that considerably enhanced
the x-ray collection efficiency in the low energy range. In the latest WDS spectrometers,
a parallel beam spectrometer with a hybrid x-ray optic are used to facilitate high x-ray
collection efficiency in high as well as low energy parts of the spectrum.

85
ELECTRON MICROSCOPY Instrumentation

Since flat diffractors (analytical crystals) are not capable of converging or focusing the
diverging x-rays, they yield a low count rate at the detector. Moreover, in the Rowland
circle WDS with Johansson diffractor (bend crystal) the x-ray collection efficiency depends
on the sample position. The optical image is in focus when the sample is placed at the
proper working distance and Z position. In case this is not accurate, the sample will not
lie on the Rowland circle and the distance between the diffractor and the sample will not
be equivalent to the distance between the detector and the diffractor. This limitation is
overcome in parallel beam WDS, where the x-ray optic is located close to the sample to
change the divergent x-rays from the sample into a parallel beam. As a result, the Rowland
circle geometry is not necessary, thus simplifying the spectrometer design and integration
with the SEM. This approach ensures that a flat diffractor can be used, which is located at
a semi-infinite distance from the sample (Fig. 4.28). The parallel beam WDS gives better
sensitivity compared to conventional WDS systems, especially at low energies. The x-ray
optics work by the principle of total reflection, which occurs off a smooth surface if the
incidence angle, is less than a critical angle. The critical angle depends on the mirror material
and the energy of the x-ray and it becomes smaller with increasing energy. For collimating
diverging x-rays, actually two x-ray optical techniques are available: poly-capillary optics and
grazing incidence. At higher energies (> ~2.5keV), a poly-capillary optic has greater efficiency
and at lower energies (< ~2.5keV), a grazing incidence optic has the greater efficiency. This
options lead to two WDS spectrometer designs: one for low energy and the other suitable
for high-energy spectroscopy.

electron beam Flat diffractor

X-ray
Detector

Hybrid optic

Bulk specimen
Fig. 4.28: A parallel beam WDS includes a hybrid optic consisting of both
polycapillary and grazing incidence optics that transforms the divergent
x-rays into a parallel beam.

86
ELECTRON MICROSCOPY Instrumentation

In parallel beam WDS systems, it is also important to make sure that x-ray optic and the
sample are aligned correctly, with the sample at the correct working distance in the SEM
to ensure that x-rays are incident on the diffractor. Any deviation from the correct working
distance decreases the detected x-ray intensity. This effect is greater for higher energy x-rays.
Optic positioning or maintaining the desired position of the sample can be done manually,
but this difficult alignment is time consuming. In modern SEM with piezo motor control
stage, the Z-axis can be precisely controlled automatically (software driven) to obtain
maximum intensity and optimal peak shape.

Recording Kikuchi pattern


In order to obtain micro-textural information from bulk crystalline specimen or thin layers,
the electron backscatter diffraction (EBSD) technique can be used in SEM. Single crystal
diffraction patterns provide more complete (although also a more complex) information

360°
than the ring patterns obtained from polycrystalline material, particularly, with respect to
the lattice type, symmetry, atomic arrangement and orientation in the crystal. EBSD is able

.
to analyze most of this information from bulk specimen, because when the electron probe

thinking
is small compared to the grain size of the polycrystalline specimen, the grain is a single
crystal for the electron probe.

360°
thinking . 360°
thinking .
Discover the truth at www.deloitte.ca/careers Dis

© Deloitte & Touche LLP and affiliated entities.

Discover the truth at www.deloitte.ca/careers © Deloitte & Touche LLP and affiliated entities.

Deloitte & Touche LLP and affiliated entities.

Discover the truth at www.deloitte.ca/careers


87
ELECTRON MICROSCOPY Instrumentation

incident electron

thick specimen

Fig. 4.29: Inelastic scattering in thick specimen is strongly forward peaked.

Accelerated electrons in the primary beam in an electron microscope can be diffracted by


atomic layers in crystalline materials as discussed in Chapter 3. Elastic scattering mainly occur
during high-energy electron diffraction in thin specimen (as in TEM). When the specimen
is thick enough, inelastic scattering will also take place. The inelastically scattered electrons
travel in all directions but their distribution peaks in forward direction (Fig. 4.29). In thick
specimen, dynamical diffraction becomes important and electrons are scattered more than
once, and this has important effects on diffraction intensities and image contrast, which
produce additional features in diffraction patterns. In the bulk specimen used in SEM,
the backscattered electrons (BSE) are inelastically scattered producing dynamic diffraction.
Phenomenologically, the EBSD pattern formation is a two-step process. Electrons which
have been inelastically scattered once can subsequently be diffracted, but only if they are
now travelling at the Bragg angle (qB) to a set of crystal planes. Two sets of electrons will be
able to do this, those at + qB and those at – qB (Fig. 4.30a). Because the electrons travel in
all directions, for each set of plane for which the Bragg condition is satisfied, the diffracted
beam lies on the surface of a cone whose axis is normal to the diffracted plane. If the cones
intersect the Ewald sphere4 or a screen placed in front of the specimen, it gives rise to a
pair of line on the screen (Fig. 4.30b). The paired arrays of lines in the diffraction pattern
are known as Kikuchi lines or Kikuchi bands [4.17].

88
ELECTRON MICROSCOPY Instrumentation

The origin of Kikuchi lines involves both elastic and inelastic scattering, and a comprehensive
treatment of the problem requires extension of the dynamical diffraction theory to include
effects due to diffuse inelastic scattering process [4.18]. The contrast in the EBSD pattern
is due to the diffraction of the backscattered electrons by the crystal. Those backscattered
electrons are generated by an incident electron beam in the SEM. This produces a divergent
source of electrons within an interaction volume in the sample that diffract according to
the Bragg’s law. The diffraction results in intensity changes in the background. As there are
more electrons at + qB position (since distribution peaks in forward direction) then at – qB
position, one bright line is developed (the excess line) together with one dark line (the
deficit line) in the Kikuchi band.

incident electron

reflecting hkl plane


excess cone

-θB deficit cone


+θB
θ θ

hkl plane

projection on Ewald sphere

excess line deficient line excess line deficit line


hkl plane
a. b.
Fig. 4.30: Origin of Kikuchi lines

The diffracted electrons and the Kikuchi pattern can be detected when they impinge on a
phosphor screen suitably placed close to the sample and generate visible lines (Fig. 4.31).
The Kikuchi patterns are essentially a projection of the geometry of the lattice planes that
satisfy Bragg diffraction in the particularly oriented grain, and give direct information about
the crystal structure and crystallographic orientation of the grain from which they originate.

89
ELECTRON MICROSCOPY Instrumentation

Fig. 4.31: Typical Kikuchi pattern recorded in EBSD


(from hexagonal cobalt phase).

We will turn your CV into


an opportunity of a lifetime

Do you like cars? Would you like to be a part of a successful brand? Send us your CV on
We will appreciate and reward both your enthusiasm and talent. www.employerforlife.com
Send us your CV. You will be surprised where it can take you.

90
ELECTRON MICROSCOPY Instrumentation

The Kikuchi pattern generated on the phosphor screen is detected (can also be recorded)
by viewing the screen with a CCD video camera (Fig. 4.32). In practice, a polished sample
is placed in the SEM, with its surface inclined, approximately 70° relative to incidence
electron beam. The detector is a high-sensitive camera equipped with a phosphor screen
integrated with a digital frame grabber. The camera along with the phosphor screen can be
inserted (usually through a motorized mechanism) to within several mm of the sample when
measurements are done. The pattern of Kikuchi lines on the phosphor screen is electronically
digitized and processed to recognize the individual Kikuchi line positions in the pattern.
These data can be analyzed to identify phases, to index the pattern, and to determine the
orientation of the crystal.

Fig. 4.32: EBSD detector schematic.

The electron beam is at grazing angle to the specimen surface and EBSD need a double-
scattering event involving the BSE electrons. First, a quasi-elastic electron scattering with
the back scattered electrons forming a coherent and divergent electron source inside the
sample. After that, the electrons (BSE) must re-scatter at the atomic planes and come out
of the sample. To ensure a good signal-to-noise ratio (SNR) the virtual electron source must
have large interaction volume and should be very close to the sample surface. Both these
conditions need a high tilt angle of the sample with respect to the incident electron beam.
A high tilt angle increases the path length of the incident electrons in the near-surface
region. This results in more scattering events than would be the case for normal incidence.
Since electrons originating at shallow depths only make the most effective contributions to
the EBSD patterns, a high tilt angle improves the SNR. The pattern intensity is optimal
around 70° tilt.

91
ELECTRON MICROSCOPY Instrumentation

4.4 REFERENCES
4.1. R.H. Geiss and A.D. Romig, Jr., Introductory Electron Optics, Chapter 2, pages 29–74:
Principles of analytical electron microscopy, D.C. Joy, A.D. Romig, Jr., J.I. Goldstein
(editors). Plenum Press, New York (1986) pages 425. ISBN 978-1-4899-2037-9

4.2. J.C. Maxwell, On reciprocal figures and diagrams of forces, Philosophical Magazine
Series 4, 27 (1864) 250–261.

4.3. J.A. Fleming, Magnets and electric currents: An elementary treatise for the use of electrical
artisans and science teachers, E. & F.N. Spon, London (1902) pages 484.

4.4. F. Zemlin, K. Weiss, P. Schiske, W. Kunath, K.H. Herrmann, Coma-free alignment
of high-resolution microscopes with the aid of optical diffractograms, Ultramicroscopy
3 (1978) 49–60 https://doi.org/10.1016/S0304-3991(78)80006-0.

4.5. O. Scherzer, Über einige Fehler von Elektronenlinsen, Zeitschrift für Physik Volume
101, Numbers 9–10, Pages 593–603 (1936) as cited in P. Hawkes – Aberration
correction past and present, Phillisophical Transaction of Royal Society A 367 (2009)
3637–3664. doi:10.1098/rsta.2009.0004

4.6. H. Rose, D. Krahl, Electron optics of imaging energy filters, Chapter 3, pages 43–149:
Energy-Ffiltering transmission electron microscopy, L. Reimer (editor), Springer, Berlin
(1995) pages 419. ISBN 978-3-662-14055-0

4.7. C. Kisielowski, B. Freitag, M. Bischoff, H. van Lin, S. Lazar, G. Knippels, P. Tiemeijer,
M. van der Stam, S. von Harrach, M. Stekelenburg, M. Haider, S. Uhlemann, H.
Müller, P. Hartel, B. Kabius, D. Miller, I. Petrov, E.A. Olson, T. Donchev, E.A.
Kenik,4 A.R. Lupini,4 J. Bentley, S.J. Pennycook, I.M. Anderson, A.M. Minor, A.K.
Schmid, T. Duden, V. Radmilovic, Q.M. Ramasse, M. Watanabe, R. Erni, E.A. Stach,
P. Denes, U. Dahmen, Detection of single atoms and buried defects in three dimensions
by aberration-corrected electron microscope with 0.5-Å information limit, Microscopy
and Microanalysis, 14 (2008) 469–477. doi:10.1017/S1431927608080902

4.8. J.M. Zuo, J.C.H. Spence, Advanced transmission electron microscopy: imaging and
diffraction in nanoscience, Springer, New York, (2017) pages 717. ISBN: 978-1-4939-
6605-9

92
ELECTRON MICROSCOPY Instrumentation

4.9. G. McMullan, S. Chen, R. Henderson, A.R. Faruqi, Detective quantum efficiency of
electron area detectors in electron microscopy, Ultramicroscopy 109 (2009) 1126–1143.
doi:10.1016/j.ultramic.2009.04.002

4.10. T.E. Everhart, and R.F.M. Thornley, Wide-band detector for micro-microampere low-
energy electron currents, Journal of Scientific Instruments 37 (1960) 246–248.

4.11. G.I. Goldstein, D.E. Newbury, P. Echlin, D.C. Joy, C. Fiori, E. Lifshin, Scanning
electron microscopy and x-ray microanalysis, Springer, New York, (1981) pages 673.
ISBN 978-1-4613-3275-6.

4.12. B.J. Griffin, A comparison of conventional Everhart-Thornley style and In-Lens secondary
electron detectors, – a further variable in scanning electron microscopy, Scanning, 33
(2011) 162–173. doi: 10.1002/sca.20255

4.13. D.B. Williams and J.W. Edington, High resolution microanalysis in material science
using electron energy loss measurements, J. Microscopy, 108 (1976) 113–145. doi:
10.1111/j.1365-2818.1976.tb01086.x

93
ELECTRON MICROSCOPY Instrumentation

4.14. D.A. Muller, Why changes in bond lengths and cohesion lead to core-level shifts in
metals, and consequences for the spatial difference method, Ultramicroscopy, 78 (1999)
163–74. doi: http://dx.doi.org/10.1016/S0304-3991(99)00029-7

4.15. S.J.B. Reed, Electron microprobe analysis (2nd Ed.), Cambridge University Press,
(1993) pages 326. ISBN 0-521-41956-5

4.16. D.A. Gedcke, The Si(Li) X-Ray Energy Analysis System: Operating Principles and
Performance, X-Ray Spectrometry 1 (1972) 129–141. doi: 10.1002/xrs.1300010403

4.17. D.B. Williams, C.B. Carter, Transmission Electron Microscopy, Springer, New York,
(2009) pages 775. ISBN: 978-0-387-76502-0

4.18. 
P.B. Hirsch, A. Howie, R.B. Nicholson, D.W. Pashley, M.J. Whelan, Electron
Microscopy of Thin Crystals, Butterworth, London, (1965) pages 549.

94
ELECTRON MICROSCOPY Microscope Types and Operating Modes

5 MICROSCOPE TYPES AND


OPERATING MODES
In this chapter you will learn about

-- Differences and similarities in TEM and SEM


-- Components of TEM
-- Components of SEM
-- Operating modes in TEM and SEM
-- Imaging modes (e.g. dark field/ bright field images, SE/BSE images)
-- Diffraction modes (e.g. SAED, CBED, EBSD)
-- Analytical modes (e.g. EDS, WDS, EELS, HAADF)

The electron microscopes are mainly of two types: 1. the transmission electron microscope
(as the name implies) in which the electrons are transmitted through a thin specimen, and
2. the scanning electron microscope, where a bulk specimen scatters electrons from the
surface. The ray paths of these two types of microscope are compared with the light optical
microscope in Fig. 4.5, which shows significant similarities in them, but also large differences.

Both SEM and TEM refer to the instrument (electron microscope) and the method (electron
microscopy). They both are a type of electron microscope where electrons generated in an
electron source interact with a specimen and the scattered electrons are used to form an
image. Fig. 4.5 shows that the electron source, the condenser lens system to collimate the
beam and the objective lens to form an image have considerable similarities in both types of
microscope, although there are also many differences in these subcomponents. The type of
specimen used (thin / bulk) are very different and the way image is formed and magnified
fundamentally differ in the two microscopes. Images produced in these instruments are
highly magnified and have high resolution, but TEM is superior to SEM in both respects.
In TEM the image can be magnified typically up to 500 thousand times, while in SEM
only about 100 thousand times. The resolution of TEM is 0.5 angstroms while SEM has
0.4 nanometers. However, SEM images have a better depth of field compared to TEM
produced images.

In this chapter, a description of the two microscopes will be given and advantages/disadvantages
in imaging and analysis will be discussed.

95
ELECTRON MICROSCOPY Microscope Types and Operating Modes

5.1 TRANSMISSION ELECTRON MICROSCOPE


The transmission electron microscope is the first of the electron microscopes to be developed.
It was built following the principles of light transmission microscope, except replacing the
light source with a beam of electrons to “see through” a thin specimen. In the early twentieth
century, a concept developed that electron beams with wavelengths 1000 times shorter than
those of light could provide a better-resolved image of an object than a microscope using
light; because magnification is limited by the wavelengths (see Chapter 1). In 1928–29,
Ernst Ruska [5.1] demonstrated that a magnetic coil could act as an electron lens and in
1931 he and Max Knoll successfully built the first electron microscope. Ruska was awarded
the Nobel Prize for Physics in 1986 for this invention.

AXA Global
Graduate Program
Find out more and apply

96
ELECTRON MICROSCOPY Microscope Types and Operating Modes

Components of TEM
A transmission electron microscope is built up of three essential systems. The first of the
component is an electron source – consisting of an electron gun that produces the electron
beam and the condenser lens system, which focuses the beam onto the specimen. In TEM,
the electron source is designed to provide collimated stream of high-energy (> 80 keV)
electrons. The different types of electron gun and the condenser lens systems are described
in chapter 4 and their advantages/disadvantages discussed. The intensity and angular spread
of the beam are controlled by the condenser lens between the gun and the specimen. For
this purpose, the condenser system contains a small physical aperture (size selectable) in the
form of a disc of metal such as platinum with a precisely circular hole at its center. The
aperture size, together with the precise level of focus of the condenser lenses control the
brightness in the beam and that in the final image. Depending on the mode of operation
of the TEM, a parallel beam or a high intensity focused beam is needed on the specimen.
Whatever the design of the gun and the condenser system, the electron beam must be capable
of adjustment both in its position and in inclination. Such adjustments are provided by
suitable sets of magnetic alignment coils, which usually include corrections for astigmatism.

The image-producing system consists of the objective lens, a specimen stage and intermediate
plus projector lenses. The specimen to be examined is mounted on a special holder. Since the
microscope column is evacuated to high vacuum, the specimen holder is inserted through an
airlock to a position just above the objective lens. The specimen goniometer stage contains
micro-drives to provide both x-y movements and tilt of the specimen with respect of the
microscope axis and the illuminating electron beam. In addition, the specimen is usually
cooled with liquid nitrogen to avoid contamination and drift. The objective lens focuses
the electrons passing through the specimen to form a real magnified image. Changing the
current flowing through the objective lens changes its focal length and this constitutes the
focus control in the microscope. The magnified image from the objective lens is further
enlarged by the projector lenses placed below the objective lens in the column. Both the
absolute and relative excitations of these lenses are controlled electronically by a simple
magnification control on the microscope’s operating console. Moveable sets of apertures
are provided in three locations in the column: (a) in the condenser system to collimate
the electron beam and modify the intensity, (b) in the objective back focal plane to select
the appropriate beam for imaging, and (c) in the objective plane to select specific area of
specimen for selected area electron diffraction (SAED). An image of a typical TEM with
its tall column is shown in Fig. 5.1.

97
ELECTRON MICROSCOPY Microscope Types and Operating Modes

Fig. 5.1: Typical transmission electron microscope. Some important components


on the microscope column that are visible in the image are labeled.

The interior of the microscope is kept under vacuum, by a set of suitable pumps. The
pumping system is so designed that differential vacuum level can be maintained in the
microscope column and the viewing chamber. The vacuum is highest in the gun area and
gets lower down the column up to the viewing chamber. Images can be recorded on photo
films or on CCD cameras and other devices as discussed in chapter 4.

Image formation and magnification in TEM


The objective lens (which is usually immersion type with very short focal length) focuses
the beams transmitted through the specimen. The diffraction pattern is formed initially in
the back focal plane of the objective, and a magnified image in its image plane (Fig. 5.2).
The projector lenses transfer either the diffraction pattern or the image on to the viewing
phosphor screen placed at the image plane.

98
ELECTRON MICROSCOPY Microscope Types and Operating Modes

When an image is formed of the beams scattered by the specimen, two main mechanisms of
contrast arise. If the transmitted and the diffracted (scattered) beams are made to recombine,
thus preserving their amplitude and phases, a lattice image of the planes that are diffracting
or even the structure images of the individual atoms may be resolved directly (phase contrast).
The principle is the same as that of the Abbé theory [5.2] for gratings in light optics. On
the other hand, amplitude contrast in the image is obtained by deliberately excluding the
diffracted beams (and hence the phase relationship). From Fig. 5.2, it can be seen, that
this can be done by shifting the objective aperture suitably to allow only the transmitted
beam (marked 0) to pass through. The image formed in this way is called a bright field (BF)
image. Alternatively, only a single diffracted beam (e.g. the one marked g) can be selected by
the objective aperture to form the image, excluding all other beams, then a dark field (DF)
image is created. The bright field and dark field images have complimentary information.

�e Graduate Programme
I joined MITAS because for Engineers and Geoscientists
I wanted real responsibili� www.discovermitas.com
Maersk.com/Mitas �e G
I joined MITAS because for Engine
I wanted real responsibili� Ma

Month 16
I was a construction Mo
supervisor ina const
I was
the North Sea super
advising and the No
Real work he
helping foremen advis
International
al opportunities
Internationa
�ree wo
work
or placements ssolve problems
Real work he
helping fo
International
Internationaal opportunities
�ree wo
work
or placements ssolve pr

99
ELECTRON MICROSCOPY Microscope Types and Operating Modes

The magnification of the intermediate image formed by the objective lens at the back
focal plane is straightforward and can be varied by adjusting the projector lenses. A single
projector lens can produce a range of magnification, typically up to 5 times and by using
interchangeable pole-pieces in the projector, a wider range of magnifications can be obtained.
Modern TEM employ two projector lenses (one called the intermediate and the other
projector lens) to permit a greater range of magnification. By this way, a greater overall
magnification is achieved in the TEM without a commensurate increase in the physical
length of the microscope column. The quality of the final image in TEM depends largely
upon the accuracy of the mechanical and electrical adjustments with which the various lenses
are aligned to one another and to the illuminating system and the various lens aberrations
that was discussed in chapter 4.

optic axis
tilted illumination

specimen


objective lens

-g 2g
0 g back focal plane
objective apperture

image plane

Fig. 5.2: Ray diagram, showing image formation in TEM when a


tilted parallel electron beam illuminates the specimen

5.2 SCANNING ELECTRON MICROSCOPE


Only thin specimens that transmit electrons through them can be used in a TEM. For
viewing in microscopes operating at 200 to 300 keV acceleration voltages the thickness
of metallic specimens is typically limited to a few hundred nanometer. To observe thicker
(bulk) specimens in the electron microscope, scanning electron microscopes are needed.

