Electron Backscatter Diffraction Strategies For Reliable Data Acquisition and Processing

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Tutorial review

Electron backscatter diffraction: Strategies for reliable data


acquisition and processing
Valerie Randle

Materials Research Centre, School of Engineering, Swansea University, Swansea SA2 8PP, UK
A R T I C L E D A T A A B S T R A C T
Article history:
Received 18 March 2009
Received in revised form18 May 2009
Accepted 20 May 2009
In electron backscatter diffraction (EBSD) software packages there are many user choices
both in data acquisition and in data processing and display. In order to extract maximum
scientific value from an inquiry, it is helpful to have some guidelines for best practice in
conducting an EBSDinvestigation. The purpose of this article therefore is to address selected
topics of EBSD practice, in a tutorial manner. The topics covered are a brief summary on the
principles of EBSD, specimen preparation, calibration of an EBSD system, experiment
design, speed of data acquisition, data clean-up, microstructure characterisation (including
grain size) and grain boundary characterisation. This list is not meant to cover exhaustively
all areas where EBSD is used, but rather to provide a resource consisting of some useful
strategies for novice EBSD users.
2009 Elsevier Inc. All rights reserved.
Keywords:
Electron backscatter diffraction
Microtexture
Orientation
1. Introduction
Electron backscatter diffraction (EBSD) is a scanning electron
microscope (SEM) based technique which has become well
known as a powerful and versatile experimental tool for
materials scientists, physicists, geologists and other scientists
and engineers. It allows the measurement of microtexture
(texture on the scale of the microstructure) [1], microstructure
quantification [2], grain and phase boundary characterisation
[3,4], phase identification [5] and strain determination [6] in
crystalline multiphase materials of any crystal structure. It
has now been more than twenty years since the inception of
EBSD as an add-on facility to the capabilities of an SEM.
Following commercialisation of the product at an early stage
EBSD has attracted increasing interest in SEM user commu-
nities, and several key publications have chronicled its
development [711]. Nowadays, EBSD systems exist world-
wide. There are two main manufacturers plus a few smaller
specialist companies.
Fig. 1 illustrates the growth in publications relating to EBSD
in the present decade. There is an exponential increase,
showing particularly a take-off in papers since 2005 as EBSD
gained in popularity and accessibility. The papers quoted here
are those obtained from a single database source which is
likely to reflect the most significant papers in the materials
field. The absolute number of publications relating to EBSD
over this time period will be therefore higher than the
numbers quoted in Fig. 1. Further analysis of recent trends
in EBSD application has shown that, although microtexture
determination used to be the primary application of the
technique, nowadays it is used for a wide range of sophisti-
cated microstructure characterisation, sometimes in conjunc-
tion with other analyses such as finite element modelling.
Furthermore originally EBSD was applied exclusively to
materials having cubic symmetry, whereas now increasingly
it is applied to materials with more complex structures.
Hexagonal materials and geological materials are examples
of such growth areas [10].
M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
Tel.: +44 1792 295841; fax: +44 1792 295676.
E-mail address: v.randle@swansea.ac.uk.
avai l abl e at www. sci encedi r ect . com
1044-5803/$ see front matter 2009 Elsevier Inc. All rights reserved.
doi:10.1016/j.matchar.2009.05.011
Operation of an EBSD system is no more difficult than
operating an SEM. However, in the comprehensive EBSD soft-
waretherearemanyuser choices bothindataacquisitionandin
data processing and display. In order to extract maximum
scientific value from an inquiry, it is helpful to have some
guidelines for best practice inconducting anEBSDinvestigation.
The purpose of this article therefore is toaddress selectedtopics
of EBSD practice in a tutorial manner. The topics covered are a
brief summary onthe principles of EBSD, specimenpreparation,
calibrationof anEBSDsystem, experiment design, speedof data
acquisition, data clean-up, microstructure characterisation
(including grain size) and grain boundary characterisation.
This list is not meant to cover exhaustively all areas where
EBSD is used, but rather to provide some useful strategies for
novice EBSD users. Throughout the text some golden rules are
mentioned. These are printed in italics for emphasis.
2. Principles of EBSD
EBSD is based on acquisition and analysis of Kikuchi diffrac-
tion patterns from the surface of a specimen in an SEM.
