You are on page 1of 12

X-Ray Diffraction (XRD)

David Alderton, Royal Holloway, Department of Earth Sciences, Egham, United Kingdom
© 2020 Elsevier Inc. All rights reserved.

Introduction 1
Historical Development of XRD 2
Instrumentation and Methods 3
Interpretation 6
Quantitative Analysis 9
Synchrotron XRD 9
Clays 10
Single Crystal X-Ray Diffraction 11
Conclusion 12
References 12
Further Reading 12

Glossary
Angstrom (Å) 1 Å ¼ 10−10 m (0.1 nm). This unit is not part of the S.I. system but is still used extensively as a measure of the
wavelengths of X-rays.
Bragg Law An equation that expresses the geometric relationship between X-ray wavelength, d spacing, and angle of incidence
during diffraction of X-rays from an atomic plane. nl ¼ 2dsiny where n is an integer, d is the interplanar spacing of atoms, and
y is the angle of incidence of the X-ray beam to the plane of atoms.
Crystallite Small single crystal of one material in powder form.
d spacing Distance between successive and similar atomic planes in a mineral.
Diffraction The action of constructive interference between waves reflecting from different crystal planes.
ICDD International Centre for Diffraction Data.
Ka, Kb X-rays produced by electron transmissions from the L to K shells and M to K shells respectively.
Lambda (l) Wavelength of X-ray.
Lattice As used here, a lattice is an infinite array of discrete atoms in a mineral generated by a set of translation operations into
three-dimensional space. (Note that there is some disagreement amongst crystallographers about the correct usage of
this term).
Miller indices The notation system in crystallography for labeling and describing planes in crystal lattices. There are (usually)
three integers written as (hkℓ), and these denote the family of planes in three-dimensional space. They represent a notation for
the intersections with the crystal axes.
Synchrotron A synchrotron generates extremely powerful X-rays from high-energy electrons as they circulate around a particle
accelerator.
Theta (u) Angle of incidence of X-rays on to a plane of atoms.
Unit cell The smallest repeating unit having the full symmetry of the crystal structure. The size of the cell is represented by the
letters a, b and c along three crystallographic axes, x, y and z.
X-ray X-rays are a form of high-energy electromagnetic radiation. Most X-rays have a wavelength between 10 and 0.1 Å.

Introduction

The method of X-ray diffraction (XRD) is based on the constructive interference of monochromatic X-rays in a crystalline sample.
Crystalline substances containing a network of atoms can diffract incident X-rays and generate an interference pattern. The resulting
pattern can be analyzed to obtain information about the atomic or molecular structure of that material, and thus the identity of the
phase. XRD is a rapid and powerful technique for identifying and characterizing materials; it is in common usage in many
geoscience and physical science research departments and is used extensively in industry. It requires minimal sample preparation
and the interpretation of the resulting data is usually straightforward.
This article will focus on the predominant XRD application in geology—that of powder XRD for identifying the minerals present
in a sample and quantifying their relative proportions in a mixture. However, mention will also be made of additional techniques
and applications to illustrate their broad range, including the determination of unit cell dimensions and single crystal XRD. Modern
XRD equipment tends to be fully automated and the user may not need to input much effort to generate a result. However, this
article will cover the principles of the method so that the user can understand how the results are generated.

Encyclopedia of Geology, 2nd edition https://doi.org/10.1016/B978-0-08-102908-4.00178-8 1