100
ELECTRON MICROSCOPY Microscope Types and Operating Modes

Although, Knoll (1935) can be attributed the credit for the first ‘‘scanning microscope’’, von
Ardenne (in 1938) clearly established the theoretical principles underlying the scanning
electron microscope, as we know it today [5.3]. The first true SEM was developed in 1942
by Zworykin, who showed that secondary electrons provide topographic contrast when the
collector (detector) is positively biased relative to the specimen [5.3]. In 1960, Everhart
and Thornley greatly improved the secondary electron detection with the development of
the ETD detectors (see Chapter 4).

Components of SEM
A scanning electron microscope produce image by probing the specimen with a focused
electron beam that is scanned (raster scanning) across an area of the specimen surface. When
the incident electron beam interacts with the bulk specimen, it loses energy by various
mechanisms, producing a variety of signals that contain information about the sample’s
surface topography and composition (see Chapter 4). Any of these signals (e.g. SE or BSE)
can be mapped to form an image, which correspond position-by-position the beam on the
specimen, when the signal was generated.

Like in TEM, the scanning electron microscope is also made up of three main systems:
the electron source including beam collimation, the image-producing system containing the
objective lens and the detector systems. In SEM, the electron source is designed for a lower
acceleration voltage (up to 40 keV) than in TEM and therefore the microscope column is
much more compact (Fig. 5.3).

Fig. 5.3: Typical scanning electron microscope.


Some components of the microscope are labeled.

101
ELECTRON MICROSCOPY Microscope Types and Operating Modes

The microscope column houses the gun at the top, followed by the condenser and objective
lenses, which control the beam (probe) size and focus it on the specimen. The beam travels
down the column to the specimen chamber at the base. A series of apertures, which the beam
passes through regulate the properties of the beam. The condenser lens controls the intensity
of the electron beam reaching the specimen and the objective lens brings the electron beam
into focus (de-magnifies) on the specimen. The objective lens aperture is used to reduce or
exclude extraneous (scattered) electrons. An optimal aperture diameter should be selected
for obtaining high-resolution electron images. A pair of scanning coils is placed below the
condenser lens system in the column, associated with the objective lens area. The beam
is deflected in the x and y-axes by the scanning coils, scanning the specimen surface in a
raster scan pattern, with consecutive dwell-times of the beam spot across a rectangular area
(Fig. 5.4). From the area of beam-specimen-interaction several types of signals (including
SE and BSE) are generated that can be detected by suitably placed detectors inside the
specimen chamber and processed to produce an image. The specimen (bulk material) is
mounted on a specimen holder fixed to a specimen stage inside the specimen chamber.
The stage can be moved along the X, Y (in the specimen plane), and Z directions (at right
angles out of the specimen plane). The Z adjustment is also known as the specimen height.
The specimen stage can also rotate continuously. The chamber is maintained in vacuum to
minimize scattering of the electron beam before they reach the specimen. This is important
as scattering or attenuation of the electron beam, if happens (e.g. in variable pressure SEM),
will need an increase in the probe size that will reduce the image resolution, especially in
the SE imaging mode. A high vacuum condition also optimizes collection efficiency at the
detectors, particularly of the secondary electrons. Like in the TEM, a differential vacuum
level is maintained in the SEM with the value in the upper column being greater than in
the specimen chamber.

Image formation and magnification in SEM


The image formation and magnification in SEM is fundamentally different from that in a
TEM. Like in an optical microscope, the image of the specimen in a TEM is formed by
the objective lens using the electrons scattered through the specimen and magnified by the
projector lenses (Fig. 4.5). It is rather unusual the way in which image is formed in SEM,
which seems to differ greatly from normal human experience with image formation by
light as viewed by the eye. In order to produce images the electron beam is focused into a
fine probe, which is scanned across the surface of the specimen with the help of scanning
coils (Fig. 5.4). Each point on the specimen that is struck by the accelerated electrons
emits signal in the form of electromagnetic radiation. Selected portions of this radiation,
usually secondary.

102
ELECTRON MICROSCOPY Microscope Types and Operating Modes

(SE) and/or backscattered electrons (BSE) are collected by a detector and the signal is
amplified and displayed on a monitor. The resulting image is generally straightforward to
interpret, at least for topographic imaging of objects at low magnifications as it has point-
to-point correspondence with the beam on the specimen.

The purpose of the objective lens in a SEM is to create a small, focused electron probe on
the specimen. Most SEMs can generate an electron beam at the specimen surface with spot
size less than a few nanometers in diameter while still carrying sufficient current to form
acceptable image. The SEM image is a 2D intensity map in the analog or digital domain.
Each image pixel on the monitor corresponds to a point on the sample, with proportional
signal intensity captured by the detector at each specific point (Fig. 5.4). Unlike, optical
or transmission electron microscopes no real image exists in the SEM. It is not possible to
place a film anywhere in the SEM and record an image. The image is generated/formed
and displayed by electronic synthesis.

93%
OF MIM STUDENTS ARE
WORKING IN THEIR SECTOR 3 MONTHS
FOLLOWING GRADUATION

MASTER IN MANAGEMENT
• STUDY IN THE CENTER OF MADRID AND TAKE ADVANTAGE OF THE UNIQUE OPPORTUNITIES
Length: 1O MONTHS
THAT THE CAPITAL OF SPAIN OFFERS
Av. Experience: 1 YEAR
• PROPEL YOUR EDUCATION BY EARNING A DOUBLE DEGREE THAT BEST SUITS YOUR
Language: ENGLISH / SPANISH
PROFESSIONAL GOALS
Format: FULL-TIME
• STUDY A SEMESTER ABROAD AND BECOME A GLOBAL CITIZEN WITH THE BEYOND BORDERS
Intakes: SEPT / FEB
EXPERIENCE

5 Specializations #10 WORLDWIDE 55 Nationalities


MASTER IN MANAGEMENT
Personalize your program FINANCIAL TIMES
in class

www.ie.edu/master-management mim.admissions@ie.edu Follow us on IE MIM Experience

103
ELECTRON MICROSCOPY Microscope Types and Operating Modes

LM
Magnification

LM
M=
LS

transfer of
information

LS

Area projected
on monitor
Area scanned on
specimen
Fig. 5.4: Image formation and magnification in SEM.

As no real or virtual optical images are produced in the SEM, thus, no optical transformation
is responsible for image magnification. Magnification is achieved simply because the
electron probe scanning an area on the specimen is smaller than the image pixel displayed
on the monitor. Since the monitor length is fixed, increase or decrease in magnification
is achieved by respectively reducing or increasing the length of the scan on the specimen.
The magnification of the SEM image is changed by adjusting the length of the scan on the
specimen (LS) for a constant length of scan on the monitor (LM). The linear magnification
(M) of the image is therefore, given by the ratio of the length on the monitor versus the
length of the scan (Fig. 5.4). For accurate magnification measurement, a calibration of the
length scale is necessary. Magnification in the SEM depends only on the excitation of the
scan coils and not on the excitation of the objective lens, which determines the focus of
the beam.

5.3 OPERATING MODES IN TEM AND SEM


Electron microscope is a versatile modern instrument and a valuable tool for material
characterization. Quite different information can be obtained from the material investigated
by TEM and SEM but they are complementary in nature. Both the microscopes can be
operated in several different modes to obtain image, diffraction and analytical information
from the specimen.

104
ELECTRON MICROSCOPY Microscope Types and Operating Modes

In a TEM, images can be obtained using the transmitted direct beam (Bright Field image)
or from one of the diffracted beams (Dark Field image) to get contrast from different
phases and defects (dislocation, stacking faults, etc.) in the specimen. Alternatively, a high-
resolution lattice resolved image (HREM) could be obtained from a crystalline material
to show the atomic arrangement in the specimen. The microscopes can be also operated
in the diffraction mode to obtain crystallographic information from the specimen. The
diffraction mode can be used on a powder sample to obtain ring patterns, familiar in x-ray
diffraction, or on a selected area of a solid specimen to obtain Laue spot patterns in a
SAED image. These modes of operation in TEM mostly use parallel incident beam on the
specimen. However, by using a converging beam on the specimen the so-called convergent
beam electron diffraction (CBED) images can be obtained in TEM. Each diffracted Laue
spot then becomes a disc within which variations in intensity can usually be seen. Such
CBED patterns contain a wealth of information about the symmetry of the crystal. One
big advantage of CBED over SAED techniques is that information is generated from a very
small region in the specimen, beyond that possible to analyze in SAED mode. The TEM
can be also used in a scanning mode, where a focused beam is scanned in a raster on the
specimen (like in SEM) to perform scanning transmission electron microscopy (STEM).
The STEM technique scans a very finely focused beam of electrons across the sample and
the interactions between the beam electrons and sample atoms generate a series of signal
stream. These signals are correlated with the beam position and can be used to build a virtual
image in which the signal level at any location in the sample is represented by the gray
level at the corresponding location in the image. Its primary advantage over conventional
SEM imaging is the improvement in spatial resolution.

One of the principal advantages of STEM over TEM is in enabling the use of other types of
signals that can be spatially correlated, including secondary electrons, scattered beam electrons,
characteristic x-rays, and electron energy loss. Therefore, the STEM/TEM can be used also
as an analytical microscope. X-rays generated by electrons bombarding the specimen can be
used for microanalysis using x-ray-energy dispersive spectrometer (EDS) in TEM to count and
sort characteristic x-rays according to their energy. In STEM, transmitted beam electrons that
have been scattered through a relatively large angle can be detected using a high-angle-annular
dark field (HAADF) detector, which gives contrast of the specimen according to distribution
of different atom types in the material (Z-contrast). Further, in the electron-energy-loss
spectrometry (EELS) analysis transmitted electrons in a STEM are used to determine the
amount of energy they have lost in interactions with the thin sample. It provides information
about the interacting atoms, including elemental identity, chemical bonding, valence and
conduction band electronic properties, surface properties, and element-specific pair distance
distribution functions. While the standard viewing screen or recording devices (e.g. phosphorus
screen, or CCD camera, etc.) can be used for most imaging and diffraction modes in TEM,
the analytical modes (EDS, HAADF and EELS) require special detectors.

105
ELECTRON MICROSCOPY Microscope Types and Operating Modes

Annular dark field imaging is a method of mapping samples in a scanning transmission


electron microscope. These images are formed by collecting scattered electrons with an
annular dark field detector. In traditional dark field imaging (in a parallel beam mode in
TEM), an objective aperture is placed in the diffraction plane so as to only collect electrons
scattering through that aperture, avoiding the main beam. By contrast, in STEM mode the
optics distinguishing between dark and bright field modes is positioned further downstream,
after the convergent beam has interacted with the specimen. Consequently, the contrast
mechanisms are different between conventional dark field imaging and STEM dark field. An
annular dark field detector collects electrons from an annulus around the beam, sampling
far more scattered electrons than can pass through an objective aperture. This gives an
advantage in terms of signal collection efficiency and allows the main beam to pass to an
EELS detector, allowing both types of measurement to be performed simultaneously. An
annular dark field image formed only by very high angle, incoherently scattered electrons (as
opposed to Bragg scattered electrons) is highly sensitive to variations in the atomic number
of atoms in the sample (Z-contrast). This technique is known as high angle annular dark
field imaging or HAADF [5.4].

106
ELECTRON MICROSCOPY Microscope Types and Operating Modes

Likewise, the SEM can also be used in several modes of operation, to provide different types of
image contrast or composition and crystallographic information of the investigated specimen.
However, unlike in TEM, since a variety of scattered signals is generated in SEM by the interaction
of the electron beam with the bulk specimen, the detector is mostly specific for the various modes
of imaging and analytical signals. This implies that in the different SEM mode, the microscope
operation (in term of scanning) mostly remains the same, but the detector is changed. Therefore,
for normal imaging with SE an Everhart and Thornley detector is used, but for a BSE image or
for a low voltage surface image the BSE detector or an in-lens detector is used.

X-ray microanalysis in SEM can be done either in energy dispersive spectroscopic (EDS)
mode or in wavelength dispersive spectroscopic (WDS) mode. WDS is particularly effective
for detection of light elements in the specimen. The scanning electron microscopes in general
is not operated in diffraction mode, however, special conditions exists when a primary
scattered incident electron generate a backscatter electron which is subsequently diffracted
by the crystalline specimen and give rise to a Kikuchi diffraction pattern (EBSD). Operating
the SEM in EBSD mode require a highly tilted (70° with respect to the beam) sample and
a high-sensitive camera with phosphorus screen to detect the diffraction pattern.

Moreover, by using a thin specimen (like in TEM), and a special setup with the detector
below the sample to allow transmission of electrons through the specimen, the SEM can
be operated in STEM mode. In STEM mode, the SEM operates like a low voltage TEM
and analytical methods like HAADF can be used. In STEM mode, it is also possible to
use dark field and bright field imaging in SEM.

In chapter 6, the different microscopy and analysis methods in TEM, SEM and STEM
that are mentioned here will be discussed in some more details. This book however only
introduces the topics and the methods to the students and for a complete knowledge and
understanding, further reading is recommended.

5.4 REFERENCES
5.1. P.W. Hawkes, Ernst Ruska, Physics Today 43 (1990) 84-85. doi: 10.1063/1.2810640

5.2. H. Köhler, On Abbe’s Theory of Image Formation in the Microscope, Optica Acta 28
(1981) 1691–1701. doi: 10.1080/713820514

5.3. A. Bogner, P.-H. Jouneau, G. Thollet, D. Basset, C. Gauthier, A history of scanning
electron microscopy developments: Towards ‘‘wet-STEM’’ imaging, Micron 38 (2007)
390–401. doi:10.1016/j.micron.2006.06.008

5.4. D.E. Jesson, S.J. Pennycook, Incoherent Imaging of Crystals Using Thermally Scattered
Electrons, Proc. Roy. Soc. A. 449 (1995) 273–293. doi:10.1098/rspa.1995.0044

107
ELECTRON MICROSCOPY Microscopy and Analysis

6 MICROSCOPY AND ANALYSIS


In this chapter you will learn about

-- Imaging and image contrast in SEM, TEM and STEM


-- SE and BSE images in SEM
-- BF and DF images in TEM and STEM
-- HREM images in TEM
-- HAADF images in STEM and SEM
-- Diffraction in TEM and SEM
-- SAED in TEM
-- CBED in TEM
-- EBSD in SEM
-- Analytical Electron Microscopy
-- EDS in TEM and SEM
-- WDS in SEM
-- EELS in TEM

Interaction of accelerated electrons with matter generates a variety of signal and in the
electron microscope; they are used for imaging or collecting diffraction and other analytical
information about the specimen. In the previous chapter, it was discussed that image
formation principle is fundamentally different in the case of SEM and TEM. While an
optical transformation by the objective lens, of electrons scattered at the specimen atoms,
is essential to form a real image in TEM, no such transformation is needed in SEM and
STEM, where a virtual image is formed by mapping an area of the sample point-by-point.
The contrast information available in TEM and SEM images are complex but are well
understood and described in many textbooks [e.g. 6.1–6.3].

108
ELECTRON MICROSCOPY Microscopy and Analysis

6.1 IMAGING
6.1.1 SCANNING ELECTRON MICROSCOPY

SE and BSE Images in SEM


SEM obtains high magnification images, with a good depth of field and high-resolution. The
two main types of scattered electrons generated in SEM by the interaction of the primary
beam with the bulk specimen are SE and BSE and both can be detected and mapped to form
image using appropriate detectors. For routine SEM imaging, SE replicates the image of the
specimen surface. Secondary electrons are low energy electrons (< 50eV) formed by inelastic
scattering and can be easily collected by Everhart-Thornley detector. Only those SE produced
near the specimen surface are able to escape the sample and can reach the ETD. However, a
major influence on SE signal-generation and collection is the shape of the specimen surface
(Fig. 4.18). As the ETD is placed on one side, the SE signals emitted from a surface that
faces away from the detector are blocked by the specimen surface topography. These surfaces
appear darker in the image than surfaces that face towards the detector. Secondary electrons
provide particularly good edge detail, which are brighter than the rest of the image because
a sharp edge scatters more electrons. This topographical contrast due to the position of the
SE detector is a major factor in building contrast in the SE images. However, it is not the
only factor, which contributes to the brightness and contrast in an SEM image.

In the past 5 years we have drilled around

95,000 km
—that’s more than twice around the world.

Who are we?


We are the world’s leading provider of reservoir characterization,
drilling, production, and processing technologies to the oil and
gas industry.

Who are we looking for?


We offer countless opportunities in the following domains:
n Operations
n Research, Engineering, and Manufacturing
n Geoscience and Petrotechnical
n Commercial and Business

We’re looking for high-energy, self-motivated graduates


with vision and integrity to join our team. What will you be?

careers.slb.com

109
ELECTRON MICROSCOPY Microscopy and Analysis

The primary function of the SE detector is to attract low energy secondary electrons,
which are generated from approximately the top layer of the specimen surface (depth of
10–15 nm). Generally, however, the image produced by the ETD will always contain an
amount of sub-surface information derived from high energy BSE. As a rule, the higher
the accelerating voltage of the incident electron beam, the more subsurface information is
picked up by the detector due to various backscattered effects.

BSE are high-energy electrons from the primary incident beam (having energies much
greater than 50 eV and close to that of the incident beam energy) that are backscattered
out of the bulk specimen. BSE produce a different kind of contrast in the SEM image than
the SE. The contrast depends on the type of atom in the investigated sample. Higher the
atomic number (heavier atom in the specimen), more primary electrons are backscattered,
which leads to a brighter image for materials constituted of higher atomic number atoms.
Moreover, the greater energy of BSE compared with SE means that the interaction volume
is larger and BSE produced from deeper within the interaction volume can escape from the
specimen and collected by the BSE detector. Thus, BSE images have lower spatial resolution
than SE images. While BSE generation is independent of the specimen topography, the
topography does interfere with the BSE signal reaching the detector, although not as
strongly as in the case of SE due to the detector position (Fig. 4.19). Therefore, to obtain
the best compositional information using BSE, it is preferable to use a flat sample. For a
smooth (e.g. polished) specimen the dense materials provide brighter signal level then less
dense material. The greyscale differences in the BSE image indicate the average Z (atomic
number) contrast of the phases present and allow the recognition and classification of
different phases in the material.

Fig. 6.1: SE and BSE Image pair obtained in SEM from a Co-based alloy with the Cr-Re-rich σ phase (bright):
a) SE image recorded with ETD, b) BSE image recorded with CBS detector.

110
ELECTRON MICROSCOPY Microscopy and Analysis

In the example shown Fig. 6.1 the contrast from SE and BSE images recorded from the
same area of a Co-based alloy specimen is compared. The influence of the topographic and
Z contrast is visible in the SE and the BSE images respectively, but each image is influenced
by both SE and BSE signals. As Re has high atomic number (# 75), the Re-rich Cr2Re3-type
s phase is bright. The topologically closed packed s phase particles are hard and they are
polished less than the Co matrix, therefore, they project out of the surface. The particles are
flat in the BSE image (Fig. 6.1b) but show topographic features in the SE image (Fig. 6.1a).

Contrast changes in SEM image also due to a change in acceleration voltage, spot size
(probe current) and incident beam tilt angle, each of which change the SE to BSE ratio.
The emission of secondary electrons remains relatively constant for a given probe current
and does not change much with the accelerating voltage of the incident beam. However,
the BSE influence in the image increases as the accelerating voltage is increased, diminishing
the contribution of SE. Higher acceleration voltage increase the interaction volume and
thereby decrease image resolution. Another area of influence is the design of the microscope,
particularly the detector position and its type. To obtain high-resolution SEM images in-lens
detectors (Fig. 4.20), low working distance and low acceleration voltages are useful. SEM
imaging is possible with only a few hundred volts in modern microscopes.

Imaging with In-lens Detectors in SEM


Due to the interaction of the primary beam with the specimen, several types of electrons
are generated in the SEM and all of these electrons carry distinct structural information
about the sample and differ from one another in their origin, energy and traveling direction.
The type I secondary electrons (SE1) emitted from close proximity of the impact point (of
the incident beam) at a high angle, carry high-resolution surface-sensitive (topographic)
information of the sample. On the other hand, type II secondary electrons (SE2), which
are commonly used in SE imaging with the ETD, are generated from a deeper and wider
volume than the SE1 and reflect at a lower angle, therefore carrying intrinsically lower-
resolution topographical information.

Since the early 1980s, SEM fitted with in-lens detector are available, which can collect the
SE1 to produce high-resolution images. The standard Everhart-Thornley detector is mounted
in the SEM specimen chamber but the in-lens detector is located inside the objective lens
(Fig. 4.20). In this configuration, the objective lens acts as an electron filter, where the lens
field and lens geometry prevents the BSE and SE2 from reaching the in-lens detector. In
conventional SEM, high-resolution imaging of the specimens with SE had been prevented
because by moving the specimen nearer to the objective lens most of the SE signals are
blocked, and cannot reach the ETD. With a detector in the objective lens positioned just
above the sample, such problems are overcome.

111
ELECTRON MICROSCOPY Microscopy and Analysis

The low energy SE1 is unable to break away from the lens axis due to the strong field of
the objective lens. These electrons, which cannot move either downwards or to the side,
spiral back up the column until they are outside of the objective lens field and at this
point, they are attracted into the in-lens SE detector. In this lens design, a pure secondary
electron image may be formed and the high-energy BSE does not contribute to the image.
However, some BSE that strikes the objective lens inner surface are converted to produce
SE. They may be incorporated in the final image or excluded, depending upon the lens
configuration. With the in-lens-detector, images may be obtained from specimens at less than
3mm from the final lens. The constructive influence of BSE to an image however, should
not be discounted, very often whilst SE offer resolution, the BSE offers more information.

Recently, in-lens BSE detectors are also available. Like the SE1, singly scattered backscattered
electrons (BSE1) tend to emit at a high angle and are closely related to compositional contrast,
while multiply scattered BSE (BSE2) take off at a lower angle and are conventionally used
to characterize composition and crystalline structures of the sample. Comparison of Standard
SE, BSE and in-lens SE and in-lens BSE images from the same area is shown in Fig. 6.2.

Excellent Economics and Business programmes at:

“The perfect start


of a successful,
international career.”