Detailed accounts of EBSD hardware, software and the
generation and indexing of diffraction patterns can be found
in several texts [9,12], and will not be repeated here. To obtain
an EBSD diffraction pattern a stationary beam of electrons is
sited on the specimen surface. Backscattered electrons are
diffracted at crystal lattice planes within the probe volume,
according to Bragg's law. The fraction of diffracted back-
scattered electrons which are able to escape from the speci-
men surface is maximised by tilting the specimen so that it
makes a small angle, typically 20, with the incoming electron
beam. The diffraction patterns arise therefore from typically
up to 50 nm depth from the specimen surface. The depth
resolution has been quoted as approximately 40 nmfor silicon
and 10 nm for nickel [1315]. The diffracted signal is collected
and viewed via a low-light video camera interfaced to a
phosphor screen.
Camera technology has improved greatly in the last few
years, which has underpinned vastly increased data acquisi-
tion speeds. The improved quality of the captured diffraction
pattern is due in part to an improved dynamic range of the
camera, i.e. the number of distinguishable grey levels, which
has increased tenfold by use of a Charge-Coupled Device (CCD)
camera compared to the older Silicon Intensified Target (SIT)
camera. The improvement is also due to the recent advance
whereby groups of pixels in the diffraction pattern can be
grouped together into super pixels (binning), which has the
effect of increasing the sensitivity of the camera. For example
a block of 88 pixels can be grouped together to produce the
same increase in camera sensitivity, i.e. 88=64. In turn, this
gives the same reduction in diffraction patterncollection time.
The gain in pattern collection speed brought about by binning
is further enhanced by recent improvements in electronic
processing, such as frame averaging, and amplification of the
captured diffraction pattern as well as computer and software
improvements. This has led to faster and faster mapping
rates, depending on the material (Section 3.4). The latest
generation of cameras also has the advantage of distortion-
free lenses and a rectangular phosphor screen, to replace the
previous circular screen, so that the whole diffraction pattern
is captured and used. This camera is also shaped so that it can
be moved close to the specimen, which increases both the
camera sensitivity and the spatial resolution.
The diffraction patterns provide crystallographic informa-
tion that can be related back to their position of origin on the
specimen. Fig. 2a shows a typical processed EBSD diffraction
pattern from an austenitic (face centred cubic) steel. Evalua-
tion and indexing of the diffraction patterns is performed in
most cases automatically and the data is output in a variety of
Fig. 1 Number of EBSD-related publications in the period
20002008 (Source: ScienceDirect).
Fig. 2 (a) Typical processed EBSD diffraction pattern froman
austenitic (face centred cubic) steel. (b) Diffraction pattern
from (a) with simulated bands added, as part of the
calibration routine.
914 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
both statistical and pictorial formats. The most versatile and
revealing of these outputs is the orientation map (sometimes
called an Orientation Imaging Micrograph, OIM), which is a
quantitative depiction of an area of microstructure in terms of
its crystallographic constituents. Orientation maps are dis-
cussed further in Section 4.
The spatial resolution of EBSD is influenced by the
microscope parameters, the atomic number of the material,
the specimen/microscope geometry, the accelerating voltage
and probe current used and the diffraction pattern quality.
These factors are discussed in some detail elsewhere
[9,11,12]. The best spatial resolution for a material such as
brass, for example, is 2550 nm for a tungsten filament SEM
and 922 nm for a FEGSEM, both measured parallel to the tilt
axis [1618].
The angular resolution or accuracy of EBSD relates directly
to the precision with which the diffraction pattern can be
indexed, which in turn is influenced by the calibration of the
system (Section 3.2), the effectiveness of the software solve
routine, the diffraction pattern quality (Section 4.2) and the
magnification of the diffraction pattern. The accuracy for
orientation measurement by EBSDhas been found experimen-
tally to be approximately 0.51.0 in a tungsten filament SEM
[19]. This precisionapplies for patterns of optimumquality and
will reduce significantly for poor quality patterns. The speed of
data collection has increased greatly in recent years, which is
discussed in Section 3.4.
3. Data Acquisition
3.1. Specimen Preparation
As mentionedinSection2, EBSDdiffractionpatterns arise from
the topfewnanometers of specimensurface, andsodiffraction
patternquality is very sensitive to crystalline perfection in this
surface layer. Any occlusion or deformation must be removed
from the surface by appropriate specimen preparation. Most
conducting materials are prepared by the normal metallo-
graphic routes of grinding and polishing to produce a flat
surface. Diamond polishing is not a suitable final step in some
materials because of the surface deformation introduced.