2 X-Ray Diffraction (XRD)

Historical Development of XRD

Wilhelm Röntgen discovered X-rays in 1895 but the initial applications of this discovery were to image the internal structure of solid
objects, especially in medical applications. German physicist Max von Laue postulated that these X-rays were a form of electro-
magnetic radiation with very short wavelengths (of the order of a nanometer) and he wondered whether crystalline material would
diffract X-rays, in the same manner as light on a scratched solid surface, in other words, that the crystalline substances would act as
three-dimensional diffraction gratings. Subsequently, in 1912, Laue and colleagues Friedrich and Knipping discovered that X-rays
were indeed scattered by the internal lattices of crystals, the first successful image being produced from a crystal of copper sulfate.
This truly dramatic and ground-breaking result demonstrated for the first time that: (1) crystalline materials consist of a regular,
three-dimensional arrangement of atoms, (2) the atomic spacings are of roughly the same magnitude as the X-ray wavelengths, and
(3) X-rays do indeed have a wave-like form.
A diagram of Laue’s original experimental set-up is shown in Fig. 1. In this, an X-ray beam is focused through a small hole in
some lead plates and directed at a stationary crystal (ideally along an axis of symmetry). The diffracted X-rays are recorded on a
photographic plate as spots. A modern “Laue” XRD photo of sodium chloride is shown in Fig. 2, and the fourfold symmetrical
arrangement of the X-ray pattern from the atomic arrangement in this cubic mineral is clearly visible.
When X-rays bombard a regular row of atoms, each atom emits a spherical wave front of X-rays. Most of these wavefronts act
destructively on each other, that is, they cancel each other out. However, in some specific directions they combine to reinforce each
other; this cooperative scattering effect, where in-phase wave fronts are produced, is known as diffraction. For a fixed angle of
incidence to the plane of atoms the emerging angle of the reinforced, diffracted wavefront will have a locus in three-dimensions in
the form a cone. In a three-dimensional network of atoms, each direction having its own atomic periodicity, three sets of cones with
different apical angles will be produced, each coaxial to the three different axes of atoms. Laue developed three equations that

Fig. 1 Diagram illustrating the set-up Laue and co-workers used to generate X-ray diffraction patterns. The X-rays were focused into a beam by passage through
small holes in lead screens and the beam was directed at a stationary crystal. The position of X-rays on the other side of the crystal was recorded on a photographic
plate. Although some of the X-rays passed straight through the crystal there were also several “spots” elsewhere (shown in red) and these often exhibited some form
of symmetrical arrangement.

Fig. 2 Laue photograph of the mineral halite (NaCl). The X-ray beam has been directed at the (001) face and the fourfold symmetry is clearly visible. Image created
with Crystal Studio, courtesy of Anthony Zhang.
X-Ray Diffraction (XRD) 3

expressed the geometry of these three intersecting cones. When diffraction occurs, these three cones will only overlap and reinforce
along one direction; this will be the diffraction line which will project towards, and can be recorded as a spot on, a photographic
plate, as in Figs. 1 and 2 (see Klein, 2002 for further descriptions).
Starting in 1912 (and soon after Laue’s discovery), father and son W.H. and W.L. Bragg discovered that, although diffraction was
taking place in three dimensions due to these overlapping cones related to rows of atoms, in a crystalline substance the diffracted
X-rays are more easily considered as being reflected from simple planes of atoms within the crystal. They developed a geometric
relationship between the incoming radiation and the diffracted X-rays similar to that of the Laue equations and went on to establish
the method for determining crystal structures. They started by solving the structure of NaCl, KCl, KBr and KI, and then soon
progressed to ZnS and diamond.
This concept is illustrated in Fig. 3. It shows that some X-rays can be reflected from a plane but others penetrate further and are
reflected from successively deeper planes. However, each set of reflected X-rays would tend to cancel each other out if they
re-emerged out of phase and thus they would not be recorded. However, under special conditions they would be in phase and
would reinforce the response. (In reality the resulting diffraction is not just due to two adjacent planes, but dependent on a large
number, each contributing a small part of the total). The geometry of this special condition is shown in Fig. 3, where two adjacent
X-ray paths are reflected and emerge in phase. The geometry of this special condition is expressed by the formula:

nl ¼ 2dsiny

where n is an integer, d is the interplanar spacing of atoms, and y is the angle of incidence of the X-ray beam to the plane of atoms.
This relationship is known as Bragg’s Law and is the fundamental equation of XRD.
In practice, each crystalline phase will have numerous sets of different planes that are able to diffract the X-rays at different angles
and satisfy Bragg’s equation. Each will have a unique separation distance (d) and intensity, the latter depending on the nature of the
plane of atoms (especially atom intensity). The characteristic set of “d-spacings” for a mineral, together with their respective
intensities, provides a unique “fingerprint” of the mineral present in the sample, one that enables identification of the material.