CLICK HERE
to discover why both socially
and academically the University
of Groningen is one of the best
places for a student to be
www.rug.nl/feb/education

112
ELECTRON MICROSCOPY Microscopy and Analysis

a. b.

c. d.
Fig. 6.2: Standard SE and BSE Images (with ETD and CBS detectors) compared with in-lens SE and BSE images.
All images obtained in SEM from a Co-based alloy. a) SE image, b) BSE image, c) In-lens SE image, b) in-lens
BSE image.

Electron Channeling Contrast in SEM


In Fig. 6.1, in addition to topographic and Z contrast, an additional contrast is visible.
The differently oriented cobalt solid-solution matrix grains / sub-grains in the images show
different grey level contrast. This contrast arises due to electron channeling mechanism and
this special imaging method is known as electron channeling contrast imaging (ECCI) [6.4].
ECCI is a SEM technique that makes use of the strong dependence of the BSE signal on the
orientation of the crystal lattice planes with respect to the incident-electron-beam direction.
ECCI uses the backscatter electron detector and can be effective for direct observation of
lattice defects in SEM (for example dislocations or stacking faults). In this technique, only
defects close to the surface of the bulk specimen can be detected.

113
ELECTRON MICROSCOPY Microscopy and Analysis

When high-energy primary beam electrons in a SEM enter a crystalline specimen, they
form a standing electron-density wave inside the lattice (the so-called primary wave field),
which is coherent with the crystal lattice. Depending on the direction of the primary
electron-beam with respect to the orientation of the lattice, the maxima of the electron-
density waves may change from a position at the atomic nuclei to one in between them.
In the first case, strong backscattering of electrons out of the primary wave field occurs,
while in the latter case only few electrons are backscattered. Consequently, the backscatter
signal carries information about the crystal lattice and its orientation relative to the primary
beam. In Fig. 6.3, the channeling concept is presented in a simple way by considering the
electrons as particles, rather than wave. To understand, channeling a crystal lattice can
be described in terms of channels or paths rather than a set of atomic points, where the
electron (particle) can preferably penetrate to a higher depth before scattering. Hence, for
some orientations of the crystal, electrons will backscatter more than other orientations
(Fig. 6.3a and 6.3b). This will give rise to orientation contrast. Minimum backscattering
occurs when the primary beam almost exactly fulfills the Bragg angle with one of the lattice
planes. The electrons then travel deep into the crystal without intense interaction with it.
This case is called electron channeling.


e¯ e¯

e¯ e¯


e¯ e¯ e¯

a. b.
Fig. 6.3: Illustration of atomic lattice relative to incoming electron trajectories showing electron
channeling (a) more BSE are generated when the channel is closed or (b) in open channel
condition electrons penetrate the lattice to a larger depth before scattering.

114
ELECTRON MICROSCOPY Microscopy and Analysis

If a defect (e.g. a dislocation) is present in the crystal, then the coherency of the channeling
primary electron wave field with the lattice is disturbed and strong backscattering occurs
at the position of the defect (Fig. 6.4). As a result, the defect is visible as a bright feature
on an otherwise dark background when the image is observed with a backscatter electron
detector. According to the dynamical theory of electron diffraction, the optimum diffraction
contrast for imaging crystal defects in ECCI is obtained by orienting the crystal exactly into
Bragg condition for a selected set of intense diffracting lattice planes [6.5]. ECCI shows
very comparable contrast features to dark-field TEM with the advantage that the images
are obtained on bulk samples rather than on thin foils. The contrast and resolution of the
ECCI images are less than in TEM images.

American online
LIGS University
is currently enrolling in the
Interactive Online BBA, MBA, MSc,
DBA and PhD programs:

▶▶ enroll by September 30th, 2014 and


▶▶ save up to 16% on the tuition!
▶▶ pay in 10 installments / 2 years
▶▶ Interactive Online education
▶▶ visit www.ligsuniversity.com to
find out more!

Note: LIGS University is not accredited by any


nationally recognized accrediting agency listed
by the US Secretary of Education.
More info here.

115
ELECTRON MICROSCOPY Microscopy and Analysis

e¯ e¯

e¯ e¯

Fig. 6.4: ECCI imaging of defect

For the ECCI technique, the SEM should be operated at high magnifications and with
low working distance. Thus, the beam tilting above the imaged specimen region is limited.
Visible contrast is caused by differences in orientation of individual grains (Fig. 6.1b) or
by local tilting of the lattice planes in presence of lattice defects (Fig. 6.4).

Depth of Field in SEM Images


One of the great advantages of SEM image is its large depth of field (which is sometimes
confused with depth of focus). The field refers to space defined on the specimen, whereas the
focus refers to space in the image plane. Since in the SEM, images are not produced by a
lens system (as in LOM and TEM where the objective lens produces the image), there can
only be a depth of field. Confusion arises in microscopy between the terms depth of field and
depth of focus because no optical transformation (Abbé theory) is involved in SEM image
formation (see Chapter 5).

116
ELECTRON MICROSCOPY Microscopy and Analysis

Large depth of field makes it possible in SEM to examine very rough surfaces with deep
topographic features at much higher magnifications than is possible in LOM and TEM. This
makes SEM the preferred choice for fracture surface analysis in material characterization.
The large depth of field is demonstrated in Fig. 6.5, where a SEM image of cuboidal and
spherical g’ nanoparticles (selectively extracted from a Ni-base superalloy [6.6]), shows
that the small spherical g’ particles deep inside the channels between the large cuboidal g’
particles are clearly in focus.

Fig. 6.5: SEM image of γ’ precipitates extracted from Ni-base superalloy shows large
depth of field in SEM images.

The final lens of the SEM focuses the electron beam to a crossover at the plane of best focus
(Fig. 6.6a). The beam diameter increases as the beam converges or diverges above and below
this plane. At some distance, D/2 above and below the focus plane the diameter of the
beam becomes twice the pixel diameter for a given magnification, whereupon the signals
from adjacent pixels overlap and blur the image. Over the distance between these limits,
however, the image remains in sharp focus. The distance D is called the depth of field. The
depth of field can usually be adjusted by changing the semi-angle of the illuminating beam
and in a conventional SEM can be controlled by using the two independent variables – the
final aperture size and the working distance (Fig. 6.6b).

117
ELECTRON MICROSCOPY Microscopy and Analysis

It is important to note that the effective signal emission diameter also varies according to
the type of signal being collected and therefore, an important parameter here is the energy
of the scattered electrons. For example, in the case of SE (low energy) signal, the emission
diameter is equivalent to the effective probe diameter but if the BSE signal (high energy)
is used to form the image, then, the emission diameter is somewhat larger (due to the
larger interaction volume). Adjusting the final aperture size (to control a) and the working
distance (W) will have separate effect on the effective probe diameter. Generally, the depth
of field (D) is higher when:

• the emission disc is larger


• the final aperture is smaller
• the working distance is longer
• the SEM magnification is low

118
ELECTRON MICROSCOPY Microscopy and Analysis

electron beam electron beam electron beam


e¯ e¯ e¯
Final aperture Final aperture Final aperture
α α2
α1
W1

W Best focus D W2

Bulk specimen
D/2
D D
Best focus Best focus

Bulk specimen Bulk specimen

Fig. 6.6: Schematic diagram explaining the depth of field in SEM. a) The distance D is called the depth of field.
b) D can be adjusted by changing the final aperture size (which changes α) and the working distance (W). With
longer working distance (W2 > W1) and smaller aperture (α2 < α1) the depth of field (D) is increased.

6.1.2 TRANSMISSION AND SCANNING TRANSMISSION ELECTRON MICROSCOPY

BF and DF Images in TEM and STEM


The ability to generate high-resolution and high-magnification images from a thin specimen
with the TEM was a major breakthrough in electron optical technology in the early 20th
century. TEM imaging provides information of microstructural features from nanometer
to micron size that include defects in crystalline lattices. These microstructural features and
defects control many important properties of materials and TEM often provide detailed
and more direct information about them, than can be obtained by any other experimental
technique. Generally, the contrast due to the variation of thickness of the specimen is
weak, except in cases where the specimen contains atoms with large differences in atomic
number or when diffraction is weak. The phase contrast and Z-contrast relating to HREM
and HAADF imaging will be discussed later. Here we discuss condition where mass thickness
contrast is generally weak and overshadowed by the stronger effects of electron diffraction.
Contrast in BF and DF TEM images, which reveal the microstructural features and defect
in the thin specimen is usually from diffraction contrast, where the variations in intensity
in the image is due to diffraction across the sample.

The development of STEM has not only added major analytical capabilities in TEM, but
also provides imaging possibilities, which is particularly useful for sensitive samples (e.g.
polymers and biological materials) which are prone to electron beam damage. STEM mode
can also generate dark and bright field images from the thin specimen however; the contrast
mechanism is somewhat different from conventional BF/DF imaging in TEM.

119
ELECTRON MICROSCOPY Microscopy and Analysis

In order to interpret electron images (and diffraction patterns) in TEM, it is essential to


understand the factors that determine the intensities of the Bragg diffracted beam. The
amplitude of the wavevector scattered by an atom in a crystal lattice can be calculated
under the so-called, kinematical approximation, which assume that the amplitude of the
scattered wave is only a very small fraction of that of the incident wave. The kinematical
theory is applied here to explain BF and DF image contrast and how contrast from defects
like dislocations arises in TEM images. The shortcomings of the Kinematical approximation
will be discussed latter with respect to contrast of HREM images and it will be shown how
they can be overcome using dynamical approximations. In order to interpret BD and DF
image contrast it is important to understand some additional concepts, e.g. Laue diffraction
conditions, structure factor, reciprocal lattice and Ewald sphere.

Laue Diffraction Conditions


The atomic scattering factor (fθ), i.e. the amplitude of coherent electron wave scattered by
an isolated atom can be expressed in terms of x-ray scattering and given by:

where Z is the atomic number of the isolated atom and fx is the atomic scattering factor for
x-rays [6.2]. The equation shows that fθ, which is a measure of the scattering power of an
isolated atom, depends on the ratio sinq/l where q is the Bragg angle and l the wavelength
of the electron wave.

When scattering of electron occurs at many points in the crystal lattice, the Laue equations
give the condition for constructive interference of the waves scattered from corresponding
lattice points (which can be essentially reduced to the Bragg’s law) [6.7]. Laue formulated
an alternative theorem to the Bragg’s law for diffraction. This theorem is useful for electron
diffraction because it does not require the assumptions used by Bragg, that diffraction
involves parallel planes of atoms, but assumes that elastic scattering occur at lattice points.
In crystallography, the Laue conditions must be satisfied for constructive interference to
occur, when incident waves are diffracted by the crystal lattice atoms. Diffraction contrast
and the appearance of features in BF and DF images depend sensitively on how the Laue
diffraction conditions are satisfied.

120
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.7 presents the vector relationship of electron waves scattered by two atoms in a crystal
lattice. The distance from atom A to atom B is described by the vector r in three-dimensional
space. The wavevector for the incoming (incident) beam is ki and for the outgoing (scattered)
beam is ko. The path-difference between the two waves is AC – AD, which in vector notation
is r.( ki – ko), where ( ki – ko) = D k is the scattering vector (also called transfer wavevector).
The wavevectors ki and ko are unit vectors with modulus 1/l, where l is the wavelength of
the incident and scattered beams. The path difference of the waves should be an integer
multiple of the wavelength for constructive interference (diffraction) to occur, i.e. CAD
= r. D k = nl. This is equivalent to the Bragg’s law. In Fig. 6.7a, the vector r is defined
in spherical coordinates. In Laue equations [6.7] the vector r is expressed in Cartesian
coordinates (x, y and z) and the distance between atoms in terms of unit vectors (a, b and
c) defined in the crystal lattice coordinate axes. It turns out that the difference between
the two wavevectors D k equals g the reciprocal lattice vector of the crystal (with magnitude
d*hkl) when the Bragg diffraction condition is satisfied (Fig. 6.7b).

121
ELECTRON MICROSCOPY Microscopy and Analysis

incident electon beam


incident
electron beam
Ewald Sphere
θ

lattice planes

s=1
radiu

(hkl)
C

Thin specimen r
A B

D
ki - ko= ∆k = g ki ko

ki ko
o g
d*hkl

a. ∆k b. Reciprocal Lattice

Fig. 6.7: a) Electron waves scattered by two atoms in a thin crystal. The incoming wavevector is ki and
the outgoing (scattered) beam is ko. b) Ewald sphere construction and reciprocal lattice satisfying the
Laue condition.

Structure Factor
The Laue equation gives the condition only for the reinforcement of waves scattered by
corresponding points in the crystal lattice. However, there is interference between waves
scattered from different points in each unit cell, such that in some cases there are orders of
reflections given by Bragg and Laue equations which are completely or partially suppressed
by this effect. The diffraction intensity is different for each crystalline plane because the
distribution of atoms per unit area for individual planes is not the same and depends on the
type of crystalline lattice and the atomic arrangements. The structure factor is a mathematical
description of how a material scatters incident radiation and gives the amplitude and phase
of electron beam diffracted by the crystallographic plane. The structure factor takes into
account the location of atoms in the scattering plane and the type of atom in the crystal
structure (i.e. their scattering power) to describe the diffraction intensity. The structure factor
(Fhkl ) can be calculated knowing the individual atomic scattering factors (fq) and is given by:

where n is the total number of atoms and (hkl) is the Miller index for the scattering plane
in the crystal lattice.

122
ELECTRON MICROSCOPY Microscopy and Analysis

Reciprocal Lattice
Atoms in crystalline materials are arranged in regular array in three-dimensional space, where
the smallest group of atoms that constitutes the repeating pattern in the material is called
the unit cell of the crystal structure. The unit cells, completely defines the structure of the
entire crystal lattice and are classified into cubic and non-cubic crystal systems according
to the symmetry of their atomic arrangement [for more details see e.g. 6.8]. The crystal
structure and symmetry play important role in determining many physical and mechanical
properties of the material. The atoms in a crystal can be mathematically represented as
points in a three-dimensional real space lattice. If these lattice points are arranged in a
periodic manner, then one can define real space unit vectors a, b, and c and the angles α,
β, and γ to locate the atom in the lattice. The atom positions within the unit cell, as well
as directions and planes in a crystal lattice can be also described by the so-called Miller
index notation, where a three-index notation is used to define particular planes {hkl} or
direction <uvw> in the crystal structure. Crystal structures and their indexing, although
important in understanding many concepts of electron microscopy will not be discussed
further in this book. Many excellent textbooks are available on this subject, and the reader
is referred to them [e.g. 6.8].

Diffraction is an ideal experimental technique to determine a real space lattice structure,


and methods include x-ray, neutron and electron diffraction. The particles in these probes
have dual properties behaving as matter and wave. If the propagation of a wave is described
by the advancement of a wave front then in plane wave propagation the wavevector k is
perpendicular to the plane wave front, where | k| = 2p/l from the de Broglie relationship
[6.9]. It is convenient to define a reciprocal space lattice in the momentum space that is
related to the real space lattice in order to understand diffraction. The symmetry of a real
space lattice and its reciprocal space lattice are related. The unit vectors in the reciprocal
space lattice have a reciprocal relationship with the unit vectors in the real space lattice. In
mathematical term, the reciprocal lattice is the Fourier transform of the real lattice.

The crystal lattice (also known as the direct lattice) exists in real-space, whereas, the reciprocal
lattice (also known as the momentum-space or the k-space lattice) exists in the reciprocal
space. The reciprocal lattice is an array of points in three-dimension like the direct lattice, in
which each point corresponds to a special plane in the crystal lattice. In fact, a plane (hkl)
in real crystal lattice is a point in the reciprocal lattice located at distance d*hkl from the
center (which is equal to the vector g = 1/dhkl, i.e. inverse of the spacing between the (hkl)
planes) as seen in Fig. 6.7b. The distance of a point in the reciprocal lattice from the origin
O, is defined by the vector g, and is called the diffraction vector. Diffraction pattern and
reciprocal lattice are related to each other and this relation is used for the interpretation of
diffraction from the crystal. The reciprocal lattice of a reciprocal lattice, then, is the original
direct lattice again, since the two lattices are Fourier Transforms of each other.

123
ELECTRON MICROSCOPY Microscopy and Analysis

Ewald Sphere
The Ewald sphere is a geometric construct conceived by Paul Peter Ewald [6.10] and used
in x-ray, neutron and electron crystallography to demonstrate relationship between the
wavevector of the incident and diffracted beams. It also defines the diffraction angle for a
given reflection in the reciprocal lattice of the crystal.

Sphere of reflection
radius = ko =1/λ

ki

Fig. 6.8: Ewald sphere construction: sphere of reflection.


ko incident wave, ki diffracted wave.

Join the best at Top master’s programmes


• 3
 3rd place Financial Times worldwide ranking: MSc
the Maastricht University International Business
• 1st place: MSc International Business
School of Business and • 1st place: MSc Financial Economics
• 2nd place: MSc Management of Learning

Economics! • 2nd place: MSc Economics


• 2nd place: MSc Econometrics and Operations Research
• 2nd place: MSc Global Supply Chain Management and
Change
Sources: Keuzegids Master ranking 2013; Elsevier ‘Beste Studies’ ranking 2012;
Financial Times Global Masters in Management ranking 2012

Maastricht
University is
the best specialist
university in the
Visit us and find out why we are the best! Netherlands
(Elsevier)
Master’s Open Day: 22 February 2014

www.mastersopenday.nl

124
ELECTRON MICROSCOPY Microscopy and Analysis

It is shown above that the incident and diffracted waves can be presented as wavevectors
in the k-space. These vectors can be used to think of the different ways it is possible to
orientate the crystal lattice with respect to the incident and the scattered wavevectors. Ewald
showed that if both the orientation of the crystal and the incident direction of the electron
beam are kept fixed then all the diffracted wavevectors will lies on a reflection of sphere whose
radius equals 1/l, i.e. inverse of the wavelength of the incident beam. In two-dimensions,
one can think that due to scattering, all the diffracted wavevectors sweeps out a circle as
shown in Fig. 6.8 (which in 3D space is actually a sphere). For a crystalline specimen,
strong scattering occurs whenever the vector g coincides with a reciprocal lattice point, i.e.
when Bragg’s law or Laue conditions are satisfied. This defines the Ewald sphere (Fig. 6.7b).

BF and DF Image Contrast


Since the diffracted beams are scattered at an angle from the direct beam by the objective
lens, it can be arranged that they do not contribute to the final image. This can be done
in TEM by inserting a suitably sized aperture in the back focal plane of the objective lens.
In such a condition, the TEM image will show a dark contrast wherever local conditions in
the thin specimen produce strong diffraction. The contrast can be reversed by changing from
bright to dark field illumination, i.e. by suitably shifting the aperture to select a diffracted
beam instead of the direct beam. Fig. 6.9 shows the aperture position in the beam path
for a) bright field and b) dark field illuminations. This method of contrast formation has
an advantage over the method of direct lattice resolution in that the atomic array of the
crystal lattice is not resolved and therefore, resolution is not a limiting factor and specimen
requirements are less critical. The contrast in the image is a result of intensity distribution
of electrons coming out of the bottom surface of the thin specimen. Since the BF image
forms with the direct beam only and eliminates all diffracted electrons, the image contrast
is strongly influenced by diffraction. A DF image from the same area of the specimen,
which is imaged only with a single diffracted beam have complimentary contrast to the BF
image. As seen in Fig. 6.10, where what is dark in the BF image is bright in the DF image.
Superimposed upon this is the contribution from the inelastically scattered electrons, which
mainly contribute to the background intensity, and for most purposes (for a thin specimen)
this can be neglected. Any strong variation in the specimen thickness however, will change
the intensity and the phase of the scattered wave coming out of the sample bottom surface
and will contribute to the contrast.

125
ELECTRON MICROSCOPY Microscopy and Analysis

Direct beam

specimen

Objective lens
I
(hkl) (hkl) (1 - I)
aperture
(1 - I) I

Direct beam Direct beam

a. b.

Fig. 6.9: Illustrating diffraction contrast for a) bright field and b) dark field illumination.

Different phases present in the specimen (pure metal, solid solution or intermetallic phase)
has different atomic arrangements and therefore will contribute to diffraction differently.
Therefore, their presence in the sample will give rise to contrast in BF/DF images. Both
the g’ and the g’’ precipitates with the L12 and DO22 crystal structure5 respectively, can be
distinguished from the fcc Ni-solid solution matrix in Fig. 6.10. Also by tilting the specimen,
the incident electron beam will encounter a different atomic arrangement and diffraction
will be affected differently, and this will change image contrast. Since diffraction contrast
is sensitive to the positions of atoms in the crystal, any displacement from their normal
positions will produce a phase difference between waves diffracted from successive scattering
centers. Therefore, strain in the crystal lattice will produce further contrast effects. Strain
can come from defects, such as dislocations, due to second phase precipitates or simply due
to bending of the specimen. In order to understand contrast effects in BF and DF images
it is essential to investigate in details the intensities of electrons scattered by the crystal.

126
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.10: BF / DF image pair of γ’ / γ’’ precipitates in an Inconel 706 Ni-superalloy. The aperture position for
the BF and the DF images are shown in the SAED pattern.

127
ELECTRON MICROSCOPY Microscopy and Analysis

Intensities of Diffracted Beam: Kinematical Theory


The electron diffraction pattern from a crystalline specimen in TEM consists of a system of
spots corresponding to the reciprocal lattice points that are intersected by the Ewald sphere.
The spots can be indexed according to their reflecting (hkl) plane in the crystal for which
diffraction is allowed by the structure factor. However, due to the special geometry of the thin
specimen in the TEM (where, one sample dimension, the thickness, is very small compared
to the other two dimensions) the intensity distribution at the reciprocal lattice point is not
a point but is spiked. The intensity distribution and the shape of the intensity spike at the
reciprocal lattice point can be calculated using kinematical theory of diffraction, with the
main assumption that the crystal is thin enough, so rescattering or absorption is negligible.
The spread in the intensity (spike) of the diffracted spot is shown in Fig. 6.11. This implies
that in TEM even when there is a slight deviation from the exact Bragg condition, the
diffraction spot is still visible on the Ewald sphere. This is a reason why electron diffraction
pattern has many spots, while for x-rays the crystal must be oriented carefully to get spot
patterns. The deviation from the Ewald sphere (in a direction parallel to the direct beam)
is termed the deviation parameter and designated by the vector s. The deviation parameter
controls the image contrast and it is frequently used for quantitative analysis of defects in
TEM. The sg for a particular imaging condition can be accurately determined experimentally
by measuring the displacement of the Kikuchi lines from its corresponding spot. Two-beam
and weak-beam imaging for analysis of dislocation burger vector (b) or stacking fault vector
(R) needs precise control of sg in the microscope operation through specimen tilting.