Electropolishing, etching and polishing with colloidal silica
are all used as final preparation steps. Ion milling, either in a
Dual-Beam instrument or ex-situ, or plasma etching can be
used for materials not amenable to conventional metallogra-
phy [20]. Non-conducting specimens can be thinly coated with
a conducting medium, as for conventional SEM examination,
and examined with a higher accelerating voltage to compen-
sate for the extra thickness.
Optimum specimen preparation is a fundamental requirement for
anEBSDexperiment. Inadequate specimenpreparationwill result
in degraded diffraction patterns which feed through to loss of
data quality. Data clean-up routines can compensate in part for
unsolved patterns (Section 4.1) but clean-up procedures should
never be selected as an alternative to obtaining the best and most
representative diffraction patterns. Furthermore real plastic strain,
whichresults inreduceddiffractionpatternquality, mayexist in
the lattice [21]. It is important to distinguish this real and
quantifiable effect from preparation-induced artefacts.
3.2. Calibration
The goal of calibrating an EBSD system is to establish the
geometry of the projection of the EBSD pattern onto the
phosphor screen and to determine the geometrical relationship
between the specimen coordinates inthe SEMchamber and the
phosphor screen. The physical factors which can vary and
therefore alter thecalibrationparameters are thetilt angleof the
specimen surface with respect to the electron beam, the
position of the phosphor/camera assembly and the position of
the specimen. Calibration is described in detail elsewhere [9,11]
and in manufacturers' documentation. Two modes of calibra-
tion can be classified: a full calibration which usually involves a
calibrationcrystal of knownorientation, andaniterativepattern
fitting routine which refines the calibration parameters and is
performed on a good quality pattern from a known material.
The former mode is used in the initial commissioning of an
EBSDsystemandthereafter onlyapproximatelyonce per year or
morefrequentlyif thephysical set-uphas beendisturbed. Fig. 2b
shows an example of a diffraction pattern from an austenitic
steel, where the system has been well calibrated and so the
simulated Kikuchi bands are an excellent match to the real
Kikuchi bands inthe EBSDpattern. Accurate calibration is essential
to obtaining reliable data. With practice a refined calibration
routine takes only minutes and should be performed at the
beginning of every investigation.
3.3. Experiment Design
The first step in an EBSD investigation is deciding what
information is required. (Here we will assume that the
investigation involves a specimen of known phases. Other-
wise, EBSD can be used to identify the constituent phases and
this is described elsewhere [22].) The investigation is usually
combinations of microtexture, interface parameters and
microstructure. The requirements of the investigation will
determine the sampling schedule, that is, the location of the
data points and how many are sampled [11].
For an unknown specimen some preliminary trial-and-error
exploration by the user will establish the scale of the orienta-
tions and howthey relate to the microstructure, whichwill then
allow the user to select appropriate areas on the specimen.
Examples of such areas include adjacent to a specimen edge, in
the vicinity of defects such as cracks, or to include a particular
phaseor graindistribution. Oncetheparameters for thelocation
of theregionof interest (i.e. selectedareaplus thestepsize) have
been determined by trial runs, most investigations can then
proceed automatically by orientation mapping.
Most EBSDdata is collectedvia anorientationmap, for which
the step size or grid size must be specified. Astep size of one-tenth
of the average grain size is a good starting point for general
microtexture, microstructure and grain misorientation measurements.
Asmaller step size than this or a smart sampling option would
be needed if the grain size distribution is inhomogeneous (e.g. a
partially-recrystallised microstructure). Smart sampling, which
is designed to eliminate the high extent of oversampling within
grains that is a drawback of the standard orientation mapping
approach, is an option in some commercially available EBSD
software. The smart sampling option is best used on unde-
formedgrains, and could be usedfor example for the analysis of
915 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
samples with a highly bimodal grain size distribution, so that
small grains can be mapped without extensive oversampling of
the larger grains. The procedure begins by analysing a coarse
grid covering the whole area. Then the algorithm calculates
whether it needs to sample between any two of the original
points (i.e. whether there is a grain boundary between the two
points). The procedure is iterated so that the sampling ends up
concentrated at grain boundaries.
It is especially important to capture all small grains if the
grain size is to be determined from the maps (Section 4.2).