Instrumentation and Methods

Laue used “white” radiation, that is, the X-rays were polychromatic (they had a range of wavelengths across the spectrum). To extract
information using the Bragg equation there needs to be monochromatic radiation (i.e. one wavelength). X-rays are generated in a
cathode ray tube by heating a filament to produce electrons, accelerating the electrons towards a target by applying a high voltage,
and impacting the electrons with the target material; all of this is carried out in a vacuum. When electrons have sufficient energy to
dislodge and eject inner shell electrons of the target material, the vacancies produced are filled by electrons from outer shells falling
back to the inner levels. As a result, X-rays characteristic of the target element are produced. These X-rays consist of several
components, “white” or “continuous” radiation with a broad wavelength range, due to the deceleration of the electrons hitting
the target, and radiation that is specific to the target element. The most common of the latter is termed Ka radiation (from electron
transmissions from the L to K shells) and Kb (from M to K shells). Ka consists of both Ka1 and Ka2; each of these has a slightly
different wavelength. Ka1 has a slightly shorter wavelength and twice the intensity as Ka2, but Ka1 and Ka2 are sufficiently close in
wavelength that a weighted average of the two can be used.
Copper Ka radiation is normally chosen for XRD use in geological applications, because a copper target is an efficient producer
of X-rays (for Cu Ka weighted average l ¼ 1.5406 Å), but samples that are rich in Fe, Cr, or Mn will fluoresce under the incident Cu
Ka beam and create polychromatic radiation and elevated backgrounds. Alternative tube targets are possible: Co is often used for

A
G

C H
P1 Θ B Θ

ΘΘ d

P2 D F
E
d
P3
Fig. 3 Diagram illustrating Bragg’s Law. Two parallel X-ray waves (AB and CD) are directed at 3 identical planes of atoms (P1, P2 and P3); these are separated by
a distance d. The rays hit the planes of atoms at an angle of y. The X-ray AB is reflected from P1 as ray BG at an angle of y, while CD has penetrated to P2 and
reflects with the same angle as ray EH. If one considers the “wave fronts” AC, BD, BF and GH, it is clear from the diagram that the ray EH will only remain in phase
with BG and provide constructive reinforcement (and thus be measured) if certain conditions are met. The lengths of DE plus DF must equal a multiple of the
wavelength of the radiation. DE + DF ¼ 2d siny ¼ nl (where n is an integer, 1, 2, 3. . .). If this requirement is not fulfilled then the waves will tend to cancel each
other out.
4 X-Ray Diffraction (XRD)

samples rich in Fe, Mo is used for material with small unit cells (such as metals) and Cr is used to study materials with large
unit cells.
Filtering, by foils or crystal monochromators, is required to remove the (much lower intensity, around one sixth) Kb radiation
and reduce the “white” radiation of the continuous spectrum to produce the monochromatic X-rays needed for diffraction. Various
methods can remove this radiation—filters (Ni for Cu radiation and Fe for Co radiation), or (more precise) monochromators and
mirrors. Filters cut out the unwanted radiation, whilst a curved-crystal graphite monochromator directs the Kß radiation away,
because it is diffracted at a slightly different angle to that of the Ka radiation.
In Fig. 3, the X-rays are diffracted at an angle of y to the plane of atoms in question; however, the angle compared to the incident
beam is 2y. This figure only shows the relationship in two dimensions; if three dimensions are considered and the planes of atoms
are rotated through 360 degrees then the emerging diffracted X-ray wavefront would have a conical shape. For a long time the main
XRD method involved producing these “cones” and recording them on photographic film. The mineral under investigation should
be ground to a very fine powder to guarantee that there are very large numbers of grains in random orientations, such that all
possible alignments of atomic planes are present and exposed to the X-ray beam, thus ensuring that a certain fraction of all crystal
planes that satisfy Bragg’s Law would be exposed to the X-rays.
The powder is loaded into a glass capillary tube and rotated on an axis during radiation to further enhance the disposition of
planes to the X-rays. This concept is illustrated in Fig. 4. The Debye-Scherrer camera is a cylindrical casing that holds a strip of
photographic film that surrounds the powder and records the X-ray cone traces. The capillary tube and sample powder can be
centered and rotated about the camera axis to expose the maximum number of planes to the X-rays. Collimators direct the incoming
X-rays on to the sample.
A typical film is shown in Fig. 5. Measurement of the position of the various lines provides a value for the d spacing for each line,
using a simple geometric conversion based on the size of the camera used; in Fig. 5 the relationship is