128
ELECTRON MICROSCOPY Microscopy and Analysis

Column in thin specimen


Incident wave which scatters incident wave

kj kg Deviation from Bragg condition

Exact Bragg condition

Ewald sphere
g+s
Sg Intensity distribution
Spike
along spike
-g +g

Spot pattern on
Ewald sphere -g 0 +g

Kikuchi lines at symmetry position

Fig. 6.11: Schematic illustration of intensity spike at diffracted reflections due to thin specimen in TEM and
deviation from exact Bragg condition showing the deviation vector s. Also shows spot pattern on Ewald sphere
with Kikuchi lines corresponding to the spot at the symmetry position.

129
ELECTRON MICROSCOPY Microscopy and Analysis

Thickness Fringes and Bend Contours – Perfect Crystals


Contrast in TEM images arises due to the difference in the phase of electron waves scattered
through the thin specimen and it is determined by the intensity of the direct and diffracted
beams at the bottom surface of the specimen. Under kinematical considerations, where
diffraction satisfies Bragg condition or deviates slightly from it, the diffracted intensity is
small compared to the intensity of the direct beam. So if the incident beam intensity is
unity and the diffracted beam intensity is I, then the intensity in the direct beam is (1 – I )
as shown in Fig. 6.9. The two-beam and column approximation are used for the calculation
of the intensities. It is assumed that only one diffracted beam together with the direct beam
is excited in the crystal and the intensity at the bottom surface of the crystal is determined
by theoretically dividing the specimen into small columns (along the thickness) across the
specimen (Fig. 6.11) and then calculating the intensity at the bottom of each column. Under
such criteria, there are only three variables to be considered for image contrast: 1) orientation
of the crystal lattice with respect to the incident beam, 2) the specimen thickness and 3) the
deviation from the Bragg condition (s). Thus, a perfect crystal will display contrast only if
the specimen thickness varies or the sample is bent.

Need help with your


dissertation?
Get in-depth feedback & advice from experts in your
topic area. Find out what you can do to improve
the quality of your dissertation!

Get Help Now

Go to www.helpmyassignment.co.uk for more info

130
ELECTRON MICROSCOPY Microscopy and Analysis

Incident beam
k

Fresnel zones

Thin specimen t

P Diffracted beam
kg

Fig. 6.12: Graphical presentation of the column approximation.

For a thin specimen with thickness t, if the diffraction beam is in the direction shown
in Fig. 6.12, the amplitude of the wave at the bottom surface of the specimen can be
calculated by Fresnel zone construction for each plane about the point of intersection of
the diffracted ray with each plane [for details refer to 6.1]. The amplitude of the wave (fg)
at P can then be obtained by adding contributions from all points along a narrow column
parallel to the diffracted wave (the so-called column approximation). Assuming the reflecting
planes to be normal to the specimen surface, the total intensity (Ig) at P, which is square
of the amplitude (fg), is given by

where, xg is the extinction distance6 and at s = 0 (exact Bragg condition) s = 1/xg

The intensity varies along the thickness i.e. it oscillates with depth of the specimen (Fig. 6.13).
The diffracted intensity is periodic in the two independent quantities, t and s.

131
ELECTRON MICROSCOPY Microscopy and Analysis

For imaging conditions where s remains constant but t varies, then fringes are produced in
the image. Fig. 6.13 shows the intensity variations of the direct and the diffracted waves
where the diffracted wave is out of phase by p/2 with respect of the direct wave. In fact,
the intensity oscillations are complementary for the DF and BF images and give rise to
thickness fringes in a wedge crystal (6.13). The depth periodicity of the oscillation is inversely
proportional to the deviation vector s, thus, the greater the deviation from the reflecting
position, the smaller the fringe spacing.

Incident electron beam BF


(intensity = I)

DF
Direct wave Diffracted wave
(1 - I) (I)

a. b.

Fig. 6.13: a) Schematic illustration of the origin of fringe contrast in a crystalline specimen due to
thickness variation. b) TEM image showing actual thickness contrast in BF and DF TEM images.

Similarly, for imaging conditions where t remains constant while s varies locally, then also
contrast result in the BF/DF images. In actual, quite often a thin TEM specimen is uniformly
thick but maybe buckled because of local bending. The Fig. 6.14a shows the basis for bend
contour formation in a foil symmetrically bent either side of the Bragg conditions. For this
geometry, when a set of (hkl) planes satisfy Bragg condition, the reflection g is excited and
the BF image shows lack of intensities at locations where the Bragg condition is exactly
satisfied (s = 0). The DF image is complementary with the BF image, as always. As the
orientation of the crystal in the local region deviates from the Bragg condition (s < 0), the
diffracted intensity is damped with oscillations, forming intensity fringes (Fig. 6.13b). Since
the location in the specimen, where the crystallographic orientations are equal with respect to
the incident beam gives an equal intensity, the fringes are termed equal inclination fringes,
but more commonly known as bend contours. The bend feature of a crystalline specimen is
revealed by the fringes in the image. When the bent specimen is tilted with respect to the
incident beam, the bend contours sweeps across in the image field of view (they are not
fixed to a position) in order to satisfy local diffraction conditions.

132
ELECTRON MICROSCOPY Microscopy and Analysis

S=0 S<0 S=0


S>0 S>0

Thin bend specimen

(hkl) planes

-2g -g 0 +g +2g
a. b.
Fig. 6.14: a) Schematic illustration of the origin of bend contour in a crystalline specimen due to local variation
of s, the deviation from exact Bragg condition. b) TEM image showing fringes due to bend contours in a BF
TEM image.

Brain power By 2020, wind could provide one-tenth of our planet’s


electricity needs. Already today, SKF’s innovative know-
how is crucial to running a large proportion of the
world’s wind turbines.
Up to 25 % of the generating costs relate to mainte-
nance. These can be reduced dramatically thanks to our
systems for on-line condition monitoring and automatic
lubrication. We help make it more economical to create
cleaner, cheaper energy out of thin air.
By sharing our experience, expertise, and creativity,
industries can boost performance beyond expectations.
Therefore we need the best employees who can
meet this challenge!

The Power of Knowledge Engineering

Plug into The Power of Knowledge Engineering.


Visit us at www.skf.com/knowledge

133
ELECTRON MICROSCOPY Microscopy and Analysis

Contrasts due to Crystal Defects


Defects in the crystal introduce displacement of atom positions in the lattice. This can be
either, a translation in the lattice due to planar defects like stacking faults (SF) or they can
be a long-range lattice strain as in the case of dislocations. The phase, the anti-phase and
the grain boundaries are also planar defects, which disrupt the crystal lattice. All of these
displacements give rise to characteristic contrast in TEM image. Contrast development
from the crystal defects in BF/DF images in TEM has been extensively studied and well
understood and have been used to characterize the defect type and their specific character.

Stacking Fault Contrast


Stacking faults are the simplest type of planar defect in a crystal. An inclined SF in a thin
specimen is illustrated in Fig. 6.15, where the crystals above and below the fault plane
are identical and have perfect lattice, but the lattice is translated in the two crystals by a
constant vector R along the fault plane. The phase factor a of the electron waves transmitting
through the crystal determine the contrast due to the fault. As the wave propagate through
the thin specimen, the a changes abruptly at the fault plane from zero above the fault to
the value a = 2p g.R below the fault.

e
an
Crystal (1)
t pl
ul
Fa

R
Crystal (2)

Fig. 6.15: An inclined Stacking Fault across a thin specimen causing the
column to be displaced by a translation vector R along the fault plane.

In TEM two-beam imaging, the g.R factor can be controlled through tilting the specimen
in the electron beam and choosing the particular diffracting beam (g). The formula (a = 2p
g.R) gives the necessary result that the two parts of the crystal (above and below the fault
plane) are in perfect register if R is a lattice translation vector. Thus, the fault is invisible
for all reflections (g) since a merely changes from the value 0 to 2np (where n is an
integer). On the other hand, even if R is not a lattice translation vector, there may still be
particular values of g for which g.R takes integer values and thus the fault will be invisible.
This invisibility criterion is a powerful tool and used for characterizing defects in TEM.
This simple vector relationship between the imaging vector and the crystal defect vector
under two-beam imaging conditions in TEM is the basis for most defect analysis including
characterizing of dislocations. In the case of a dislocation, the invisibility criterion is g.b =
0 where b is the burgers vector of the dislocation.

134
ELECTRON MICROSCOPY Microscopy and Analysis

The contrast in the case of dislocation is complex and will be discussed later. However,
in the case of SF, due to its planar nature a series of dark/bright fringes are produced in
the TEM image. The fringes are parallel to the line of intersection of the fault plane with
the foil surface. The BF / DF contrasts in the imperfect crystal containing a SF can be
determined by applying the same kinematical treatment as for the perfect crystals. Putting
kg = g + s (Fig. 6.11) the phase factor of the diffracted beam becomes

In this, the displacement R due to the SF is at depth z in the crystal. Assuming the column
approximation to hold, the amplitude of the scattered wave at the bottom surface of the
specimen is obtained by integrating over the thickness t of the specimen and given as

Here the imperfection in the crystal due to the SF introduces an additional phase factor
e-iα, where a = 2p g.R. The result of this is schematically illustrated in Fig. 6.16. As the
electron beam encounters the SF at different depths in different columns, the phase of
the emerging wave at the bottom surface is not the same for a specimen with SF that is
inclined to the beam and this gives rise to the fringe contrast. Stacking faults parallel to the
plane of the foil will not exhibit fringes as it encounters all columns at the same depth and
therefore all columns have the same phase at the bottom surface. In general, the image of
such a SF looks either darker or brighter than the background. Fringe contrast occurs due
to the inclination of the planar defect, i.e. because the magnitude of the phase shift varies
according to the depth (z) of the fault in the column and not due to the value of R. Thus,
it occurs in presence of other planar defects and boundaries also, e.g. anti-phase boundary
(APB), grain boundary (GB) etc., all shows special characteristics fringe contrast.

135
ELECTRON MICROSCOPY Microscopy and Analysis

partial dislocation phase shift


R SF
partial dislocation

phase shift

DF SF
BF R

a. b.

Fig. 6.16: Fringe contrast in BF image due to presence of SF a) resulting from phase shift in the direct wave.
Contrast calculation based on kinematical theory and column approximation using two-beam condition,
b) experimental TEM image showing overlapping SFs.

136
ELECTRON MICROSCOPY Microscopy and Analysis

Contrast from Dislocations


Dislocations are a very important type of defect in materials and their movement essentially
controls strength and deformation in the material. The dislocation is a line defect in the
crystal lattice, which give rise to a long-range strain field around its core. The presence of
dislocation therefore causes displacement in the atoms from their original position in the
crystal lattice. Unlike in the case of the SF, where the displacement R has constant value,
for a dislocation the displacement is a continuous function of distance from the core, i.e.
R varies with z.

A dislocation is characterized by two parameters, its Burgers vector7 (b) that measures the
slip displacement associated with the line defect and its unit line vector (l ) that points
locally in the direction of the dislocation line. The nature of the distortion associated with
the dislocation depends upon the relative orientation of b and l. When the two vectors
are parallel, the dislocation is known as a screw dislocation and when they are orthogonal
to each other, the dislocation is an edge dislocation. A mixed dislocation has an arbitrary
angle between its Burgers vector and line vector. The (hkl) plane in the crystal on which
a dislocation can glide/slip (move) is called the glide or slip plane. It is the plane, which
contains both the line and the Burgers vector of the dislocation.

137
ELECTRON MICROSCOPY Microscopy and Analysis

In their paper published in 1960 Hirsch, Howie and Whelan [6.12] dealt with the problem
of contrast formation due to a dislocation in the TEM image following the kinematical theory
and the column approximation. Fig. 6.17a presents an edge dislocation in a simple crystal
lattice, where the line vector is perpendicular to the plane of the paper. Clearly, due to the
presence of the extra half plane in the edge dislocation, the atoms below the plane of the
dislocation are displaced relative to those above by half the Burgers vector (½ b). Imagine
such a dislocation line in a thin crystal lying at a depth z with the line parallel to the foil
surface. If an electron beam through a column passes the center of the dislocation, then the
calculation of the contrast is similar to that of a SF. There will be an abrupt change in the
phase of the wavevector when the wave encounters the plane containing the dislocation. For
other columns not passing through the dislocation center (core), the columns are continuously
deformed and the phase angle a = 2p g.R becomes a continuous function of the position.
The value of a will be different for each column that is located at various distances from
the dislocation core, because R is a function of distance. For a dislocation, the translation
vector R is actually the Burgers vector b. In this simple and easy representation, one can
see how a complex contrast from a dislocation develops in the DF/DF TEM images. The
presence of a dislocation gives rise to dark or bright line contrast respectively in BF or
DF images. The atomic displacements on the two sides of the dislocation are symmetrical
but in opposite direction (Fig. 6.17a), therefore phase angle a due to the displacement
will be added to that with respect to the depth on one side and subtracted on the other
side. Therefore, the image of the dislocation core will be shifted in position to one side in
the micrograph and will not be at the projection of the core position. A typical image of
dislocation lines is shown in Fig. 6.17b.

a. b.
Fig. 6.17: a) Local rotation of lattice planes in a crystal near an edge dislocation. Atoms are displaced from
their original position due to the introduction of the dislocation in the lattice. b) Typical dislocation contrast
in a BF TEM image.

138
ELECTRON MICROSCOPY Microscopy and Analysis

Examples of Defect Characterization


The dot product g.R (for SF) or g.b (for dislocation) determines the contrast of the defect
in the TEM image, where the diffracted beam vector (g) influence the microscopic imaging
condition and the defect vector R or b defines the defect in the crystal. The so-called invisibility
criterion is a powerful method for accurate defect characterization in TEM imaging. Here
through an example the analysis of defects in metallic alloys is described. Essentially, a thin
specimen is tilted in the electron beam and several images obtained of the same defect in
different orientation of the foil. In some of these images, with the particular choice of g
the defect is invisible – i.e. the g.R = 0 or g.b = 0 condition is satisfied. By analyzing all
the image data and knowing the imaging beam vector g in each image, it is possible to
determine the defect vector. The presented example is for overlapping stacking faults in a
titanium alloy, which are bound by partial dislocations.

139
ELECTRON MICROSCOPY Microscopy and Analysis

The results of SF and partial dislocations analysis by S.H. Chen, et al [6.13] in a Ti-Al
alloy after deformation at elevated temperature are presented. The Ti-47Al-2Mn-2Nb alloy
studied had duplex microstructure consisting of primary equiaxed g grains and lamellar a2
+ g colonies. It is a common observation that deformation in the TiAl g phase (with L10
crystal structure) occurs by dislocation activity, which leads to stacking faults bounded by
partial dislocations. Through detailed contrast analysis in TEM Chen et al showed, that
these planar stacking faults are on (111) planes and are associated with different fcc variants
of the Shockley partial dislocations of the type 1/6[112]. This was in contrast with the
observations in stoichiometric binary Ti-Al alloys, where the 1/6[112] type Shockley partials
are found to be associated with the twins.

Imaging was done in TEM under BF two-beam condition to characterize the deformation
substructure (Fig. 6.19). An orientation map of the foil was first established by tilting the
specimen to at least three non-coplanar zone axes in order to index the diffraction patterns
uniquely [6.14]. Further, the correct indexing of the {111} fault planes was verified by checking
the fault plane in the edge-on position. Additionally, indexing was done by considering the
visibility or invisibility of the fault contrast using the established criteria [6.15]. For the
phase angle a = 2pg.R = 0, 2p or an integer multiple of 2p, the SF contrast is invisible,
while for a = ±1/3p, ±2/3p, etc., the SF is visible (where R is the displacement vector
associated with the SF). The visibility/invisibility criteria for partial dislocations used were
those initially described by Howie and Whelan [6.16] and later modified by Clarebrough
[6.17]. That is, for the deviation parameter w = xgs ≥ 0.7, the dislocations are invisible for
g.b = 0, ±1/3, -2/3 and visible for g.b = + 2/3. To remove ambiguity in the visibility of
partials associated with a fault, many imaging conditions where the fault contrast is invisible
were tested. For this purpose, different {113} imaging, vectors were used in the analysis.
The results are shown in Table 6.1 and 6.2.

140
ELECTRON MICROSCOPY Microscopy and Analysis

a.

b.

c.
Fig. 6.18: Transmission electron micrograph used for the analysis of the Burgers vector
of the different partial dislocations bounding stacking faults, where the various partial
dislocations bounding the overlapping stacking faults are labelled a, b, c and d: (a) beam
direction B near >101@, g 11 1 ; (b) B near >101@, g 020; (c) B near >101@, g 1 31. (For more
information, see [6.13]).

141
ELECTRON MICROSCOPY Microscopy and Analysis

Operating Beam Experimental Fault on 111 plane Fault on 1 1 1 plane


reflections Direction visibility a
D = 2Sg.R D = 2Sg.R
(± g) B of the SF
R 1
3
>111@ R 1
3
>1 1 1@
2 02 >101@ i 0 0

111 >101@ v 2
3
S 2
3
S
020 >101@ v 4S , 43 S 4
3
S
13 1 >101@ i 2S 2S
131 >101@ i 2S 2S
2 20 >1 1 2@ v 8
3
S 0

311 >1 1 2@ i 2S 2
3
S
31 1 >1 1 2@ i 2S 2
3
S
a
v = visible, i = invisible
Possible SF plane (111)

Table 6.1: Experimentally observed visibility of the SF (Fig. 6.19) and the calculated values of α for (1 1 1) and
(1 1 1) planes under different imaging conditions (for more information see [6.13]).

142
ELECTRON MICROSCOPY Microscopy and Analysis

Operating Experimental visibility a g.b for 16 112 111 g.b for super
reflections of the partials partial
(± g) a b c d 1
6
>1 2 1@ 1
6
>1 1 2@ 1
6
>2 11@ 1
2
>0 1 1@
2 02 i v i v 0 1 1 1

1 11 i i/v i v 1/3 2/3 1/3 1

020 v/i i v/i v 2/3 1/3 1/3 1

13 1 v i v v 1 0 1 1

13 1 v v v v 1 1 0 2

2 20 i i/v i v 1/3 2/3 1/3 1

3 11 i v i v 0 1 1 1

31 1 i v i v 0 1 1 1
a
v = visible, i = invisible
Possible Burger vectors; > @
ba = bc = ± 16 1 2 1 >
bb = ± 16 1 1 2 @ bd = ± 12 >0 1 1@
Table 6.2: The observed visibility of various dislocations (Fig. 6.19) and the calculated values of g.b for all the
partial dislocations on the (1 1 1) plane (for more information see [6.13]).

Bright Field and Dark Field Image Contrast in STEM


Scanning transmission electron microscopic mode is essentially a combination of the
concept of TEM and SEM. Where maintaining the advantages due to the high acceleration
voltage and the transmission of the electron wave through a thin specimen, as in a TEM,
the image is formed on a suitably placed semiconductor detector by scanning a focused
beam on the specimen, like in SEM. The STEM mode in TEM can be used for image
formation, diffraction pattern generation and microanalysis (EDS and EELS). It is possible
to form scanning images using transmitted and diffracted electrons (BF and DF) as well as
secondary and backscatter electrons. In these modes, magnification is achieved by reducing
the area scanned, like in SEM. Since magnification is achieved electronically and no
magnetic lenses are used for this purpose, there is no rotation of the image involved. The
equivalence of TEM and STEM bright field imaging can be shown from geometric optics
using the Reciprocity Theorem8 [6.18]. In Fig. 6.19, the essential elements of a TEM and
STEM are compared. Due to the reciprocity, the amplitude of wave at the image point B
in TEM due to an electron source at A is equal to the amplitude at A* due to a source
B* in STEM. In the STEM, a BF detector is placed in a conjugate plane to the back focal
plane to intercept the direct beam while a concentric Annular Dark Field (ADF) detector
intercepts the diffracted electrons for a DF image (Fig. 6.20).

143
ELECTRON MICROSCOPY Microscopy and Analysis

TEM STEM
A B*

specimen

specimen

B A*
Fig. 6.19: The geometric optics and reciprocity in TEM and STEM compared.

Transmitted electrons that leave the specimen at relatively low angles with respect to the
optic axis (smaller than a, the semi-convergence angle of the incident beam) are detected
by the bright field detector to form the BF image. While those electrons that leave at
relatively high angles (usually at an angle several times the a) are detected by the annular
dark field (ADF) detector to form the DF image in STEM. Those electrons that diffract to
even higher angles are detected by the HAADF detector. They produce Z-contrast images.
HAADF imaging will be discussed later in the book.

Fig. 6.20: Positions of different detectors for STEM imaging.

144
ELECTRON MICROSCOPY Microscopy and Analysis

The focused probe is a convergent electron beam and the BF and DF detectors are radially
symmetric as shown in Fig 6.20. Therefore, all beams in STEM are characterized by
angles – angle of semi-convergence of the incident beam (a) and the collection angles of
the detectors (b) for the scattered beams. Knowledge of these angles is important for STEM
imaging contrast. BF detector is a solid disc and the dark-field detectors are ring shaped
(annular – with certain inner and outer collection semi-angles b). If the inner b is set-
up to collect only diffracted beams, we obtain a diffraction contrast DF image where the
diffraction intensity is integrated over all the selected beams within these angles. It is like an
integration of multiple conventional DF TEM images. An example is shown in Fig. 6.21.
A consequence of this is that a DF TEM image has strong contrast as only a few grains/
particles (which are in Bragg orientation) shows strong intensity. In contrast, in the ADF
STEM image, more grains satisfy Bragg orientations and show intensity, therefore more
grains are visible but the overall image has less contrast. Moreover, in the STEM image
contrast is influenced by the convergence angle (2a) of the incident beam. Increasing a by
using larger condenser aperture leads to overlap of the diffraction discs (in STEM diffraction
spots are spread in a disc because of the convergent beam). Interference between discs can
lead to phase contrast image in STEM.