Fig. 3 shows an example of a large grain size spread caused by
secondary recrystallisation in a silicon iron sample. A small
step size, relative to the grain size, is also required for high
resolution definition of the interface network, e.g. for bound-
ary reconstruction (Section 4.3).
Selection of appropriate sample areas and an appropriate step
size is critical to the investigation. Unsuitable selections cannot be
changed retrospectively. The high throughput speeds avail-
able with the latest generation EBSD cameras and systems
(Section 3.4) facilitate data experiment design in that even if
the smallest step sizes and large sampling areas are chosen
mapping can be accomplished in reasonable time frames.
A note needs to be added on the significance of micro-
texture and macrotexture measurements using EBSD. Micro-
texture is defined as a sample population of orientation
measurements which can be linked individually to their
location within a specimen [11]. It refers essentially to the
orientations in the sampled area only, and does not imply the
texture of the entire specimen. Onthe other hand macrotexture
is defined as an average texture determined from many grains
obtained without necessarily having reference to the location of
individual grains within a specimen [11]. In the past macro-
texture (referred to simply as texture) could only be measured
using X-ray or neutron diffraction, in order to obtain sufficient
grain sampling. Nowadays the increased speed of data acquisi-
tion means that it is viable to obtain macrotexture by EBSD, but only
if due attention is paid to the sampling schedule and quantity of data.
The statistical reliability of texture measurements by EBSDhave
been compared to those obtained by X-rays in steel specimens
[23]. The results from the two techniques were comparable.
Approximately 10,000 grains were found to produce a very good
sampling in these materials. It should be emphasised that this
data requirement refers to numbers of grains, and not simply
data points. Prior knowledge of the grain size is therefore
needed. Texture measurements via EBSD have many advan-
tages over X-ray methods, for example differentiating textures
in large and small grains.
Insummary although there is a great diversity inapplication
of EBSD for microtexture-related investigations, the basic steps
of experiment design and procedure are generic. These are
1. Selection of a suitable candidate material and specimens
for analysis
2. Optimum specimen preparation
3. EBSD system calibration
4. Manual checking of diffraction patterns for clarity and
expected phase match
5. Appraisal of microstructure for grain size and phase distri-
bution, etc
6. Selection of sampling area(s) on specimen
7. Test runof datacollection; checkappropriateness of stepsize
8. Data collection, usually via automated orientation mapping.
3.4. Speed of Data Acquisition
In recent years the speed of EBSD measurements has
increased enormously. This is due to improvements in camera
technology (especially the introduction of digital cameras),
Fig. 3 Orientation map illustrating a large grain size spread, in silicon iron. Small, mediumand large sized grains are coloured
green, yellow and red respectively. (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)
916 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
Fig. 4 Sequence of orientation maps from alpha-brass, illustrating the effect of data clean-up. The maps are shaded grey
according to a pattern quality parameter, where darker shading corresponds to poorer quality. Unsolved pixels are black.
(a) Raw map with no clean-up, containing 5%zero solutions. (b) Part cleaned map containing 2%zero solutions. (c) Part cleaned
map containing 0.5% zero solutions.
917 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
changes in the interface between the camera and the
computer and upgrades in the EBSD pattern solve software.
All these features are described in detail elsewhere [12,24].
Fast EBSD has widened the scope of applications. SEM time is
used more efficiently because EBSD scans are much faster,
and statistically-significant quantities of data are easily
obtained so that for example grain size and macrotexture
measurements are facilitated. Furthermore dynamic experi-
ments such as in-situ tensile testing and hot stage experiments
[25] would not be possible without fast EBSD.
At the time of writing the fastest pattern acquisition speed
quoted is 750 points per second [26], on suitable specimens.
EBSD manufacturers quote indexing rate of 400750 patterns
per second. A proviso is that error limits may have to be
widened in order to achieve such speeds, which feeds through
to lower reliability. Whether or not this can be tolerated
depends on the nature of the investigation. For example, fast
mapping would probably be appropriate in order to gain an
impression of the microtexture whereas to obtain grain
boundary parameters often the highest accuracy is the
paramount concern rather than speed. It is important to ensure
the mapping speed is appropriate to the inquiry.
4. Data Processing
4.1. Clean-up
Standard clean-up or noise reduction options in commercial
EBSD packages allow the user to remove both points where
indexing was not possible and also isolated points that have
been incorrectly indexed. The points that are removed are
fixed by filling in using copies of neighbouring points.