S ð360 − 4yÞ
¼
2pr 360
where S is the measured distance between a pair of arcs and r is the radius of the camera. Theta (y) can thus be derived for each
reflection and added to Bragg’s equation to obtain the relevant d value.

X-ray
path

Fig. 4 Diagram showing how X-ray diffraction powder patterns are produced in a Debye-Scherrer camera. The X-ray reflections from one plane form a cone with
an angle of 4y.The lower image illustrates how the conical forms of individual reflections are recorded on the strip of film as pairs of arcs, separated by a distance S.

Fig. 5 An X-ray diffraction photo produced from a mineral powder (halite, NaCl) in a Debye-Scherrer camera. The geometry of this photographic film is known as
the Straumanis setting (and is different to that illustrated in Fig. 4). Here the X-rays enter the camera via a hole in the film (to the left, not visible) and exit the hole on
the right side. The strong lines on the right side of the film represent reflections at 2y angles <90 degrees, the weaker lines to the left >90 degrees. The y values
(and thus d spacing) for each line can readily be determined using the equation given in the text, if the camera radius is known, or through measurement of the
distances between the two holes and between pairs of similar reflection lines (S in Fig. 4) (Zussman, 1967).
X-Ray Diffraction (XRD) 5

An advancement on this camera design is the Gandolfi camera which rotates the sample along two-axes, thus exposing more
planes to the X-rays and allowing smaller samples to be analyzed (this camera method can use <0.1 g of material).
These cameras traditionally used photographic film for recording but, even though the camera and film method has some
advantages (e.g. small sample volumes can be determined), this recording medium slowly fell out of favor in the 1970s. However,
with the development of alternative detecting technologies and the introduction of software packages that convert a 2-D image into
a 1-D diffractogram, the camera methodology has witnessed some recent resurgence (e.g. Ross et al., 2014).
The standard methodology used at present is the powder diffractometer that utilizes moving detectors and data output on to
computer screen displays (formerly chart recorders) (actually the Braggs built and used the first such instrument with this design in
1913). With appropriate software, the resulting patterns can be converted to digital data and analyzed in greater detail and with
greater accuracy.
For the powder diffractometer, the finely-ground (ideally to <10 mm) powder is pressed into a holder or smeared with water or
acetone, ethanol, etc. on to a slide (usually glass), ensuring a flat upper surface. For normal identification purposes there should be
no preferred orientation of the grains as this would cause some reflections to have increased intensity (this can be a problem with
platy minerals such as mica). Approximately 1 g or more of powder is typically needed. Less material can be used if the sample is
smeared on to a special silicon crystal wafer, as this reduces the background and enhances the peaks. Alternatively, it is also possible
to improve the reflection response by rotating the sample in a “spinner” to expose more planes to the X-ray beam.
The basic geometry of an X-ray powder diffractometer is illustrated in Fig. 6 and involves a source of monochromatic radiation
and a moving X-ray detector situated on the circumference of a graduated circle centered on the powder specimen (Fig. 7). Slits

Fig. 6 Diagrammatic representation of the working of a y–2y powder diffractometer. The X-rays (yellow) are collimated by passage through slits (S), before and
after striking the sample mounted on a holder. The sample rotates through an angle of y whilst the detector moves around the measuring circle at an angle of 2y.
This maintains the geometry for satisfying Bragg’s Law.