Challenge the way we run

EXPERIENCE THE POWER OF


FULL ENGAGEMENT…

RUN FASTER.
RUN LONGER.. READ MORE & PRE-ORDER TODAY
RUN EASIER… WWW.GAITEYE.COM

1349906_A6_4+0.indd 1 22-08-2014 12:56:57

145
ELECTRON MICROSCOPY Microscopy and Analysis

a. b. c.
Fig. 6.21: BF / DF image pair recorded in STEM. a) BF, b) DF and c) Diffraction images respectively.
The position of the BF and the ADF detectors are shown on the diffraction image.

Two-beam Dynamical Theory of Contrast


Although the kinematical approach provides a practical and convenient method for
understanding many of the contrast effects in TEM image of crystalline specimen, it has
limitations particularly in explaining contrast from a relatively thicker foil, where absorption
cannot be neglected. Moreover, as the diffraction spots, contain intense diffraction maxima,
close to the spot, where s is very small (i.e. the crystal is very close to the exact Bragg
condition), the kinematical condition of weak diffraction intensity actually breakdown and
dynamical diffraction become important. Complex mathematical treatment is involved
in dynamical theory [6.19] and such discussions are not in the scope of this book. The
theory considers extensive interaction of the diffracted beams with each other and with the
transmitted beam. Thereby accounts for any intensity exchanges between the direct and
the diffracted beams that take place during the passage of electrons through the specimen.
Under dynamical conditions, the intensity of the diffracted electron beam generally has
intensity in the same order of magnitude as the direct beam.

The starting point for the dynamical theory is the solution of the Schrödinger equation,9
which describe elastic scattering of high-energy electrons in a periodic potential [6.20].
Under the two-beam approximation, the electrons propagating in a column of a prefect
crystal (Fig. 6.22) can be described by a wavefunction10 y(r) according to the Schrödinger
equation and is given by

In contrast to the kinematical theory, where the amplitude of the incident wave f0 was
taken constant, both the wave amplitudes f0 and fg in dynamical theory vary with z, the
depth in the specimen.

146
ELECTRON MICROSCOPY Microscopy and Analysis

dz
t

ф0 фg

Fig. 6.22: Transmitted and diffracted waves φ0 and φg


propagating through an element dz of a column of crystal.

In general, the coupled differential equations of the dynamical theory do not have explicit
analytical solutions. However, the differential equations do have explicit solutions when the
object is a perfect crystal or regions of perfect crystal separated by a fault plane. Howie and
Whelan11 [6.21], in 1961 derived a numerical solution by applying the two-beam column-
approximation for electron diffraction in imperfect crystals. It has become the basis for the
understanding contrast in electron micrographs of crystal defects such as dislocations and
stacking faults. The dynamical theory for electron diffraction under the form developed
by Howie and Whelan [6.21] leads to predictions, which agree remarkably well with the
observations even when only one strongly diffracted beam is taken into account. This is true
for fringe patterns due to planar interfaces [6.19] as well as for dislocation image-profiles
[6.2]. This good agreement suggests that the phenomenological way of taking into account
anomalous absorption by assuming a complex potential is a suitable approach.

147
ELECTRON MICROSCOPY Microscopy and Analysis

6.1.3 HIGH RESOLUTION ELECTRON MICROSCOPY

So far, in all the imaging methods discussed in the book, contrast is formed either by the
direct beam alone (undeviated transmitted beam – BF image), or diffracted beams alone
(single or many beams, DF or ADF images). In the high-resolution-transmission-electron-
microscopy or HREM, both the scattered and the transmitted beams together create an
interference image. If the image is formed with only two coherent beams (the direct beam
and one diffracted beam), by allowing the beams from the same area of the specimen to
pass through the objective aperture, a periodic image can be formed. In this case, instead
of the mass-thickness contrast or diffraction contrast, where the amplitude of the electron
wave varies, the electron optics and the specimen introduce a phase difference between
the scattered and the transmitted beams. This is also true when many diffracted beams are
included in the formation of the image (Fig. 6.23).

This e-book
is made with SETASIGN
SetaPDF

PDF components for PHP developers

www.setasign.com

148
ELECTRON MICROSCOPY Microscopy and Analysis

specimen

objective lens

aperture

image plane
Fig. 6.23: Ray path for HREM image formation.

The incident parallel electron beam, ideally a plane wave, interacts elastically while passing
through the specimen, and the resulting modulations of its phase and amplitude are present
in the electron waves leaving the specimen (Fig. 6.24). The exit wave Ψ(r) thus contains the
information about the object structure. The transmitted and the diffracted waves each have
different phase because they travel different distances in the crystal. Actually, each diffracted
wave has its own phase. This is because each diffracted wave represents a different solution
to the Schrödinger equation for the electrons in the crystal. The resulting phase depends on
the strength and spacing of the periodic potential of the lattice along a given direction in the
crystal. This leads to an interference pattern showing phase contrast in the image. To obtain
a proper lattice image, a large objective aperture must be selected that allows many beams
including the direct beam to pass (Fig. 6.23). In general, more beams that are collected by
the aperture, higher the resolution of the image. If the point resolution of the microscope is
sufficiently high and the crystalline specimen is oriented along a zone axis, then the atomic
structure of the crystal is directly resolved in the HREM image (Fig. 6.25). Thus, HREM is
an imaging mode of the TEM that allows for direct imaging of the lattice structure [6.22].

149
ELECTRON MICROSCOPY Microscopy and Analysis

incident electrons
(plane wave)

thin specimen

transmitted and
diffracted waves
(each has different phase)
Phase contrast HREM image

Fig. 6.24: HREM image formation – interaction of incident electron wave with thin specimen.

Fig. 6.25: HREM lattice image from a Ni-super alloy IN706


which shows the presence of a second phase Ni3Al (γ’) precipitate.

When the incident electron wave passes through a sample, the electron beam interacts
with the sample via scattering and experiences a phase shift, which is manifested in the
electron wavefunction exiting from the bottom of the sample. If it is assumed that the
scattering causes a phase shift but no amplitude shift while traversing the specimen, then
this is known as the Phase Object Approximation. Further, if the specimen is very thin and
a weak scatterer (so that the phase shift is very small) then this approximation is the Weak
Phase Object Approximation. Under these approximations, an analytical expression can be
found for the phase contrast and the wavefunction can be mathematically calculated. The
resultant scattered and transmitted electron beam is then focused by an objective lens and
imaged by a detector in the image plane.

150
ELECTRON MICROSCOPY Microscopy and Analysis

Unfortunately, the magnetic lenses in TEM are poor and especially the objective lens is not
an ideal lens but has aberrations (strong spherical aberration Cs in particular) that reduce
image quality. This aberration causes the diffracted waves to be phase shifted by the objective
lens. The magnitude of the phase shift varies with distance from the optic axis and thus
the diffraction angle. Therefore, each diffracted wave undergoes a different phase shift and
modified in their spatial frequencies. There is a complex dependence of the phase shift on
the wavelength (l), spherical aberration (Cs), diffraction vector (g) and defocus (Df ). This
means that the phases of the diffracted waves are also changed by the objective lens focus
and the lens effectively scrambles the information embedded in the exit wave. The amount of
scramble depends on the Cs and Df. The phase-contrast is thus very sensitive to many factors
and the appearance of the image is sensitive to small changes in the thickness, orientation,
or scattering factor of the specimen, as well as the variations in the focus or astigmatism of
the objective lens. The intensity distribution of the exit wave function is described by the
so-called contrast transfer function12 (CTF).

Free eBook on
Learning & Development
By the Chief Learning Officer of McKinsey

Download Now

151
ELECTRON MICROSCOPY Microscopy and Analysis

Contrast Transfer Theory provides a quantitative method to translate the exit wavefunction
to the final image based on the Fourier transformation13 of the electron beam wavefunction.
When an electron wavefunction passes through a lens, the wavefunction undergoes a Fourier
transform. To understand all this, one simple way is to consider the TEM as an optical
device transferring the information from the specimen to the image. There are losses and
restrictions on the way the information is transferred and additionally, some information is
distorted in the process. Hence, a point in the specimen is distorted and imaged as a disc.
This smearing effect is often referred to as the point spread function (PSF). The transfer of
information can be described in the Fourier space based on spatial frequencies (q):

where G(q) corresponds to the Fourier transform of the disk in the image, F(q) the Fourier
transform of the point in the specimen and H(q) the Fourier transform of the PSF. Further,
the PSF can in turn be expressed as the product of different functions, i.e. the aperture
function, A(q), the envelope function, E(q) and the aberration function, B(q) = e− iχ(q). An
aperture effectively reduces the size of the beam and prevents part of it from being fully
transferred. The envelope function is an intrinsic property of the lens similar to A(q) and the
beam is attenuated as it passes through the lens. The aberration function B(q) is particularly
interesting and corresponds to the distortion of the beam caused by the imperfect objective
lens. It depends on the defocus (∆f ), the wavelength (λ), the spherical aberration constant
(Cs) described in the Fourier space based on spatial frequencies) and the spatial frequency
(q). When expressed in sin χ(q), it is referred to as the contrast transfer function, where:

152
ELECTRON MICROSCOPY Microscopy and Analysis

The sin χ(q) as a function of q is plotted in Fig. 6.26. The complicated curve sinχ (q) strongly
depends on defocus. While the sine function is zero at the origin, it becomes non-zero for
intermediate values of q. In this region of spatial frequencies, all information is transferred
with positive phase contrast and the phase contrast follows the fluctuations of the CTF in the
formed image. Thus, the scattering centers in the specimen, i.e. atom positions in the crystal
lattice, appear with dark contrast. The position of the first crossover corresponds to the point
resolution of the acquired image (Fig.6.26). Up to this point, the CTF has the same phase.
Therefore, the information in HREM images is consequently directly interpretable until
the point resolution. Depending on the defocus value, the CTF may oscillate strongly. After
the crossover, the oscillations in the CTF make the information at high spatial frequencies
difficult to interpret. Moreover, at larger values of q, it is strongly damped due to the effect
of chromatic aberration, spread of focus and instabilities in energy. The best point resolution
for the TEM instrument however, depends on the defocus condition. As the defocus is a
variable and can be adjusted at the microscope, the best imaging condition can be obtained
during microscopy by maximizing the crossover frequency (qres) at an optimum focus keeping
the loss in contrast at an acceptable level. Scherzer defocus14 is an optimum focus for imaging
and its value can be chosen to optimize the imaging conditions. Scherzer defocus can actually
be used to study the degree of useful information transferred by the microscope. Both the
Scherzer defocus Dfs and the point resolution Dc are functions of Cs and l:

However, if the CTF is fully known for the instrument used, image reconstructions from
focal or tilt series can make use of this information as a function of frequencies. All spatial
frequencies are then sampled and transferred to the same phase in the CTF to form the
reconstructed image. The resolution limit in such an image is in theory the information
limit of the microscope as determined by the envelope damping function. This is why,
image simulation is very essential for interpreting and analysis of HREM images. Details
how image simulation is done and the software used are not part of discussion in this book.

153
ELECTRON MICROSCOPY Microscopy and Analysis

Envelope function information


of the TEM limit

point resolution

sin χ(q) q

Fig. 6.26: The contrast transfer function plotted as sin χ(q) vs q. The
undamped function is shown by red dotted line, the symmetrical envelope
damping function by green dashed lines and the resulting damped CTF by
the blue full line.

HREM is a phase contrast imaging technique, which requires high coherence. As spatial
coherence is lost due to the increasing beam divergence, a loss in contrast is observed and
it is essential to keep the spatial coherence as high as possible in conventional TEM.

www.sylvania.com

We do not reinvent
the wheel we reinvent
light.
Fascinating lighting offers an infinite spectrum of
possibilities: Innovative technologies and new
markets provide both opportunities and challenges.
An environment in which your expertise is in high
demand. Enjoy the supportive working atmosphere
within our global group and benefit from international
career paths. Implement sustainable ideas in close
cooperation with other specialists and contribute to
influencing our future. Come and join us in reinventing
light every day.

Light is OSRAM

154
ELECTRON MICROSCOPY Microscopy and Analysis

6.1.4 HIGH ANGLE ANNULAR DARK FIELD IMAGING

Recently, STEM has become a powerful tool for quantitative image analysis at atomic
resolution. Especially, with the introduction of HAADF imaging, incoherent Z-contrast STEM
imaging can provide direct, chemically sensitive structure imaging without the associated
phase problem [6.23]. With this advantage over conventional coherent HREM imaging,
HAADF is not only capable of lattice imaging of crystals, but also the imaging of single
dopant atoms in the specimen [6.24].

Coherent and Incoherent Imaging


In his classic paper on the resolution limit of the light microscope, Lord Rayleigh described the
difference between coherent and incoherent imaging [6.25]. Even for the electron microscopy,
the resolution limit of the imaging systems is usually evaluated in terms of the classical
Rayleigh’s criteria and it is generally concluded that incoherent imaging is better than coherent
imaging. Conventional HREM is a coherent mode of imaging. As described in the previous
section, when an electron transparent specimen is illuminated by coherent plane wave
illumination and the exit surface wavefunction is magnified by an objective lens, a highly
magnified and lattice resolved image could be formed. Due to the inherent lens aberrations
(especially the spherical aberration), there is blurring of the exit surface wavefunction as it
propagates through the objective lens to the magnified image plane. Further, the scattering
from spatially separated parts of the specimen can interfere and add to the blurring process.
Thus, a contrast reversal can take place depending on whether constructive or destructive
interference is occurring and there is uncertainty over the image contrast, i.e. whether atoms
should appear as bright or dark contrast in the image [6.26]. The contrast can further change
depending on the focusing condition of the objective lens and the existence of dynamical
diffraction effects. Therefore, the reading of the image contrast is not straight forward and
image simulation is essential for proper interpretation.

155
ELECTRON MICROSCOPY Microscopy and Analysis

Lord Rayleigh showed that if a transmitting specimen is illuminated with light from an
extended source such that the illumination is spread over a wide range of angles, then the
specimen could be treated as being self-luminous [6.25]. In such cases, the interference
between radiations emitted from spatially separated parts of the specimen will not interfere
and an incoherent image will be formed. The majority of images that we see with our naked
eyes are incoherent, because most light sources are extended and incoherent. Lord Rayleigh
pointed out that incoherent images do not show sharp interference bands that characterize
coherent images and are therefore much simpler to interpret. HAADF Z-contrast images in
STEM are incoherent with a minimum dependence on specimen thickness and microscope
focus. Remembering now the reciprocity (Fig. 6.19), a large source in a conventional TEM
is equivalent to a large detector in the STEM. So, in order to understand contrast of the
incoherent image in STEM, one needs to include the effects of the detector on the image
contrast. Due to the reciprocity, a STEM can be thought of as a conventional TEM with
the electrons propagating in the reverse direction.

High-Resolution Incoherent Imaging with STEM


The lens system in a modern STEM is designed to provide enough demagnification of the
finite sized electron source (generally a FEG source) in order to form an atomic-scale electron
probe at the specimen. The scan coils in STEM, scan this probe over the thin specimen in
a raster and the scattered signal is detected and plotted as a function of the probe position
to form a magnified image. An HREM lattice resolved image could be formed with the
STEM detector with such a small probe.

156
ELECTRON MICROSCOPY Microscopy and Analysis

In the STEM, a partial plane wave in the cone of the coherent illumination is focused by
the objective lens to form the illuminating probe (Fig. 6.27). Contrast in the image arises
from interference between beams in the incident cone (e.g. marked as Ki and Ki+q in
Fig. 6.27) that are scattered into the same final wavevector (Kf) and interfere. Here the two
partial plane waves (Ki and Ki+q) are separated by the reciprocal space vector q. Since the
probe position (r0) is defined in reciprocal space by a linear phase variation over the incident
wave, the phase difference between the two incident waves will be 2pq.r0. If the scattering
is purely Bragg diffraction from a crystalline specimen, a series of Bragg discs (diffraction
spot is not a point but spread to a disc because the incident beam is convergent) will be
seen in the form of a coherent convergent beam electron diffraction (CBED) pattern (for
more details about CBED see next section). For the Bragg beams, interference can only
occur in regions where diffraction discs overlap (Fig. 6.27). A very thin specimen can be
considered to be simply multiplying the illuminating beam by a complex wavefunction
y(r) that describes the magnitude and phase change of the transmitted wave. The Fourier
transform of y(r0) is then Y(q) which is the image intensity for the incident probe at
position r0. It will be recorded on the ADF detector for a point at position Kf in the far
field of the image plane. Scanning the probe, therefore, means that the interference between
the two incident partial plane waves will cause the intensity to oscillate at a rate given by
q, leading to image contrast. The Fourier transform of the complete image is then simply
the integration of all wavefunctions corresponding to the various probe positions over the
detector function. Thus, STEM lattice imaging depends on the detection of interference in
the overlap regions of diffracted discs, as noted by Spence and Cowley [6.27].

Incident convergent beam

Ki Ki + q

thin specimen

Kf
STEM Annular detector
diffraction disc

-g 0 g

Fig. 6.27: Interference of Bragg diffraction beams on the STEM detector.

157
ELECTRON MICROSCOPY Microscopy and Analysis

For a given spatial frequency q, a large number partial plane wave pairs are present in the
incident convergent cone that can contribute to the interference and this is the key of the
incoherent model. Thus, for the incoherent model to apply in STEM HAADF imaging, all
of the available pairs must contribute to the image in a similar way so that a simple integral
can be performed over these pairs. To achieve this, a detector area that is much larger than a
single diffracted disc must be used. This criterion applies to both the detector and its inner
hole. Hartel et al. [6.29] suggested that the inner radius of the ADF detector should be at
least three times the beam convergence angle. Just as Lord Rayleigh proposed that a large
source leads to incoherent imaging in a light microscope, this shows that a large detector
leads to incoherent imaging in STEM.

It is much easier to interpret data from incoherent imaging, as contrast reversal and delocalization
usually associated with conventional HREM images are absent. Generally, bright feature
in the HAADF image can be associated with the presence of atoms (or atomic columns)

360°
in an aligned crystal. Combined with the strong Z-contrast that arises from the high-angle

.
scattering it leads to a high-contrast, chemically sensitive imaging mode. Optimizing the

thinking
conditions for incoherent imaging in STEM is simply a matter of getting the smallest, most
intense probe possible. Use of aberrations to generate contrast (as seen in BF imaging) is
not required.

360°
thinking . 360°
thinking .
Discover the truth at www.deloitte.ca/careers Dis

© Deloitte & Touche LLP and affiliated entities.

Discover the truth at www.deloitte.ca/careers © Deloitte & Touche LLP and affiliated entities.

Deloitte & Touche LLP and affiliated entities.

Discover the truth at www.deloitte.ca/careers


158
ELECTRON MICROSCOPY Microscopy and Analysis

Resolution Limit
The overlap of the transmitted beam (marked 0) and the two diffraction discs (g and –g) in
the white area in Fig. 6.27 is in the BF STEM detector area. The triple overlap conditions
can occur only if the magnitude of g is less than the radius of the aperture. The BF imaging
in STEM shows the usual phase contrast imaging described above, with a CTF that is
controlled by the lens aberrations in a similar way to phase contrast imaging in conventional
TEM. The principle of reciprocity between TEM and STEM is thus confirmed, however, the
BF imaging in STEM is much less efficient than in TEM because the size of the detector is
small and does not collect the majority of the electrons in the detector plane. Whereas, the
ADF detector in STEM (the grey area in Fig. 6.27) covers an annular sensitive region that
detects electrons scattered over an angular range where the coherent effects of elastic scattering
can be neglected because the scattering is almost entirely thermally diffuse scattering (TDS)
[6.28]. This idea has led to the use of the HAADF imaging for HREM. To understand the
nature of this incoherent imaging and the resolution limits that apply, it is useful to consider
a lattice with no thermal vibrations so that the overlapping disc model applies. Fig. 6.27
shows that an ADF detector will not only sum the intensity over entire disc overlap regions
but also sum the intensity over many of such overlap regions. It might then be expected
that this will wash out most of the available image contrast, but surprisingly this is not the
case. Since the overlap region of the diffraction discs is small compared with the physical
size of the ADF detector, the domain of the K integral (limited to the disc overlap region)
is small compared with the domain of the Kf integral. In other words, the contribution of
any overlap regions that are partially detected by the ADF detector is small compared with
the total signal detected. The incoherent image is then regarded as being formed from an
object function that is convolved with a real-positive intensity point-spread function. The
Fourier transform of the image will therefore be a product of the Fourier transform of the
probe intensity and the Fourier transform of the object function. The Fourier transform of
the probe intensity is known as the optical transfer function (OTF). The OTF for incoherent
imaging is compared to the phase CTF for the same defocus and spherical aberration in Fig.
6.28. The radius of the objective aperture used for the OTF is also marked in the figure.
It is seen that unlike the phase CTF for BF imaging in STEM and conventional HREM
imaging, there is no contrast reversals in OTF and it decays monotonically as a function of
spatial frequency. One striking feature of Fig. 6.28 is that the resolution limit for incoherent
imaging appears to be twice that for the point-resolution in conventional HREM imaging.

159
ELECTRON MICROSCOPY Microscopy and Analysis

OTF

Transfer function
aperture radius
q

CTF

Fig. 6.28: The optical transfer function (OTF) for incoherent imaging in STEM
is compared with phase contrast transfer function (CTF) in coherent imaging
in conventional TEM for the same defocus and spherical aberration. The
radius of the objective aperture used for OTF is marked.

A broad statement for STEM resolution is that for a spatial frequency q to show up in
the image, two beams incident on the sample separated by q must be scattered by the
specimen so that they end up in the same final wavevector Kf where they can interfere. This
model of STEM imaging is applicable to any imaging mode, even when TDS or inelastic
scattering is included. It can be immediately concluded that STEM is unable to resolve
any spacing smaller than that allowed by the diameter of the objective aperture, no matter
which imaging mode is used.