Incorrect indexing could occur because of incorrectly set
indexing parameters or misindexing due to pseudosymmetry
in diffraction patterns. Zero solutions (indexing not possible)
could occur at grain boundaries (because of double diffraction
pattern sampling or etching), or if no phase match is available,
or if there is too much deformation in the specimen or if the
specimen surface is blemished or occluded.
Fig. 4 shows a sequence of orientation maps from alpha-
brass which illustrates the effects of diffraction pattern quality,
indexing and data clean-up. The maps are shaded grey
according to diffraction pattern quality, with light grey shading
depicting the best pattern quality (i.e. the sharpest Kikuchi lines
in the diffraction pattern) and conversely dark grey shading
depicting poor pattern quality (i.e. diffuse Kikuchi lines). Zero
solutions are black. Pattern quality is always more diffuse at a
defect than within the lattice, and so grain boundaries are
delineated with darker shading than within grains. Fig. 4a
shows the as-collected map, which contains 5% zero solutions.
These unsolved pixels occur as isolated pixels within grains, at
grain boundaries, at minor surface scratches and extensively
within just a few grains. Fig. 4b and c illustrates the effect of
clean-up to leave 2% and 0.5% unsolved pixels respectively.
Evidence of scratches remains as lines of poorer patternquality.
This series of maps illustrates the acceptable use of noise
reduction routines which retain the integrity of the data.
Excessive clean-up can distort the data and therefore clean-up
routines must always be used with caution.
The map in Fig. 5, which is from an austenitic steel,
consists of the pattern quality parameter and has not under-
gone any clean-up. Only 0.6% of pixels are zero solutions, and
these are virtually all sited at random grain boundaries. This
map illustrates that in this instance specimen preparation has
been carried out to the highest standard and there is virtually
no clean-up required, except at grain boundaries where some
unsolved pixels are inevitable. It is also interesting to note that
Fig. 5 Uncleaned pattern quality map froman austenitic steel. Randomboundaries, annealing twins, other 3 boundaries, 9
boundaries and 27 boundaries are black, white, blue and yellow respectively. Zero solutions are sited mainly at random
boundaries. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)
918 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
unsolved pixels are sited at random grain boundaries but not
at annealing twins [27].
4.2. Microstructure Characterisation
By far the greatest use of EBSD is to depict either micro-
structure or microtexture, or frequently both. The nature of
orientation mapping means that quantitative microstructural
descriptors are generated in addition to the general portrayal
of the microstructure. Whereas use of EBSD for texture
determination is still a widely used area [11], a higher fraction
of reported EBSD investigations related to general microstruc-
ture characterisation than microtexture in 2006 than in 2003
[10]. This trend is continuing. Strategies for the application of
EBSD to texture determination will not be expounded here,
although they are discussed elsewhere [11].
The pattern quality map is often the best choice for the depiction of
microstructure. The mainadvantage of the patternquality mapis
that the map is derived solely from contrast in the diffraction
pattern, i.e. its diffuseness, rather than any indexing para-
meters. As seen in Figs. 4 and 5, grain boundaries are rendered
visible and hence the grain structure is portrayed. There are
various patternquality metrics that canbe employed, whichare
discussed elsewhere [28]. Asecond advantage of pattern quality
maps is that the viewing of different phases may be facilitated,
as illustrated in Fig. 6. Fig. 6a is pattern quality map from an
Fig. 6 Orientation maps from an electroplated nickel composite coating co-deposited with titanium oxide (rutile) particles.
(a) Pattern quality map. The rutile particles are large and recognised by their pattern quality parameter. (b) Inverse pole figure
(normal direction) map of the nickel phase from the same area as in (a). The rutile phase is coloured green, and high and low
angle boundaries are coloured black and grey respectively (maps courtesy of Lampke and Dietrich). (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)
919 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
electroplated nickel composite coating co-deposited with tita-
niumoxide (rutile) particles [29]. Herethe grainstructure of both
phases is very clearly visible. Fig. 6b is an inverse pole figure
(normal direction) map of the nickel phase from the same area
as in Fig. 6a. The rutile phase is coloured green.