Fig. 7 Image of the sample chamber in a powder diffractometer. Note that the door to the chamber has been removed for viewing; a safety mechanism would not
allow this to happen if X-rays were being transmitted.
6 X-Ray Diffraction (XRD)

located between the X-ray source and the specimen, and between the specimen and the detector, cut out some of the non-diffracted
radiation, reduce background noise, and focus the desired X-rays on to the sample. A goniometer rotates the detector and specimen
holder, the former through twice the angle of the latter, fixing a 2:1 ratio. This is the most common geometry and is known as the
Bragg-Bentano diffractometer y–2y system. Other systems are also used (although less commonly)—for instance the y–y geometry,
where the sample is stationary in the horizontal position, and the x-ray tube and the detector both move simultaneously through
the angle theta.
At the (2y) angles that satisfy the Bragg relationship for a particular plane of atoms there will be a peak in X-ray intensity
recorded. At other angles there will be no diffraction and so only background scattered radiation is recorded.
The signals from the detector are filtered, scaled to measurable proportions, and sent to a ratemeter for conversion into a
continuous current. The output devices include strip-chart recorders, printers, and (mostly now) computer monitors and the output
consists of a plot of intensity of X-rays versus 2y angle (e.g. Fig. 8A and B).

Interpretation

The main geological use of powder X-ray diffraction is to identify unknown mineral phases in a sample. Traditionally the user would
undertake the identifications manually: they would measure the angle of each reflection and convert that to a d spacing using the
relevant wavelength of radiation in Bragg’s Law. The resulting d values would then be compared to a minerals database.
There are several databases available, some commercial, some free of charge. One of the most commonly-used is produced by
the International Centre for Diffraction Data (ICDD). This organization (formerly known as the Joint Committee on Powder
Diffraction Standards, JCPDS), is the organization that maintains the most comprehensive database of inorganic and organic

(A)
Sample identification
[counts]
14400

10000

6400

3600

1600

400

0
0 10 20 30 40 50 [°20] 60
(B)
Sample identification
[counts]
196

144

100

64

36

16

0
0 10 20 30 40 50 [°20] 60

Fig. 8 XRD traces (X-ray intensity against angle (2y)) (A) Quartz. This mineral is well crystallized and therefore shows a high intensity of diffracted X-rays and very
sharp peaks. (B) Manganese oxide. This mineral is poorly crystallized; the peaks are not well-defined and have low intensities. The lines represent where the
computer software has determined that there is a peak.
X-Ray Diffraction (XRD) 7

patterns and since the early 1940s has collected and maintained an archive of single-phase X-ray powder diffraction patterns in the
form of tables of characteristic interplanar spacings and corresponding relative intensities, along with other pertinent physical and
crystallographic properties. Their minerals database now has nearly 50,000 separate entries, often with numerous entries for variants
of the same mineral species. Initially, searching was achieved by placing the set of d values in order of decreasing intensity of peaks
(the Hanawalt method) or decreasing d values (the Fink method). The printed search manuals with tabulated mineral datasets
would then be used to identify likely candidates for the unknown material based on a few of the most intense peaks and the identity
could finally be confirmed using the complete dataset for that mineral (e.g. Fig. 9). Much of this is now carried out automatically
using computer software (Fig. 10), but occasionally the user will need to resort to printed tabulations for unclassified peaks or
unlikely candidates that have been suggested by the software.

Fig. 9 XRD and crystallographic information for Halite (NaCl). The upper image lists the d values, their respective, relative intensities, and the Miller indices for the
relevant crystallographic planes. This card is taken from the Mineral Powder Diffraction File Databook, Sets 1–42, published by the ICDD (1993). The lower image
shows the output for the same reference trace from the electronic database (Gates-Rector S and Blanton T (2019) Powder Diffraction, 34: 352–360). Both images
are produced courtesy of ICDD.
8 X-Ray Diffraction (XRD)