160
ELECTRON MICROSCOPY Microscopy and Analysis

6.2 DIFFRACTION
Diffraction is the constructive and destructive interference of waves. Due to a much stronger
interaction of electrons with the matter compared to x-rays, electron diffraction has some
advantages over XRD, e.g. a strong interaction with matter results in strong diffraction
intensity. Secondly, the wavelength of fast moving electrons are much smaller than the
spacing of atomic planes (e.g. at 200 keV the wavelength λ = 0.0025 nm), thus with a
fine probe, spatially localized information can be obtained in electron microscope and
nano-scaled objects that are too small for conventional x-ray diffraction experiment can be
studied by electron diffraction. Further, due to the close relationship between diffraction and
imaging in a selected area, it is possible to combine complementary techniques and obtain
local orientation information of the specimen. However, the very stronger interaction of
electrons with matter and very small diffraction angles in electron diffraction cause strong
dynamical diffraction effects, such as multiple diffraction/scattering, thereby hindering
accurate structural interpretation from electron diffraction patterns.

Diffraction analysis in electron microscopes can be categorized mainly into three groups,
namely: 1. Selected Area Electron Diffraction (SAED), 2. Convergent Beam Electron Diffraction
(CBED) and 3. Electron Back Scatter Diffraction (EBSD). The first two diffraction methods
can be performed in a TEM and the third usually in a SEM.

We will turn your CV into


an opportunity of a lifetime

Do you like cars? Would you like to be a part of a successful brand? Send us your CV on
We will appreciate and reward both your enthusiasm and talent. www.employerforlife.com
Send us your CV. You will be surprised where it can take you.

161
ELECTRON MICROSCOPY Microscopy and Analysis

6.2.1 SELECTED AREA ELECTRON DIFFRACTION

SAED in TEM
In SAED, a parallel beam (plane wave travelling in one direction) interacts with the thin
specimen. An aperture is used to define the area from which the diffraction pattern is to be
recorded, thus the name selected area for this method. This aperture is typically located in
the first image plane below the sample (Fig. 6.29). The size of the area studied by selected
area diffraction is typically a few hundred of nanometers only. The SAED patterns mostly
consist of a simple spot pattern corresponding to single-crystal diffraction (Fig. 6.30a).
Due to multiple diffractions, kinematically forbidden reflections sometime also appear in
the SAED pattern. Quite often however, the specimen is polycrystalline and/or the selected
area contains more than one grain/phase. In such a case, more than one spot pattern is
superimposed in the diffraction image (Fig. 6.30b). If the specimen is a powder sample,
then a ring pattern is formed (Fig. 6.30c), where spots corresponding to randomly oriented
multiple crystals form the diffraction ring pattern. SAED in TEM is commonly used for
phase identification and determination of the orientation of the local grain/phase with
respect to the incident electron beam. The orientation information from the spot pattern
is also used for tilting the specimen to specific crystal orientations in the electron beam.
Imaging in such a tilt series is often used for analysis of defects like dislocations, SF etc. (as
was described in section 6.1). However, compared to XRD, the SAED is less accurate. This
is mainly because the spots are visible on the Ewald sphere even when there is a deviation
from the Bragg condition (Fig. 6.11). Therefore, SAED is hardly used for accurate lattice
parameter measurements of the crystalline phase.

162
ELECTRON MICROSCOPY Microscopy and Analysis

optical axis

electron source

condenser lens

specimen

objective lens
back focal plane/diffraction plane
intermediate image
selected area aperature

intermediate lens

projector lens

diffraction image

Fig. 6.29: Selected area electron diffraction (SAED) image formation.

As the wavelength of high-energy electrons is very small, a few thousandths of a nanometer


and the spacing between atoms in a crystalline solid is about a hundred times larger, the
atoms in the crystal act as a diffraction grating to the electrons, which are eventually
diffracted. When the specimen is so oriented in the electron beam that a set of (hkl) plane
in the crystal satisfy Bragg diffraction condition (see section 3.2), some fraction of the
electron beam will be diffracted to a particular angle, determined by the crystal structure
of the sample, while others continue to pass through the thin specimen without deflection.
This is the so-called two-beam diffraction condition (Fig. 6.31), and it is frequently used
in dislocation/SF characterization. If the specimen is tilted from this position, such that
the same crystal still stay under illumination, then different diffraction conditions will be
activated and different diffraction spots will appear or disappear. Imaging a particular defect
under different two-beam diffraction conditions, change the contrast of the defect in the
image and can make the defect invisible. This information is used to determine the nature
and the character of the defect (see Tables 6.1 and 6.2).

163
ELECTRON MICROSCOPY Microscopy and Analysis

a. b. c.
Fig. 6.30: Examples of selected area electron diffraction (SAED) images. a) from single crystal oriented in a
zone axis, b) from a selected area that contain more than one phase showing superimposed diffraction
patterns, and c) from a powder sample.

164
ELECTRON MICROSCOPY Microscopy and Analysis

Diffraction from one sample orientation in one pattern gives spots from a selected set of
planes in the specimen. When the electron beam is parallel to a low index crystal orientation
[UVW] – which is called a Zone Axis15, multiple beam scattering condition is reached and
many (hkl) planes in the crystal satisfy the Bragg condition. The resulting diffraction pattern
contains many spots in a regular pattern (Fig. 6.30a). For the high-energy electron beam
in TEM, the radius of the Ewald sphere is large and diffraction angles small, therefore,
the section of the Ewald sphere intersected by the diffraction image plane is relatively flat
(Fig. 6.7). Moreover, since the diffraction spot intensity is spread in a spike (Fig. 6.11),
many spots are visible in the zone axis SAED. Thus, the information in the SAED pattern
is only two-dimensional. The SAED represent the reciprocal lattice of the crystal and the
zone axis SAED patterns have symmetry closely related to the symmetry of the crystal lattice
(Fig. 6.32).

Fig. 6.31: Two-beam electron diffraction,


with the transmitted electron beam (0) and
one diffracted beam (g) is frequently used
in defect (e.g. dislocation or SF, see
section 6.1) analysis. The image shown here
is not a perfect two-beam condition in the
strict sense, as other weak spots (especially
in the row) are visible, but it is typically the
image in practical microscopy when only
the two beams has strong intensities.

165
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.32: Zone axis SAED patterns give symmetry information of the crystal lattice. Example shown here is
from an fcc crystal (e.g. aluminum specimen).

CBED in TEM
Instead of parallel illumination with selected area aperture, convergent beam electron
diffraction (CBED) uses highly converged illumination to select a much smaller specimen
region. Typically, a couple of nanometer spot size is possible with convergent beam electron
diffraction, compared to the area of about a couple of hundred nanometers selected by the
selected area aperture in SAED. The convergent incident beam results in a disc of confusion
due to spherical aberration [6.30], which means that the diffraction spots are not points
anymore, but spread in a disc (Fig. 6.33). Thus, although a probe may have a nominal size
of only a few nanometers the disc of confusion, and hence the area from which a CBED
pattern is obtained, is typically hundreds of nanometers in size. Anyway, due to the small
illuminated area by the probe in CBED, little or no thickness and orientation variations in
the specimen are likely. For CBED large illumination angles (2a) are required.

SAED CBED
2α 2α

thin specimen

objective lens

back focal plane

diffraction plane
-g 0 g -g 0 g

diffraction spots diffraction discs

Fig. 6.33: The diffraction pattern and the optical geometry of CBED compared to SAED.

166
ELECTRON MICROSCOPY Microscopy and Analysis

Convergent beam electron diffraction was first described by Kossel and Möllenstedt in
1939 [6.31]. However, it was not until the advent of the STEM in the 1970s that the
technique became more widely used in materials science. Clearly one application of CBED
is acquisition of conventional diffraction information (i.e. determining crystal system, lattice
type, interplanar spacing, and angles) from very small specimen areas. For this application, the
technique is often known as microdiffraction [6.32]. If the specimen is very thin, kinematic
conditions prevail and the discs show uniform contrast. In thicker specimen, dynamical
interactions become important and contrast appears within the discs. Provided the area of
the specimen selected by the beam is strain free the dynamical contrast makes it possible
to derive point and space group information of the crystal. The contrast in the diffraction
disc is packed full with much information.

AXA Global
Graduate Program
Find out more and apply

167
ELECTRON MICROSCOPY Microscopy and Analysis

Geometry of CBED patterns from higher-order Laue zones


In microdiffraction analysis, the diffraction disk diameter can be reduced by using a nearly
parallel electron beam (but still converging), thereby improving the angular resolution and
reducing the diffuse scattering in the diffraction pattern. The zone axis pattern (ZAP), in
microdiffraction is composed of small diameter diffracted disks that exhibit several Laue zones
grouped into the zero order Laue zone (ZOLZ) and the higher order Laue zones (HOLZ)
(Fig. 6.34). It may also contain Kikuchi lines, and sometimes-even HOLZ deficiency lines
or other features due to dynamical effects. A low camera length and wide angular view of
the reciprocal space is generally necessary to record the HOLZ.

incident beam

thin specimen

SOLZ
Ewald sphere
FOLZ
ZOLZ
a. b.

Fig. 6.34: The geometry of CBED: a) the interception of the reflecting sphere with zero-, first- and second-
order Laue zones, b) microdiffraction showing HOLZ diffraction discs.

The structure factors for reflections in higher order Laue zones are far smaller than for reflections
in the zero order Laue zone because the atomic scattering factor for electrons drops off very
rapidly as the Bragg angle increases (far more rapidly than for X-rays). With a parallel electron
beam, as in SAED, therefore, only the ZOLZ reflections are strongly excited because their g
vectors are normal to the electron beam direction, b [UVW] and they satisfy Bragg condition.
In CBED, the reflections in the HOLZ only have a small component of g normal to the
beam direction, b. When a convergent beam is used, the range of incident angles (2a) leads
to significant excitation of the HOLZ reflections as well as causing the reciprocal lattice
points to become discs (Figs. 6.33). Moreover, due to the thin nature of the specimen and the
convergence of the electron beam, the reciprocal lattice points are extended in a cylindrical
shape, with their diameter proportional to the beam convergence angle, a, and the length
inversely proportional to the specimen thickness. The interception of the Ewald sphere with
zero-, first- and second-order Laue zones thus, occurs as shown in Fig. 6.34a. The resulting
CBED pattern is shown in Fig. 6.34b. The CBED patterns contain three-dimensional
crystallographic information, despite being viewed and recorded in two dimensions only.

168
ELECTRON MICROSCOPY Microscopy and Analysis

Dynamical diffraction effects on HOLZ-pattern


There are two kinds of detail seen within the CBED pattern when dynamical effects are
present, namely, diffuse scattering within the discs and sharp lines which are visible both
within the discs and outside them (Fig. 6.35). The diffuse contrast arises from dynamic
interactions within the ZOLZ (therefore, referred as zero-order information). The contrast
is sensitive to the thickness of the specimen and the information can be used to determine
accurately the local foil thickness16. On the other hand, the sharp lines within the discs arising
from three-dimensional diffraction are called HOLZ-lines. They are equivalent to Kossel lines
in X-ray diffraction and are the result of elastic scattering by the planes in higher order Laue
zones. These HOLZ line positions (Fig. 6.36) are sensitive to very small changes in lattice
parameters and their positions are sensitive to small lattice strains [6.33]. CBED is therefore,
often used for accurate lattice parameter determination [6.34]. They occur in pairs (like the
Kikuchi lines) with a bright (excess) line associated with the HOLZ disc (Fig. 6.35b) and
a dark (deficiency) line present in the bright field (0 0 0) disc (Fig. 6.35a). The sharp lines
outside the discs are the result of inelastic scattering from the HOLZ planes (Fig. 6.35b).
They are analogous to the Kikuchi lines that arise from ZOLZ planes in both conventional
SAED and in CBED [6.37]. The direction of the HOLZ line pair is always parallel to each
other. The most important feature of relevance for the lattice parameter measurement is that
the angular position of these lines (Fig. 6.36) is sensitive to the accelerating voltage and the
lattice parameter. Hence, if the accelerating voltage of the incident electrons is maintained
constant and accurately known, the changes in the angular position of these HOLZ lines
can be directly correlated to the variations in the lattice parameter.

Fig. 6.35: CBED patterns: a) Diffuse, zero-order information is visible in all reflections and sharp
HOLZ lines are visible in the (000) spot. b) Bright (excess) line associated with the HOLZ discs.

169
ELECTRON MICROSCOPY Microscopy and Analysis

The HOLZ lines arrangement in a bright field disc of a CBED pattern (Fig. 6.36) reflects the
crystal symmetry. Although, in the kinematical theory of electron diffraction, the position of a
HOLZ line is determined only by the lattice parameters and the electron wavelength (i.e. the
accelerating voltage of the electron microscope), actually, the lines often shift systematically
from the positions predicted by the kinematical approximation [6.38]. The systematic shift
is caused by dynamical effects and can be well compensated in the kinematical simulation if
an effective accelerating voltage (Ee) is adopted instead of the actual one (Ea). The effective
voltage for the microscope operating condition can be calibrated using a measurement on
a standard sample (e.g. Si) with known lattice parameter.

�e Graduate Programme
I joined MITAS because for Engineers and Geoscientists
I wanted real responsibili� www.discovermitas.com
Maersk.com/Mitas �e G
I joined MITAS because for Engine
I wanted real responsibili� Ma

Month 16
I was a construction Mo
supervisor ina const
I was
the North Sea super
advising and the No
Real work he
helping foremen advis
International
al opportunities
Internationa
�ree wo
work
or placements ssolve problems
Real work he
helping fo
International
Internationaal opportunities
�ree wo
work
or placements ssolve pr

170
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.36: A typical CBED HOLZ line pattern in the bright field (000) disc from
Ni3Al g’ precipitate in single crystal Ni-base superalloy. The sample was oriented in
ZA [014]. The HOLZ lines position are sensitive to the unit cell size and the
crystal symmetry.

EBSD in SEM
While both the SAED and the CBED are diffraction analysis methods in TEM, the electron
backscatter diffraction (EBSD), is a diffraction analysis method in SEM. The EBSD technique
gives crystallographic information of the microstructure at the surface of a bulk specimen.
The information in the form of Kikuchi line pattern is characteristic of the local crystal
orientation in the sample region where it is generated. In EBSD, a stationary electron beam
interacts with a tilted crystalline sample (generally tilted to 70° angle with respect to the
beam) and the diffracted electrons form a pattern that can be detected with a fluorescent
screen (Fig. 4.32). Why the specimen needs to be tilted and how EBSD patterns are recorded
is described in this book in section 4.3. Essentially, when the electron probe is raster-scanned
over an area of the specimen, the local crystal orientation can be determined. The Kikuchi
patterns can discriminate between crystallographically different phases, characterize grain,
sub-grain and twin boundaries, and provide information about local strain and texture in
the specimen.

The discovery of the fundamental diffraction phenomenon on which EBSD is based can
be traced back to 1928 and the experiment by Nishikawa and Kikuchi [6.39]. However,
only in 1954, Alam et al. published their paper on High-Angle Kikuchi Patterns [6.40].
Moreover, it was not until the introduction of the commercial SEM that allowed greater
progress during the 1970s decades when three notable discoveries were achieved: namely,
selected area channeling (SACP) [6.41], Kossel diffraction [6.42] and electron backscatter
diffraction (EBSD) [6.42]. In this period a phosphor screen and TV camera was used for
the first time to record EBSD pattern.

171
ELECTRON MICROSCOPY Microscopy and Analysis

In EBSD, the Kikuchi patterns take the form of an array of intersecting bands (Fig. 4.31)
that describes the crystal structure of the material, with the bands corresponding to the (hkl)
planes in the crystalline material. The widths of the bands are directly related to the spacing
of atoms in the diffracting crystal planes and the angles between the bands relates to the
angles between the planes. The type of atoms present in the crystal determines the intensity
of the Kikuchi lines. All information about the crystalline structure of the material can thus
be determined from these patterns. Prior to the development of automated software tools,
the Kikuchi bands had to be identified manually by the operator, locating and drawing lines
on the recorded image. This process was both tedious and prone to error and the analysis
could be done only offline. The problem was finally resolved by using a technique known
as the Hough transform17 to convert the bands in the image into points in a mathematical
construct known as Hough space. The final indexing step involves simply determining the
width of the bands and the angles between them to reveal the crystal structure of the original
material. The first software tools for automatically relating the geometric arrangement of the
Kikuchi bands to the crystalline structure, a process known as indexing was available in the
1990s [6.44]. In today’s systems, the video camera has been replaced by a digital camera,
and the whole system is computer-controlled, with both the SEM beam and the stage on
which the material is held able to be moved automatically. Such systems can capture images
with a resolution as low as 10 nm in just a fraction of a second, thus EBSD analysis can
be performed online.

In general, images or maps reconstructed from automated EBSD data provide an excellent
way to characterize the orientation information in the microstructures of a polycrystalline
specimen. This technique is also referred to as Orientation Imaging Microscopy (OIM).
By scanning the electron probe on the specimen surface, EBSD can build up a picture of
the crystal structure across the entire surface of the material. In polycrystalline specimen,
the same basic crystal structure occurs in a variety of different orientations, as each grain
is oriented differently to the electron beam. The grains in the microstructure are separated
by grain boundaries. Similarly, if the microstructure contains more than one phase, their
crystal structure and therefore, their orientation with respect to the beam is also different.
However, the crystal orientation can only be calculated if the crystal structure of the phase
is known. Therefore, it is necessary first to determine the phases present in the material.
If this is already known, orientations can be calculated. An example of OIM image of
differently oriented grains in a Co-Re alloy specimen is shown in Fig. 6.37. In Fig. 6.38,
OIM images from a multi-phase structure are shown. The figures include the so-called
inverse pole figure (IPF) and phase maps.

172
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.37: EBSD image of different grain orientation and the determination
of grain boundaries and small angle boundaries in a Co-Re alloy, from
local misorientation based on the boundaries rotational angle.

93%
OF MIM STUDENTS ARE
WORKING IN THEIR SECTOR 3 MONTHS
FOLLOWING GRADUATION

MASTER IN MANAGEMENT
• STUDY IN THE CENTER OF MADRID AND TAKE ADVANTAGE OF THE UNIQUE OPPORTUNITIES
Length: 1O MONTHS
THAT THE CAPITAL OF SPAIN OFFERS
Av. Experience: 1 YEAR
• PROPEL YOUR EDUCATION BY EARNING A DOUBLE DEGREE THAT BEST SUITS YOUR
Language: ENGLISH / SPANISH
PROFESSIONAL GOALS
Format: FULL-TIME
• STUDY A SEMESTER ABROAD AND BECOME A GLOBAL CITIZEN WITH THE BEYOND BORDERS
Intakes: SEPT / FEB
EXPERIENCE

5 Specializations #10 WORLDWIDE 55 Nationalities


MASTER IN MANAGEMENT
Personalize your program FINANCIAL TIMES
in class

www.ie.edu/master-management mim.admissions@ie.edu Follow us on IE MIM Experience

173
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.38: EBSD phase map image of a two-phase structure in a Co-Re alloy.

6.3 ANALYSIS
Microanalysis methods identify the chemical elements present either within or on the surface
of a specimen and both TEM and SEM are suitable instruments for performing such an
analysis. In most microanalysis methods, the sample can be analyzed in-situ and the results
can be either qualitative (i.e. identification and the distribution of different atoms in the
sample) or quantitative (determination of composition in the local area or in the phase).
The analysis technique in electron microscopes involve focusing an electron beam on the
specimen and measuring an output beam (e.g. x-ray/electron) that results from the interaction
of the input beam with the atoms and molecules making up the sample. X-rays are generated
when the electron beam strikes the specimen in the analytical electron microscope (AEM),
giving rise to through thickness composition information (Section 3.3). The x-ray source
is localized in the interaction volume where the electron probe interacts with the specimen
(Fig. 3.9) and for TEM and SEM, this volume is small, therefore the spatial resolution is
high and much improved than bulk x-ray analysis methods. In modern microscopes, the
AEM technique is well developed and the information generated is accurate, reproducible
and can be used to understand the material character and solve engineering problems.
In many literature characterization of crystal structure information of the specimen, like
microdiffraction or CBED are also included in the microanalysis methods. In this book by
microanalysis, we mean only the chemical analysis of the composition at micro/nano scale.
The main techniques include energy dispersive spectroscopy (EDS), wavelength dispersive
spectroscopy (WDS) and energy loss spectroscopy (EELS). The first two of these methods
analyze the x-ray spectra emitted from the specimen on electron bombardment and differ
mainly in the characteristic of the spectra that is analyzed. While in the third method of
analysis the distribution of electron energies emergent from a thin-foil specimen are analyzed.

174
ELECTRON MICROSCOPY Microscopy and Analysis

6.3.1 ENERGY DISPERSIVE SPECTROSCOPY

EDS in TEM and SEM


Energy dispersive spectroscopy (EDS) analyzes the x-rays emitted from the specimen, and
therefore is variously known as energy dispersive x-ray-spectroscopy (EDX) or x-ray energy
dispersive spectroscopy (XEDS). In this method, information can be obtained on the chemical
composition of the specimen for all elements with atomic number (Z) greater than 3,
i.e. Beryllium (Z = 4) and above in the periodic table, although not all EDS instruments
are equipped for the analysis of light elements (Z < 10). As EDS analysis relies on x-ray
generation in the specimen due to the interaction of the primary electron beam, there are
some limitations in the method. Firstly, the elements of the first period, i.e. Hydrogen (Z =
1) and Helium (Z = 2) cannot be detected by EDS. A paper by Stojilovic [6.46] describes
why hydrogen cannot be detected in x-ray photoelectron spectroscopy (XPS). The reason is
simple, that is the EDS measurement is essentially related to x-ray generation from the K
shells, which are not the valence shell. Since H and He have K shells in covalent bonding, their
electrons are shared. Thus, they do not give characteristic x-rays. The atoms of Li (Z = 3)
although have a K shell which leads to x-ray emission, the characteristic energy is too low to
be detected. Thus, Li also cannot be detected by EDS. Moreover, the analysis of the light
elements, i.e. Be, B, C, N, O and F by EDS is difficult because of their low photon energies.
X-rays from Beryllium (Z = 4) to Neon (Z = 10) can be detected by EDS, but there are two
problems. Firstly, they are low energy x-rays and have a high absorption in the specimen
and in the detector. Secondly, light elements give a low yield of x-rays and the low energy
peaks are positioned close to about 0 keV, where the electronic noise of the detection system
is high and thus the peak to background ratio is low. Heavier elements (Z > 10) are easier
to analyze by EDS, but their quantification also depends on the amount of the particular
atoms present in the alloy/phase. Moreover, the EDS system uses a window in front of the
detector to provide a barrier to maintain vacuum within the detector. In the older systems,
this window was made of beryllium, thus it was not possible to detect Be in older system.
Modern EDS system uses polymer-based ultra-thin window and thus Be can be detected.
Qualitative analysis involves only the identification of the characteristic energy lines in the
spectrum, but for quantitative analysis (determination of the concentrations of the elements
present) measuring the line intensities (photon counts) for each element in the sample is
needed. This intensity for the same elements in a standard specimen of known composition
is also measured under the same microscopy condition (voltage, current) for calibration.