Pattern quality maps are particularly useful for the discri-
mination of phases that are difficult to distinguish by other
means. For example pattern quality maps have been used
successfully to determine the volume fraction of the micro-
structural constituents bainite, ferrite and austenite, in a TRIP
steel [30]. Similarly, volume fractions of recrystallised and
unrecrystallised microstructure can be determined. This task
has been carried out automatically in aluminium [31].
Increasingly EBSD is the tool of choice for the determina-
tion of grain size distribution and furthermore gives access to
accompanying derivative analyses such as grain axis ratio and
microtexture according to grain size class. For example, grain
size and microtexture distributions have been determined for
nanostructured electrodeposited NiCo at both the deposit
bath interface and at the depositsubstrate interface [2]. It is
important for grain size determination that small grains are
recognised unambiguously in maps, and that an appropriate choice
is made for the lower limit of grain boundary misorientation to define
a grain. The topics of map step size selection (Section 3.3) and
map clean-up (Section 4.1) have direct bearing on accurate
grain size determination. These topics are discussed and
illustrated in detail elsewhere [32].
Fig. 7a shows a simple example of grain size determination
in nickel. Grains are displayed in random colours and
interfaces, including annealing twins, are black lines. This
example illustrates the point, from observation of the grain
colours, that in this case annealing twins have been ignored
for grain size determination. The grain size statistics are
shown in Fig. 7b. Grains with two sides or less have been
omitted from the sample population. The average grain size is
141 m.
Fig. 7 (a) Orientation map from annealed nickel wherein grains are depicted in random colours and all interfaces are black.
(b) Grain size histogram from the map in (a).
920 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
4.3. Grain Boundary Characterisation
For many years data fromEBSDorientationmaps has beenused
to characterise grain boundary misorientation distributions,
especially in cubic materials. Typically grain boundaries are
classified according to low/high angle type and often by the
popular coincidence site lattice (CSL) notation. Such analyses
provide partial information on the possible properties of the
boundary [33].
Proportions of grain boundary types can be calculated as a
fractionof either total grainboundarylengthor total numbers of
grainboundaries. Thelatter statistic requires that all grains (and
hence grain boundaries) have been explicitly detected, usually
as part of the grainsize determination routine. The number and
lengthfraction statistics derived fromgrain boundary classes in
the map in Fig. 7a are 30.4% length fraction 3, 20.8% number
fraction 3, 0.9%length fraction 9, 1.7%number fraction 9. It
is clear that there are differences between the length and
number statistics, whichhas beendiscussed indetail elsewhere
[34]. This fact is not always acknowledged in publications. It is
therefore vital that when quoting grain boundary proportions, it
should be clearly stated if they refer to a number or a length fraction.
Like other areas of EBSD application, in recent years the
area of interface studies has evolved. Examples of this more
sophisticated usage include application to grain boundary
networks rather than individual boundaries [35], phase
boundaries and orientation relationships [4] non-cubic mate-
rials [36] and measurement of the grain boundary plane in
addition to the misorientation [3,37]. The distribution of grain
boundary planes can be measured using the so-called five-
parameter analysis, which is described in detail elsewhere
[38,39]. The boundary plane distribution is derived either by a
statistical method or by serial sectioning.
The first step in determination of the boundary plane
distribution is to apply a grain boundary reconstruction
algorithm to reproduce faithfully the outline of grain bound-
ary segments. Fig. 8 shows an example from lightly deformed
rocksalt [40]. Fig. 8a is an orientation map from the salt
specimen. Low angle boundaries are red lines. Fig. 8b shows
the corresponding reconstructed microstructure wherein
each grain is assigned a single average orientation value
and represented as a random colour. Grain boundary line
segments used in the analysis are shown as black lines. It is
clear that the reconstructed boundary network is a good
match to the real boundary network. Fig. 8c shows, in stereo-
graphic projection, the overall grain boundary plane distribu-
tion recorded from this specimen. The density units are
multiples of a random distribution, MRD. It can be seen that
there are well-defined maxima at {100} planes. The square-
shaped grains seen in Fig. 8a and b (one is depicted with a
cross) also have been shown by manual measurements to
have planes close to {100} [41].
The five-parameter analysis and analysis of boundary
planes is a relatively new extension of EBSD characterisation,
and already is extending knowledge of the relationship
between grain boundary crystallography and properties [3].
The recent development of a Dual-BeamEBSDand focused ion
beam microscope (FIB) allows in-situ serial sectioning of small
regions hence facilitating the characterisation of grain bound-
ary planes and other three-dimensional aspects of micro-
structure characterisation [42].