Output from EDS analysis


The EDS measurement can be performed in different ways and the output of the measurement
therefore can be varied. That is to say, analysis can be made from a spot were the primary electron
beam strikes the specimen, or the electron probe can be scanned over an area on the surface of
the specimen or along a line and the EDS spectrum obtained at each scanned spot position.

175
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.39: EDS spectrum from a spot analysis in a Co-Re-Cr alloy. Measurement was done with an electron
acceleration voltage of 20 kV in SEM. The characteristics K and L line peaks from the elements Co, Re and Cr
are seen.

176
ELECTRON MICROSCOPY Microscopy and Analysis

EDS Spectrum
The EDS spectrum consists of a plot of the number of x-ray counts detected from various
atoms in the specimen versus their energies. Characteristic x-rays form peaks is superimposed
on the Bremsstrahlung x-ray as described in section 3.3 (Fig. 3.13). A typical experimentally
obtained EDS spectrum is shown in Fig. 6.39. The Characteristic x-rays allow the elements
present in the specimen (in the example shown, Co, Re and Cr) to be identified from their
energy values and quantified by measuring the area under the peaks.

EDS Map and Line Scan


The EDS map is an image showing the distribution of phases (Phase Map) or the elements
(Elemental Maps) over an area measured on the specimen. Phases are distinguished by their
different chemical composition and the concentration variation of each element is plotted
in the elemental maps. In the phase map, each different color represents area with same
composition (i.e. similar spectra) and it is obtained by comparing the spectra at each point
on the raster. Usually several scans of the area are made during measurement and the spectra
obtained at each point are integrated to improve the peak to background ratio of the x-ray
counts. With modern SDD detectors, very high x-ray counts can be obtained and the
advancement in computer software allows comparing spectra live during the measurement
to identify the different phases in the specimen. Typical examples of phase and elemental
maps are shown in Fig. 6.40. It is however, seen in Fig. 6.40 that the boron map does not
reveal enough B counts because this is one of the light elements and additionally, in the
investigated alloy B is only present in very low quantity (1000 ppm). The phase map could
distinguish all the 3 phases present, namely e Co matrix and s and Cr2B phases.

Fig. 6.40: EDS map from a Co-Re-Cr-B alloy specimen, showing phase map and elemental maps from
Co, Re, Cr and B.

Similar presentation can also be done for measurement along a line. In Fig. 6.41, an example
of the Line scan is presented. Again, it is seen that EDS could not reveal the difference in
B content between the matrix and the Cr2B phase and the B counts are very low.

177
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.41: EDS line scan from a Co-Re-Cr-B alloy specimen, showing elemental variation along
the line. Co, Re, Cr and B scans are shown.

Quantitative EDS Microanalysis


In quantitative EDS microanalysis, the weight/atomic percentage of the elements present
in the specimen is calculated. EDS is particularly practicable for quantification because the
number of x-rays emitted by each element present in the specimen bears a direct relationship
with the concentration of that element (mass or atomic fraction) in the specimen. This is
why it is possible to convert the x-ray measurements into a final x-ray spectrum and assess
the chemical composition. The acquired spectra are processed to remove Bremsstrahlung
x-rays (background intensity), and then the characteristic x-rays from the particular elements
are compared with data measured on a standard reference materials. The EDS employs pulse
height analysis and the detector outputs a pulse proportional in height to the x-ray photon
energy. This is used in conjunction with a pulse height analyzer. A solid-state detector is
generally used because of its good energy resolution. As described in section 4.3, x-ray
photons cause ionization in the detector (Si(Li) or SDD) and produce an electrical charge.
An SDD has less electronic noise than the Si(Li) detector and therefore has a better energy
resolution at high count rates. The acquired spectrum is displayed with the x-axis representing
x-ray energy (usually in channels of 10 or 20 eV wide) and the y-axis representing the
number of photon counts per channel (Fig. 6.39). The characteristic x-ray line (essentially
of mono-energetic photons) is actually broadened to a Gaussian profile by the response of
the system. Energy resolution is defined as the full width half maximum (FWHM) height
of the peak. Generally, this is specified for the Mn Kα peak at 5.89 keV and for Si(Li) and
SDD detectors, the value is in the range of 130–150 eV.

178
ELECTRON MICROSCOPY Microscopy and Analysis

The optimum choice of the accelerating voltage in the microscope for the EDS measurement
is determined by the element in the specimen, which has the highest atomic number. In
order to obtain adequate intensity, the electron accelerating voltage (in kV) should be not
less than twice the highest excitation energy Ec (in keV) of the element. In the example
of Fig. 6.39, an accelerating voltage of 20 kV was adequate to excite the Re Lβ-lines. The
electron beam – specimen – detector geometry is generally optimized for the microscope,
so the other important variable is the beam current. A higher beam current increase the
x-ray intensity, but it also means more damage to the specimen.

Castaing [6.47] first showed that the relative intensity of an x-ray line is approximately
proportional to the mass concentration of the element concerned. Since it is difficult to
measure an absolute intensity, he compared the measured value to a standard sample containing
the element. An apparent concentration (Csp) can be derived using the following relationship:

In the past 5 years we have drilled around

95,000 km
—that’s more than twice around the world.

Who are we?


We are the world’s leading provider of reservoir characterization,
drilling, production, and processing technologies to the oil and
gas industry.

Who are we looking for?


We offer countless opportunities in the following domains:
n Operations
n Research, Engineering, and Manufacturing
n Geoscience and Petrotechnical
n Commercial and Business

We’re looking for high-energy, self-motivated graduates


with vision and integrity to join our team. What will you be?

careers.slb.com

179
ELECTRON MICROSCOPY Microscopy and Analysis

where, Cstd is the concentration in the standard and Isp and Istd are the x-ray intensities
measured in the specimen and the standard sample. To obtain the true concentration, certain
corrections are required. K is a sensitivity factor and not a constant, determined (inversely)
by: the atomic number (Z), the absorption of x-rays within the specimen (A) and the
fluorescence of x-rays within the specimen (F ), i.e. the so-called ZAF correction factors. In
1975, Cliff and Lorimer showed that a standard sample is not needed, if intensities for two
elements are gathered simultaneously and compared [6.48] and thereby the EDS system is
calibrated. This method of standard-less analysis came to be known as Cliff-Lorimer method.

C + C = 100%
A B

where, kAB is a different sensitivity factor, called the Cliff-Lorimer factor, which can be
determined both theoretically and experimentally (with standards). CA and CB are the weight
fractions of elements A and B respectively, and IA and IB are the characteristic x-ray intensities
of the two elements measured simultaneously and kAB is a factor which is constant for a
given acceleration voltage. To make the microanalysis reliable, it is important to determine
accurate values of the kAB factors for the elements of interest. A typical quantitative analysis
results obtained from the spectrum shown in Fig. 6.39 using k ratio and ZAF correction
is shown in Table 6.3.

180
ELECTRON MICROSCOPY Microscopy and Analysis

6.3.2 WAVELENGTH DISPERSIVE SPECTROSCOPY

WDS in SEM
While data collection and analysis with EDS is a relatively quick and simple process because
the complete spectrum over the whole energy range is acquired at the same time, the method
is not very sensitive to light elements (Z < 10). On the other hand, the spectrum acquisition
in WDS is sequential and slow, but it has a much improved energy resolution compared
to EDS. Typically, FWHM of a peak in WDS is in the range of 2 to 35 eV, whereas the
EDS detector has resolution in the order of 60 to 130 eV (depending on the element). This
combination of better resolution and a large signal amplification (typically103–105 times) in
the gas proportional counters used in the detector to count photon energy, allows WDS to
detect elements at an order of magnitude lower concentration than in EDS. The detection
limit in WDS is in the level of parts per million (e.g. 200–1000 ppm) as compared to
0.1–0.5 weight percentage in EDS. The higher peak resolution in WDS is compared to
EDS in Fig. 6.42. The elemental peaks (e.g. Si and Re) which are overlapped in the EDS
spectrum are well resolved and separated in the WDS spectrum. Therefore, in addition to
the analysis of light elements and trace elements, WDS is also useful for peak separation
and deconvolution for heavier elements.

Fig. 6.42: EDS and WDS peaks compared in a Co-Re-Cr-Si alloy. The overlapped peaks of Si and
Re in EDS spectrum are well separated in the WDS spectrum.

181
ELECTRON MICROSCOPY Microscopy and Analysis

Like in the EDS spectroscopy, the WDS system comprises of three basic components that
are designed to work together: the x-ray spectrometer, measurement electronics and the
analytical software. Although WDS generally requires a higher beam current than EDS,
the x-ray data in both EDS and WDS can be obtained simultaneously if needed. Thus, a
combined EDS + WDS analysis, including mapping is possible. In WDS, analyzing crystal
(diffractor) with specific lattice spacing is used to diffract the x-rays generated in the specimen
on to the detector, where the wavelength of the reflected x-rays can be varied by changing
the position of the analyzing crystal relative to the sample (Fig. 4.27). At each position of
the diffractor, the detector counts x-ray photon of a single wavelength (energy) at any given
time and by changing the crystal and the detector positions sequentially, a plot of x-ray
counts as a function of wavelength is recorded. A typical wavelength dispersive spectrum for
the element Re (Re Lα1 line at 8.652 KeV) is shown in Fig. 6.43. Generally, a measurement
over the entire energy range is done to cover the complete peak (Fig. 6.43a). However,
since WDS measures sequentially at each wavelength, this is time consuming. Moreover, for
quantitative analysis only the peak height and the peak to background ratio are important.
Therefore, to optimize the measurement time, it is adequate to measure the peak height in
a much narrower wavelength (energy) range in the so-called reduced scanning mode and
to estimate the background count levels at extreme energies (Fig. 6.43b).

Excellent Economics and Business programmes at:

“The perfect start


of a successful,
international career.”

CLICK HERE
to discover why both socially
and academically the University
of Groningen is one of the best
places for a student to be
www.rug.nl/feb/education

182
ELECTRON MICROSCOPY Microscopy and Analysis

a. b.
Fig. 6.43: WDS spectrum in the energy range 8.480 and 9.000 KeV measured to detect Re L peaks in a Co-Re
base alloy. a) Continuous measurement over the whole range of energy, b) measurements in reduced scan mode:
the background counts are recorded at the two extreme energies and measurement of the peak in selected
energy range.

In order to cover a wide range to wavelengths (energies) for measurements and detection of
different light and heavy elements, different analyzing crystals with specific lattice spacing’s
are used. Some of the commonly used crystals are lithium fluoride (LiF: 2d = 2.848 Å,
analyzing energy range 4712–15330 eV, for element range V to Y), pentaerythritol tetrakis
(PET: 2d = 8.74 Å, 1535–4994 eV, for Si to Ti) and thallium acid phthalate (TAP: 2d =
25.9 Å, 521–1695 eV for O to Si). However, for lower energies (long wavelength) x-rays,
larger d spacing are required, which are difficult to obtain in natural crystals and layered
synthetic crystals (LSM – e.g. Ni-C type: 2d = ~78 Å, 172–564 –eV, for B to O) are used.

Fig. 6.44: WDS peaks from the elements Co, Re, Cr and Si in a Co-Re-Cr-Si alloy and are superimposed on the
EDS spectrum for comparison.

183
ELECTRON MICROSCOPY Microscopy and Analysis

A typical WDS measurement covering a range of different elements (Co, Re, Cr and Si) in
a Co-Re-Cr-Si alloy is shown in Fig. 6.44, which compares the WDS measurements with
the EDS. WDS measurement required measuring different elements separately using the
most appropriate analyzing crystal for the particular energy range. In modern WDS system
with parallel beam optic such measurement can be made in one setting, as different crystals
are mounted in a turret and the detector moves in an arc (Fig. 6.45). The operator then
chooses the particular analyzing crystal for each element to be measured and the system
makes measurement for each element at a time, suitably changing the diffractor for the
different wavelength ranges. As mentioned before (see section 4.3) the x-ray counts on the
detector is sensitive to the alignment of the x-ray optic with the sample, particularly the
correct working distance in the SEM must be ensured, so that x-rays are incident on the
diffractor. For this, a height calibration prior to WDS measurement is necessary, which can
be manually done separately, but modern software control in WDS systems and piezo motor
driven SEM stage makes the height adjustment automatically (software driven) possible.

Axis of turret rotation

LIF TAP
L PET

Different diffractors
on the faces of turret

Detector

Arc of detector motion


Fig. 6.45: Schematic representation of a parallel beam spectrometer, showing a fixed L
length movement of the detector with respect to the diffractor turret.

184
ELECTRON MICROSCOPY Microscopy and Analysis

Quantitative WDS Microanalysis


Quantitative WDS microanalysis (weight/atomic percentage) of the elements present in
the specimen can be determined. In order to perform quantitative analysis with a WDS
spectrometer, the k-ratios for each element must be determined, like in the EDS. The k-ratio,
which is the ratio of the intensity of the unknown element divided by the pure element
intensity, is measured under the same conditions of accelerating voltage and beam current.
In a first approximation, the k-ratio is the weight fraction of the element in the sample,
but it can be in error due to interference of other elements in the matrix. Therefore, it is
essential to use standard samples, which can be either pure element standards or compounds
having the appropriate element in known quantity. The standards and unknown sample are
measured under the same conditions, i.e. same kV, step-size (eV/channel), scan-size and
the same start/end eV values, but the dwell times and beam current can be different. Since
x-ray detection by the detector is sensitive to sample height with respective to the x-ray
optics, a height optimization routine is performed before the actual measurement. Once the
optimum stage height is determined by maximizing the intensity and the resolution on a
relatively high-energy x-ray line (an element with sufficient counts is usually chosen for this),
then the working distance is noted, and the same working distance is used for the sample
to be measured. The reproducibility of the stage height in the SEM is important due to
the switching between the sample and the standard in the beam and it should be within a
few microns only for quantitative WDS measurement. Generally, high beam current (> 10
nA) is preferable for WDS quantification, which ensure enough x-ray counts. In addition,
the value of the beam current (not the specimen current) must be known accurately and
therefore, the current is measured using a Faraday cup setup before or after each sample
measurement. Usually a high magnification or a spot mode is used for quantification by
WDS because of the limited field of view of the spectrometer. Quantification of results in
combined EDS + WDS measurements is however, also possible, for example, the heavier
elements measured by EDS and light elements or overlapped elemental peaks by WDS.
Similarly, EDS + WDS combined mapping is possible (Fig. 6.46). When the EDS and
WDS elemental maps are compared, the light elements B and C clearly show higher counts
in the WDS maps.

185
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.46: EDS + WDS combined map from a Co-Re-Cr-B-C alloy specimen, showing elemental maps from Co,
Re, Cr, B and C. The elements B and C are measured both, by the EDS and the WDS and compared.

American online
LIGS University
is currently enrolling in the
Interactive Online BBA, MBA, MSc,
DBA and PhD programs:

▶▶ enroll by September 30th, 2014 and


▶▶ save up to 16% on the tuition!
▶▶ pay in 10 installments / 2 years
▶▶ Interactive Online education
▶▶ visit www.ligsuniversity.com to
find out more!

Note: LIGS University is not accredited by any


nationally recognized accrediting agency listed
by the US Secretary of Education.
More info here.

186
ELECTRON MICROSCOPY Microscopy and Analysis

6.3.3 ELECTRON ENERGY LOSS SPECTROSCOPY

EELS in TEM
Electron energy-loss spectroscopy offers unique possibilities for material analysis at nanometer
resolution, due to the broad range of inelastic interactions of the high-energy electrons with
the specimen atoms (which range from phonon interactions to ionization processes). It can
be used to map the elemental composition of a specimen, but also for studying the physical
and chemical properties of a wide range of materials. The analysis of the distribution of
electron energies emergent from a thin-foil specimen constitutes the EELS method. The
spectrometry is performed by a magnetic prism spectrometer positioned beneath the final
viewing screen of the TEM, which disperses the electrons entering it according to their energy
(Fig. 4.21). The wide use of energy filtering in TEM (EFTEM) also led to its application
as microanalysis method for nanoscale characterization of thin samples. The method is
applicable for analysis of most chemical elements and particularly sensitive to light elements.

One of the most commonly used applications of STEM-EELS or EFTEM is to derive


compositional information by recording energy-filtered images using the element characteristic
ionization edges. Two examples of ELLS analysis are shown below. The first example is the
analysis of Ti-carbide in a ß-titanium alloy and in the second example carbon nanotube
produced by chemical vapor deposition (CVD) are analyzed by EELS.

The titanium carbide particles in a Ti-15V-3Sn-3Cr-3Al alloy was studied by EELS. The
EELS spectrum from the carbide particle is shown in Fig. 6.47. It identifies the presence
of Ti and C energy edges in the EELS spectrum. The corresponding C and Ti jump ratio
images (from EELS) are shown in Fig. 6.48 respectively, which reveal carbon enrichment
in the carbide particle (Fig. 6.48b) and almost uniform distribution of Ti in the carbide
and the matrix (Fig. 6.48a).

Fig. 6.47: EELS spectrum showing carbon and titanium edge from a carbide particle [6.49].

187
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.48: (a) Carbon and (b) titanium jump ratio images from EELS analysis showing distribution
of C and Ti in TiC and matrix [6.49].

Carbon nanotube grown by CVD on iron catalyst film was deposited both as iron islands
and continuous film and their growth was aligned and patterned. The analysis by EELS of
one such nanotube is shown as the second example. A single carbon nanotube (which is
finer than 100 nm in diameter) is imaged by TEM (Fig. 6.49). The nanotube is multiwalled
and a particle (dark contrast) is seen encapsulated at the tip of the nanotube. Electron
energy loss spectrum (Fig. 6.50) confirmed that the particle at the tip as Fe particle, which
came from the catalyst film. Figure 6.51 shows edge of carbon peak from carbon nanotube,
which is sp2, hybridized. Moreover, a fine difference between carbon from the nanotube
and that from carbon in the amorphous carbon coated grid can be detected by EELS, as
seen from the Fig. 6.52.

Fig. 6.49: A TEM micrograph showing a large particle


at the tip of a carbon nanotube [6.49].

188
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.50: Electron energy loss spectroscopic study of the particle at tip of the carbon
nanotube in Fig. 6.49 shows presence of Fe peak [6.49].

189
ELECTRON MICROSCOPY Microscopy and Analysis

Fig. 6.51: EELS spectrum showing carbon edge from the carbon nanotube (CNT) [6.49].

Fig. 6.52: No distinct edge can be seen from the amorphous carbon of the carbon grid [6.49].

6.4 REFERENCES
6.1. 
P.B. Hirsch, A. Howie, R.B. Nicholson, D.W. Pashley, M.J. Whelan, Electron
Microscopy of Thin Crystals, Butterworth, London, (1965) pages 549.