5. Concluding Remarks
EBSD has matured into a powerful experimental tool, encom-
passing many strands of microstructure characterisation. This
paper has sought bothto illustrate the richness of data available
from modern EBSD analysis and to provide, via selected
examples, a resource for novice users on how to get the best
out of an EBSD inquiry.
Fig. 8 (a) Pattern quality orientation map from lightly
deformed rocksalt. Low angle boundaries are red lines. One
of several square-shaped grains is depicted with a cross.
(b) Reconstructed microstructure corresponding to (a). Each
grain is assigned a single average orientation value and
represented as a random colour. Grain boundary line
segments used in the analysis are shown as black lines.
(c) Density distribution of grain boundary planes in the
rocksalt specimen expressed as multiples of a random
distribution, MRD in stereographic projection. (For
interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
921 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2
R E F E R E N C E S
[1] Jiang J, Godfrey A, Liu W, Liu Q. Microtexture evolution via
deformation twinning and slip during compression of
magnesium alloy AZ31. Mater Sci Eng 2008:483484:576579.
[2] Bastos A, Zaefferer S, Raabe D, Schuh C. Characterization of
the microstructure and texture of nanostructured
electrodeposited NiCo using electron backscatter diffraction
(EBSD). Acta Mater 2006;54:245162.
[3] Randle V, Rohrer G, Miller H, Coleman M, Owen G.
Five-parameter grain boundary distribution of commercially
grain boundary engineered nickel and copper. Acta Mater
2008;56:236373.
[4] Gourgues-Lorenzon AF. Application of electron backscatter
diffraction to the study of phase transformations. Int Mater
Rev 2007;57:65128.
[5] Perez M, Kenik E, O Keefe M, Miller F, Johnson B. Identification
of phases in zinc alloy powders using electron backscatter
diffraction. Mater Sci Eng 2006;424A:23950.
[6] Wilkinson A, Meaden G, Dingley D. High-resolution elastic
strain measurement from electron backscatter diffraction
patterns: new levels of sensitivity. Ultramicros
2006;106:30713.
[7] Randle V. Microtexture determination and its applications.
Second edition. London: Institute of Materials; 2003.
[8] Schwartz AJ, Kumar M, Adams BL, editors. Electron
backscatter diffraction in materials science. New York:
Kluwer Academic; 2000.
[9] Schwartz AJ, Kumar M, Adams BL, Field DP (Editors). Electron
backscatter diffraction in materials science, Second Edition,
Springer, New York, in press.
[10] Randle V. Recent developments in electron backscatter
diffraction. Adv Imaging Electron Phys 2008;151:363416.
[11] Randle V, Engler O. Introduction to texture analysis:
macrotexture, microtexture and orientation mapping. Taylor
and Francis, USA, second edition, in press.
[12] Schwarzer R, Field D, Adams B, Kumar M, Schwartz A. Present
state of electron backscatter diffraction and prospective
developments. InElectronbackscatter diffractionin materials
science, Second Edition. Schwartz AJ, Kumar M, Adams BL,
Field DP (Editors), Springer, New York, in press.
[13] Venables J, Bin-Jaya R. Accurate microcrystallography using
electron back-scattering patterns. Phil Mag 1977;35A:131732.
[14] Dingley DJ. Progressive steps in the development of electron
backscatter diffraction and orientation imaging microscopy.
J Micros 2004;213:21424.
[15] Michael J, Goehner R. Advances in backscattered-electron
Kikuchi patterns for crystallographic phase identification.
In: Bailey G, Garratt-Reed A, editors. Proc. 52nd annual
meeting of the Microscopy Society of America. San Francisco
Press Inc.; 1994. p. 5967.
[16] Venables J, Harland C. Electron backscattering pattern a
new technique for obtaining crystallographic information in
the scanning electron microscope. Phil Mag 1973;27:1193.
[17] Dingley D. In: Schwartz AJ, Kumar M, Adams BL, editors.
Electron backscatter diffraction in materials science.
New York: Kluwer Academic press; 2000. p. 1.
[18] Humphreys FJ. Review grain and subgrain characterization
by electron backscatter diffraction. J Mater Sci 2001;36:3833.
[19] El-Dasher B, Adams B, Rollett A. Viewpoint: experimental
recovery of geometrically necessary dislocation density in
polycrystals. Scripta Mater 2003;48:141.
[20] Electron backscattered diffraction explained. Oxford
instruments analyticaltechnical briefing. http://www.ebsd.
com/.
[21] Wilkinson A, Dingley D,Meadon G. Strain mapping using
electron backscatter diffraction. In Electron backscatter
diffraction in materials science, Second Edition, Editors
Schwartz AJ, Kumar M, Adams BL, Field DP, Schwartz., in press.
[22] El-Dasher B, Deal A. Application of electron backscatter
diffraction to phase identification. In Electron backscatter
diffraction in materials science, Second Edition, Editors
Schwartz AJ, Kumar M, Adams BL, Field DP. Springer,
New York, in press.
[23] Wright S, Nowell M. A comparison of texture measurements
via EBSD and XRay. ICOTOM15, ed. A.D. Rollett, American
Ceramic Soc. and TMS (on CD). ISBN 978-1-57498-296-1.
[24] Schwarzer R. A fast ACOM/EBSD system Arch. Metall Mater
2008;53:5.
[25] Wright S, Nowell M. A review of in situ EBSD studies.
In Electron backscatter diffraction in materials science,
Second Edition, Editors Schwartz AJ, Kumar M, Adams BL,
Field DP. Springer, New York in press.
[26] Hjelen J, rsund E, Hoel E, Runde P, Furu T, Nes E. EBSP,
progress in technique and applications. Textures and
Microstruct 1993;20:2940.
[27] Jones R, Owen G, Randle V. Carbide precipitation and grain
boundary plane selection in overaged type 316 austenitic
stainless steel. Mater Sci Eng 2008;A496:25661.
[28] Wright S, Nowell M. Image quality mapping. Microsc
Microanal 2006;12:7284.
[29] Lampke T, Dietrich D, Leopold A, Alisch G, Wielage B.
Cavitation erosion of electroplated nickel composite coatings.
Surf Coat Technol 2008;202:396774.
[30] Petrov R, Kestens L, WasilkowskaA, Houbaert Y. Microstructure
and texture of a lightly deformed TRIP-assisted steel
characterized by means of the EBSD technique. Mater Sci Eng
2007;A447:285.
[31] Wu G, Juul Jensen D. Automatic determination of
recrystallization parameters based on EBSD mapping. Maters
Char 2008;59:794800.
[32] Mingard K, Roebuck B, Bennett E, Gee M, Nordenstrom H,
Sweetman G, Chan P. Comparison of EBSD and conventional
methods of grain size measurement of hardmetals. Int J
Refract Met Hard Mater 2009;27:21323.
[33] Randle V. Twinning-related grain boundary engineering
(overview). Acta Mater 2004;52:406781.
[34] Randle V. Sigma-boundary statistics by length and number.
Interface Sci 2002;10:2717.
[35] Schuh C, Kumar M, King W. Universal features of grains
boundarynetworks inFCCmaterials. J Mater Sci 2005;40:84752.
[36] Vonlanthen P, Grobety B. CSL grain boundary distribution in
alumina and zirconia ceramics. Ceram Int 2008;34:145972.
[37] Creuzigera A, Croneb W. Grain boundary fracture in CuAlNi
shape memory alloys. Mater Sci Eng 2008;A498:40411.
[38] Saylor D, El-Dasher B, Adams B, Rohrer G. Measuring the
five-parameter grain-boundary distribution from observations
of planar sections. Met Mat Trans 2004;35A:19819.
[39] Rohrer G, Randle V. Measurement of the five-parameter grain
boundary distribution from planar sections. In Electron
backscatter diffraction in materials science, Second Edition,
Editors Schwartz AJ, Kumar M, Adams BL, Field DP. Springer,
New York, in press.
[40] Pennock G, Coleman M, Drury M, Randle V. Grain boundary
plane populations in minerals; the example of wet NaCl after
low strain deformation. Contrib Mineral Petrol
2009;158:5367.
[41] Pennock G, Drury M, Spiers C. Grain boundary populations in
wet and dry NaCl. Mat Sci Tech 2006;22:1307.
[42] Munroe PR. The application of focused ion beam microscopy
in the materials sciences. Maters Char 2009;60:213.
922 M A T E R I A L S C H A R A C T E R I Z A T I O N 6 0 ( 2 0 0 9 ) 9 1 3 9 2 2

You might also like