6.2. G. Thomas, Transmission Electron Microscopy of Metals, John Wiley and Sons, New
York (1962) pages 299.

6.3. L. Reimer, Scanning Electron Microscopy: Physics of Image Formation and Microanalysis,
3rd edition, Springer Series in Optical Sciences, volume 45, Springer-Verlag, Berlin
Heidelberg (1998) pages 545. ISBN 978-3-540-56849-0

6.4. J.P. Spencer, C.J. Humphreys, P.B. Hirsch, A dynamical theory for the contrast of perfect
and imperfect crystals in the scanning electron microscope using backscattered electrons,
Phil. Mag 26 (1972) 193–213. Doi 10.1080/14786437208221029

190
ELECTRON MICROSCOPY Microscopy and Analysis

6.5. 
S. Zaefferer, N.N. Elhami, Theory and application of electron channeling contrast
imaging under controlled diffraction conditions, Acta Materialia 75 (2014) 20–50. doi:
10.1016/j.actamat.2014.04.018

6.6. D. Mukherji, G. Pigozzi, F. Schmitz, O. Näth, J. Rösler, G. Kostorz, Nano-structured


materials produced from simple metallic alloys, Nanotechnology 16 (2005) 2176–2187.
doi:10.1088/0957-4484/16/10/034

6.7. M. Laue, Eine quantitative Prüfung der Theorie für die Interferenz-Erscheinungen bei
Röntgenstrahlen, Annalen der Physik 346, (1913) 989–1002. M. Laue Version of
Record online: 14 MAR 2006, doi: 10.1002/andp.19133461005

6.8. L.H. Schwartz, J.B. Cohen, Diffraction from Materials, 2nd edition, Springer-Verlag,
Berlin Heidelberg (1987) pages 549. ISBN 3-540-17114-2

6.9. G.-C. Wang, T.-M. Lu, RHEED Transmission Mode and Pole Figures: Thin Film and
Nanostructure Texture Analysis, Chapter 2: Crystal Lattices and Reciprocal Lattices,
Springer-Verlag, Berlin Heidelberg (2014) pages 227. ISBN 978-1-4939-5366-0

6.10. P.P. Ewald, Introduction to the dynamic theory of X-ray diffraction, Acta Cryst. A25
(1969) 103–108. https://doi.org/10.1107/S0567739469000155

6.11. D. Hull, D.J. Bacon, Introduction to Dislocations, Butterworth-Heinemann, Oxford,


(2001) pages 257. ISBN 978-0-08-096672-4

6.12. P.B. Hirsch, A. Howie and M.J. Whelan, A Kinematical Theory of Diffraction Contrast
of Electron Transmission Microscope Images of Dislocations and other Defects, Phil. Trans.
R. Soc. Lond. A 252 (1960) 499–529. doi: 10.1098/rsta.1960.0013

6.13. S.H. Chen, D. Mukherji, G. Schumacher, G. Frohberg, R.P. Wahi, Observation of


planar stacking faults in a Ti-rich two-phase Ti-Al alloy after deformation at elevated
temperatures, Philosophical Magazine Letters, 80 (2000) 19–26. http://dx.doi.
org/10.1080/095008300176416

6.14. A.K. Head, P. Humble, L.M. Clarebrough, A.J. Morton, C.T. Forwood, Computed
Electron Micrograph and Defect Identification, North-Holland Publishing Company,
Amsterdam (1973) pages 400. ISBN 0-7204-1757-0

191
ELECTRON MICROSCOPY Microscopy and Analysis

6.15. J.W. Edington, Monographs in Practical Electron Microscopy in Materials Science,


Vol. 2: Practical Electron Microscopy in Materials Science, Macmillan, London
(1975) pages 136. ISBN-13: 978-0333182925

6.16. A. Howie and M.J. Whelan, Diffraction Contrast of Electron Microscope Images of
Crystal Lattice Defects. III. Results and Experimental Confirmation of the Dynamical
Theory of Dislocation Image Contrast, Proceedings of the Royal Society A, 267
(1962) 206–230. doi: 10.1098/rspa.1962.0093

6.17. L.M. Clarebrough, Australian Journal of Physics 24 (1971) 79.

6.18. H. Rose and C. Kisielowski, On the Reciprocity of TEM and STEM, Microscopy and
Microanalysis 11 (2005) 2114–2115. doi: 10.1017/S1431927605507761

6.19. M.J. Whelan, P.B. Hirsch, Electron diffraction from crystals containing stacking faults:
I, Philosophical Magazine 2 (1957) 1121–1142. doi: 10.1080/14786435708242742

6.20. E. Schrödinger, An Undulatory Theory of the Mechanics of Atoms and Molecules, The
Physical Review, 28, (1926), 1049–1070. doi: https://doi.org/10.1103/PhysRev.28.1049

192
ELECTRON MICROSCOPY Microscopy and Analysis

6.21. A. Howie, M.J. Whelan, Diffraction Contrast of Electron Microscope Images of Crystal
Lattice Defects. II. The Development of a Dynamical Theory, Proceedings of the Royal
Society A, 263 (1961) 217–237. doi: 10.1098/rspa.1961.0157

6.22. J.C.H. Spence, Hgh-resolution electron microscopy, Oxford U. Press. New York (2013)
pages 729. ISBN 0-19-505405-9.

6.23. S.J. Pennycook and L A. Boatner, Chemically sensitive structure imaging with a scanning
transmission electron microscope, Nature 336 (1988) 565–567. doi:10.1038/336565a0

6.24. P. D. Nellist and S. J. Pennycook, Direct imaging of the atomic configuration of ultra-
dispersed catalysts, Science 274 (1996) 413–415. doi: 10.1126/science.274.5286.413

6.25. 
Lord Rayleigh, On the theory of optical images with special reference to
the microscope, Philosophical Magazine 42 (1896) 167–195. http://dx.doi.
org/10.1080/14786449608620902

6.26. J.C.H. Spence, Experimental High-Resolution Electron Microscopy, 4th edition, Oxford
University Press, Oxford, UK, (2013). ISBN: 9780199668632

6.27. J.C.H. Spence, J.M. Cowley, Lattice imaging in STEM, Optik 50 (1978) 129-142.

6.28. A. Howie, Image contrast and localised signal selection techniques, Journal of Microscopy
117 (1979) 11–23. doi: 10.1111/j.1365-2818.1979.tb00228.x

6.29. P. Hartel, H. Rose, C. Dinges, Conditions and reasons for incoherent imaging in STEM,
Ultramicroscopy 63 (1996) 93–114 https://doi.org/10.1016/0304-3991(96)00020-4.

6.30. G. Cliff, P.B. Kenway, The effects of spherical aberration in probe forming lenses on
probe size, image resolution and x-ray spatial resolution in scanning transmission electron
microscopy, in Microbeam analysis – 1982, K.F.J. Heinrich (editor) San Francisco
Press, San Francisco, (1982) pages 107–110.

6.31. W. Kossel, G. Möllenstedt, Elektroneninterferenzen im konvergenten Bündel, Annalen


der Physik 36 (1939) 113–140. doi: 10.1002/andp.19394280204

6.32. J.P. Morniroli, J. W. Steeds, Microdiffraction as a tool for crystal structure identification and
determination, Ultramicroscopy 45 (1992) 219–239. http://dx.doi.org/10.1016/0304-
3991(92)90511-H

193
ELECTRON MICROSCOPY Microscopy and Analysis

6.33. D. Mukherji, R.P. Wahi, On the nature of lattice distortion near the g/g’ interfaces in
Ni-base superalloys, Scripta Mater., 36, (1997) 1233–1238. http://dx.doi.org/10.1016/
S1359-6462(97)00033-X

6.34. M. Tanaka, Conventional transmission-electron-microscopy techniques in convergent-


beam electron diffraction, Journal of Electron Microscopy 35 (1986) 314–323 DOI:
https://doi.org/10.1093/oxfordjournals.jmicro.a050584.

6.35. P.M. Kelly, A. Jostens, R.G. Blake, J.G. Napier, The determination of foil thickness by
scanning transmission electron microscopy, Physica Status Solidi (a) 31 (1975) 771–780.
doi: 10.1002/pssa.2210310251

6.36. D.B. Williams, Practical Analytical Electron Microscopy in Materials Science, VCH
Publishing, New Jersey, (1984) pages 153. ISBN-13: 978-0895733078

6.37. J.W. Steeds, Convergent beam electron diffraction, Chapter 15: in Introduction to
Analytical Electron Microscopy, J.J. Hren, J.I. Goldstein and D.C. Joy, (eds.) Plenum
Press, New York, (1979) 387–422. ISBN: 978-1-4757-5583-1

6.38. D. Mukherji, R.P. Wahi, On the measurement of lattice mismatch between g and g’
phases in Ni-base superalloys by CBED technique, Scripta Mater., 35, (1996) 117–122.
http://dx.doi.org/10.1016/1359-6462(96)00092-9

6.39. D.B. Williams, C.B. Carter, Kikuchi Diffraction, Chapter 15: in Transmission Electron
Microscopy, Springer, (2009) 311–322. ISBN: 978-0-387-76500-6

6.40. M.N. Alam, M. Blackman, D.W. Pashley, High-Angle Kikuchi Patterns, Proceedings
of the Royal Society A, 221(1954) 224–242. doi: 10.1098/rspa.1954.0017

6.41. D.C. Joy and G.R. Booker, Simultaneous display of micrograph and selected-area
channelling pattern using the scanning electron microscope, Journal of Physics E:
Scientific Instruments, 4 (1971) 837–842. http://iopscience.iop.org/0022-3735/4/11/001

6.42. D.J. Dingley, On-line determination of crystal orientation and texture determination
in an SEM, Proceedings Royal Microscopy Society, 19 (1984) 74–75.

6.43. J.A. Venables, C.J. Harland, Electron back-scattering patterns –A new technique for
obtaining crystallographic information in the scanning electron microscope, Philosophical
Magazine, 27 (1973) 1193–1200. http://dx.doi.org/10.1080/14786437308225827

194
ELECTRON MICROSCOPY Microscopy and Analysis

6.44. S.I. Wright, B.L. Adams, Automatic-analysis of electron backscatter diffraction patterns,
Metallurgical Transaction A, 23 (1992) 759–767. doi:10.1007/BF02675553

6.45. P.V.C. Hough, Method and means for recognizing complex patterns, U.S. Patent 3069654
A, Dec. 18, (1962).

6.46. N. Stojilovic, Why can’t we see hydrogen in x-ray photoelectron spectroscopy?, Journal
of Chemical Education, 89 (2012) 1331–1332. doi: 10.1021/ed300057j

6.47. R. Castaing, Application of electron probes to local chemical and crystallographic analysis,
(1951) PhD Thesis, University of Paris [English translation by P. Duwez and D.B.
Wittry, California Institute of Technology, (1955)].

6.48. G. Cliff and G.W. Lorimer, The quantitative analysis of thin specimens, Journal of
Microscopy, 103 (1975) 203–207. DOI: 10.1111/j.1365-2818.1975.tb03895.x

6.49. Private communication: all data for the EELS in figures 6.47 to 6.52 are from the
measurements by Dr Partha Goshal, head of the electron microscopy laboratories at
the Defence Metallurgical Research Laboratories, Hyderabad, India.

195
ELECTRON MICROSCOPY Endnotes

ENDNOTES
1. The Ewald sphere is a geometric construction used in electron, neutron, and x-ray crystallography
which demonstrates the relationship between the wavevector of the incident and diffracted beams.
More details will be discussed in Chapter 6 in the book.
2. Dispersion is caused by a spread in the wavelength of the light passing through the lens coupled with
any variation in the refractive index of the glass lens.
3. Meridional and Sagittal planes: The meridional plane (also called the tangential plane) of the lens
represents the meridian of the lens that radiates out from the optical center; these planes are analogous
to the spokes of a bicycle wheel. The sagittal plane of the lens represents the meridian of the lens that is
perpendicular to the tangential plane (i.e., at a 90° angle to it) at any point; these planes circumscribe
the optical center.
4. The Ewald sphere is a geometric construction used in electron, neutron and x-ray diffraction, which
demonstrates the relationship between the incident and diffracted wavevectors and the reciprocal lattice
of the crystal. It is further discussed in Chapter 6.
5. In crystallography, crystal structure gives the description of atomic arrangements in a unit cell and they
are classified according to various symmetry elements. The crystal structure is generally categorized in
different ways and assigned names, numbers or symbols, according to e.g. Bravier Lattices (e.g. cubic,
hexagonal, etc.), Space group (Pm-3m, I4/mmm) and Space Group numbers (# 221, 139), Pearson
symbols (cP4, tI8) or Strukturbericht notations (e.g. L12, DO22 etc.). The examples used here were for
the L12 and the DO22 crystals, i.e. the g‘ and the g‘‘ phases.
6. For electrons passing through a crystalline specimen causing one Bragg reflection to occur (i.e.
under two-beam approximation), when the incident wave reaches a certain depth (x), its amplitude
becomes 0 (zero) and the reflected wave amplitude becomes maximum. Furthermore, if the incident
wave continues to travel the crystal and reach the depth 2x (twice as deep as the previous depth) the
amplitude of the reflected wave becomes 0 again and the amplitude of the incident wave becomes
maximum. Thus, the amplitudes of the incident and the reflected waves oscillate with the depth. The
length of one periodicity of this oscillation is called extinction distance (x). The extinction distance is
inversely proportional to the crystal structure factor of the operating reflection (g) and the wavelength
of the incident electron beam (l).
7. The Burgers vector, named after Dutch physicist Jan Burgers, is a vector, often denoted b, that represents
the magnitude and direction of the lattice distortion resulting from a dislocation in a crystal lattice.
The Burgers vector of the dislocation can be found through the Burgers circuit construction – atom-
by-atom path that forms a closed loop. Firstly, a closed circuit around the dislocation line is made
which encloses the dislocation in the crystal lattice. Then the same chain of base vectors is applied in
a perfect reference lattice. The second loop will not close. The vector needed for closing the circuit in
the reference crystal is the Burgers vector. For more details please see [6.11].

196
ELECTRON MICROSCOPY Endnotes

8. Reciprocity is a wave optical argument first raised by Helmholz in 1860 for optical instruments
stating that ray trajectories can be inverted because of the nature of the stationary wave equation.
The reciprocity theorem applies for the wave function of elastically scattered electrons and not for
inelastically scattered electrons or intensities. The reciprocity states that when coordinates are chosen
along the optic axis, the image and the source points can be exchanged. In particular, the source
plane in the TEM corresponds to the detector plane in the STEM and this equivalence is due to the
Reciprocity Theorem [6.18].
9. A wave equation can explain the behavior of particles like electrons and Schrodinger was the first person
in 1926 to write down such a wave equation [6.14]. The Schrödinger equation is a partial differential
equation that describes how the wavefunction of a physical system evolves over time (he also derived a
Time-independent form of the equation). The Schrödinger equation predicts analytically and precisely
the probability of events or outcome.
10. Each particle is represented by a wavefunction y (position, time) such that y*y = the probability
of finding the particle at that position in that time. Y contains all information about the particle
and establishes the probability distribution in three-dimension. It is used in the formulation of
Schrödinger equation.
11. Howie-Whelan equation gives a method for calculating the intensities of transmitted and diffracted
waves at the bottom plane of a crystalline specimen when the incident electron beam interacts with
the specimen. In the “Howie-Whelan equation”, transmission and diffraction are dealt by dividing the
specimen into many thin layers. The transmitted and diffracted waves are given at the upper surface
of a layer (on the top layer, the diffracted wave amplitudes are assumed to be 0 (zero)). These waves
undergo transmission and diffraction according to the respective crystal structure factors in the layer.
The new transmitted and diffracted waves are obtained at the lower surface of the layer of thickness dz.
These waves are incident on the next layer. This process is repeated for successive planes. Finally, the
amplitudes of the transmitted and diffracted waves at the bottom plane of the specimen are obtained.
This method (equation) can be used to explain the image contrast due to lattice defects such as stacking
faults and dislocations.
12. Contrast Transfer Function (CTF) mathematically describes how lens aberrations in a transmission
electron microscope modify the image of the specimen. It is a mathematical function sin c(q), which
modulates the amplitudes and the phases of the electron diffraction waves formed in the back focal
plane of the objective lens. CTF essentially gives the phase change of diffracted beams with respect to
the direct beam.
13. The Fourier Transform is an important mathematical tool that breaks a waveform into an alternate
representation, characterized by sine and cosines components. Thus, any waveform can be re-written as
the sum of sinusoidal functions using Fourier Transform. The output of the transformation represents
the image in the Fourier or frequency domain, while the input image is the spatial domain equivalent.
In the Fourier domain image, each point represents a particular frequency contained in the spatial
domain image.
14. Scherzer defocus – when a high-resolution structure image of a phase object is taken in the TEM
mode, Scherzer focus is used as the defocus condition. It is determined by the spherical aberration
(Cs) of the objective lens so that the phase of diffracted waves is shifted by 1/4 of wavelength l (or a
phase of p/2) of the electron wave over a wide range of spatial frequencies (q).

197
ELECTRON MICROSCOPY Endnotes

15. Zone Axis – a term sometimes used to refer to “high-symmetry” orientations in a crystal, most generally
refers to any direction referenced to the direct lattice (as distinct from the reciprocal lattice) of a crystal
in three dimensions. Therefore, it is indexed with direct lattice-indices, instead of with Miller-indices.
High-symmetry zone axes through a crystal lattice, in particular, often lie in the direction of tunnels
through the crystal between planes of atoms. This is because, as we see below, such zone axis directions
generally lie within more than one plane of atoms in the crystal.
A zone axis is a lattice row parallel to the intersection of two (or more) families of lattices planes. It is
denoted by [u v w]. A zone axis [u v w] is parallel to a family of lattice planes of Miller indices (hkl)
if: uh + vk + wl = 0.
16. The most accurate method for determining the thickness requires the measurement of the spacing of
the diffuse Kossel-Möllenstedt fringes that occur in a dark-field disc under two-beam conditions [6.35].
The method is described in detail by Williams [6.36].
17. The Hough transform, is a mathematical transformation used to detect lines, circles or other parametric
curves in an image. It was introduced in 1962, by Hough [6.45]. The goal is to find the location of lines in
the images. The simplest case of Hough transform is detecting straight lines. In general, the straight-line y =
m x + b can be represented as a point (b, m) in the parameter space. However, vertical lines pose a problem,
as they would give rise to unbounded values of m, the slope parameter. Thus, for computational reasons,
the line is represented by r = x cos q + y sin q, where r is the distance from the origin to the closest point
on the straight line and q is the angle between the horizontal axis and the line connecting the origin with
that closest point. It is therefore possible to associate a (r, q ) pair values with each line of the image. The
resulting (r, q ) plane corresponding to the line pattern image is sometimes referred to as Hough space.

Join the best at Top master’s programmes


• 3
 3rd place Financial Times worldwide ranking: MSc
the Maastricht University International Business
• 1st place: MSc International Business
School of Business and • 1st place: MSc Financial Economics
• 2nd place: MSc Management of Learning

Economics! • 2nd place: MSc Economics


• 2nd place: MSc Econometrics and Operations Research
• 2nd place: MSc Global Supply Chain Management and
Change
Sources: Keuzegids Master ranking 2013; Elsevier ‘Beste Studies’ ranking 2012;
Financial Times Global Masters in Management ranking 2012

Maastricht
University is
the best specialist
university in the
Visit us and find out why we are the best! Netherlands
(Elsevier)
Master’s Open Day: 22 February 2014

www.mastersopenday.nl

198
ELECTRON MICROSCOPY About the author

ABOUT THE AUTHOR


I started my research career as a Junior Scientific Officer at the Defence Metallurgical Research
Laboratory, Hyderabad, India after my graduation in Metallurgical Engineering from the
Indian Institute of Technology (IIT), Kharagpur, India. Transmission electron microscope
was one of the major tools for my research on high temperature Titanium alloys and I
started learning to use the instrument from the day one of my research career but before
that, I only had a vague idea what a TEM is. TEM and Titanium alloy became the main
subject for my doctoral thesis. I got my PhD from IIT, Kharagpur in 1987. Shortly after
completing my PhD, I moved to Berlin, Germany and worked first as a postdoc fellow and
later as a research scientist at the Hahn Meitner Institut and the Technische Universität
Berlin. The material for research changed to Ni-base superalloys, but the major research
tool remained the same – TEM.

Since then, I have moved around a bit, working at Swiss Federal Institute of Technology
Zurich (ETH), Zurich, Switzerland and the Technische Universität Braunschweig (TU-
BS), Germany. I also expanded my characterization methods to include techniques beyond
electron microscopy to include neutron scattering methods, but the research topic mainly
was high temperature alloys for gas turbines. In between, some research on high temperature
Ni-superalloys, lead to a patent and some years of research (mainly at TU- BS and ETH
Zurich) on nano-particles / nanotechnology, where again electron microscopy (both SEM
and TEM) was an invaluable tool. My present research interest is high temperature alloy
development, beyond nickel based superalloys for future gas turbines.

199
ELECTRON MICROSCOPY About the reviewer: Prof. Dr. Gerhard Schumacher

ABOUT THE REVIEWER:


PROF. DR. GERHARD SCHUMACHER
I completed my Diploma and my PhD at Hahn-Meitner Institut Berlin in the area of
Radiation effects in metallic glasses. After my PhD I worked two more years at the Hahn-
Meitner Institute in the same research field and then moved to Physkalisch-Technische
Bundesanstalt and Technische Universität München before I came back to the Hahn-Meitner
Institute for a few years. Then I changed to Argonne National Laboratory as a Max-Kade
fellow where I continued my studies on radiation effects in metallic glasses with Brioullin-
Scattering and dilatation measurements.

Coming back to Germany, I worked at Technische Universität Berlin and Bundesanstalt


für Materialforschung in the area of high temperature alloys before I reached my final
termination Helmholtz-Zentrum Berlin, the former Hahn-Meitner Institute.

In 2004, I received the postdoctoral lecture qualification at Technische Universität Berlin


where I got a professorship in 2015.

Currently, I work as a group leader and deputy leader of the department of microstructure
and residual stress analysis at Helmholtz-Zentrum Berlin in the research fields of battery
materials, membranes, high temperature alloys, metallic glasses and high entropy alloys.
For characterization of these materials we use mainly X-ray and neutron diffraction, X-ray
absorption spectroscopy and electron microscopy.

200
ELECTRON MICROSCOPY About the reviewer: Dr Partha Ghosal

ABOUT THE REVIEWER:


DR PARTHA GHOSAL
After my graduation in Solid State Physics from Banaras Hindu University, I received my
PhD from IIT, Banaras Hindu University in 1996 in the field of Metallurgical Engineering
and worked on the effect of crystal defects in the diffraction theory for ordered hcp alloys.
I worked extensively on Ti alloy development for aerospace application using electron
microscopy. I currently head the Electron Microscopy Group in DMRL, Hyderabad in the
area of advanced characterization techniques, along with SEM, EBSD, FIB, TEM, HRTEM
and Raman-AFM systems. My current interests include in-situ mechanical testing and in-
situ heating experiments of nano and advanced materials inside electron microscopes and
advanced Raman analysis for these materials.

In 2001-2002, I worked at Technical University, Clausthal, Germany as DAAD visiting


scientist on orientation imaging microscopy using SEM and TEM on electronic materials.

Extensive electron microscopic works on Ti base alloys, W based alloys and nano composites
are also in the scope of my interests. I have published around 98 international papers, have
been invited to over 100 presentations and have 3 patents to my credit with 26 years of
research experience. Recently, I was elected fellow of EMSI and International member of
CEC (IUCr) from India. I am also the reviewer of 7 international journals and a life member
of professional national bodies such as IIM, MRSI, EMSI and MSI, in India.

201
ELECTRON MICROSCOPY review by dr partha ghosal

REVIEW BY DR PARTHA GHOSAL


The back bones of any material development is characterization and among this, electron
microscopy has revolutionized our understanding. Along with advanced research, one of
the important parts is to update the knowledge of the next generation of graduates and
researchers of physical and material sciences. There existed a requirement for a book that is
useful and referred to by all young researchers to start doing electron microscopy of a variety
of materials. It is very satisfying to read the book and I am very pleased to note that the
instrument and technique details are elaborated with beautiful photographs and references
that students will be very excited to see. The beauty of the book lies in the simplicity of
language used for explanations, the photographs, minimum mathematical equations and
sufficient references. This will make students more curious about the subject and start
working on it. From electron microscopes to advanced techniques, most of the areas are
covered for any researcher to use. It is even more important to note that undergraduates
are now expected to know the basics, applications and techniques of electron microscopes,
which can be utilized to understand the science and engineering of various materials which
will be part of tomorrow’s world. The topics in the chapters precisely presented with color
photographs. The author has undergone the great task of covering so much in such a
short book, including an interesting introduction, operational knowledge and concepts of
various techniques.

202

You might also like