(A)
2000

Counts
00-011-0078 Dolomite

00-008-0048 Orthoclase

00-026-0801 Pyrite, syn


1600 00-002-0056 lllite

00-001-0649 Quartz

1200

800

400

0
10 20 30 40 50
Degrees 2-Theta
(B)
l rel.
1000

950 Calculated pattern (exp. peaks) (Rp=38.4 %)


900 Background
[96-901-0145] O2 Si Quartz (55.8%)
850
[96-901-5000] Al2 O9 Si2 Kaolinite (22.6%)
800 [96-901-3720] Al2 H2 KO12 Si4 lllite (21.7%)
750

700

650

600

550

500

450

400

350

300

250

200

150

100

50

5.00 10.00 15.00 20.00 25.00 30.00 35.00 40.00 45.00 50.00 55.00
Cu-Ka (1.541838 A) 2theta

Fig. 10 XRD traces of two separate rock samples with peak markers indicating likely mineral phases present. (A) Using the commercial programme Traces and
the ICDD Mineral database. (B) Using the programme Match. Match is commercial software that uses the open-access Crystallography Open Database (COD). The
proportions are approximate and not derived from Rietveld analysis.
X-Ray Diffraction (XRD) 9

As well as identifying unknown phases, it is also possible to use powder diffractometry to determine or refine the cell dimensions
for the mineral. Ideally the crystal system and the (h,k,l) values of each of the peaks should be known; if these are unknown then it
still may be possible to determine them from the peak values. In the case of simple structures, such as minerals in the cubic crystal
system, peaks can be assigned Miller indices manually by looking for integer relationships between the resulting interplanar
spacings. In more complex cases it is probably easier to use single crystal XRD (see below). There are simple formulae which relate
the (h,k,l) indices for specific peaks to cell dimensions and d spacings. For the orthorhombic crystal system the relationship is:

1 h2 k 2 l2
¼ + 2 + 2
d2 a2 b c
Therefore, if the crystal system and (h,k,l) indices for some appropriate peaks are known, then unit cell dimensions can be
calculated. These relationships are simpler for cubic and tetragonal crystal systems because of the reduced cell dimension
unknowns, but they get progressively more difficult to use for minerals of the monoclinic and triclinic systems (see Zussman,
1967) and it is better to use the available computer programs for this purpose.

Quantitative Analysis

Quantitative analysis can be performed on mixtures of minerals using XRD, if their mineral content has already been confirmed.
Several approaches have been used in the past, for instance by mixing with an internal standard. The aim is essentially to relate the
intensity of the measured peak to concentration. To do this, there needs to be a calculation of the various factors that can influence
this, in addition to concentration. These are numerous and complex, but include physical factors (the intensity of the incident beam,
polarization of the reflected beam, multiplicity of reflections from crystallographically equivalent planes, absorption of X-rays by
the sample) and structural factors (determined by the number, kind and position of atoms in the unit cell) (see Zussman, 1967).
One of the most common methods now used for quantification was initiated by Hugo Rietveld (1969) when he developed a
methodology for atomic structure refinement using neutrons. It has since found wider usage and is now very common in XRD
applications; for instance it is much used in the cement industry in situations where more than 30 phases need to be quantified in a
sample. The calculations are complex and are therefore applied using computer packages (many commercial, some freely available).
Instead of the conventional approach to XRD quantification using individual peaks the Rietveld method analyses the whole XRD
profile. The method computes theoretical phase XRD profiles from crystallographic data for the individual identified minerals
stored in a database. Equations developed from this data incorporate the atomic structure of each phase and the various other
parameters related to both the minerals and diffractometer aberrations (unit cells, polarization of X-rays, temperature factors,
absorption, preferred orientation) that can influence the height of a peak. The theoretical profile can then be adjusted by modifying
the relevant parameters to get a best fit of the trace (the pattern is “refinable”).
Rietveld analysis can provide an estimate of accuracy through “goodness of fit,” that is comparing the experimental trace with the
theoretical trace. The difference (“residual”) gives a measure of the difference between the two traces. The limits of quantification for
individual minerals by XRD will vary depending on the minerals in question; for a well-crystallized mineral like quartz with
minimal unit cell variations it may be as low as about 1 wt%, but this value will be higher for other phases. An example of Rietveld
analysis of a sediment sample is shown in Fig. 11.

Synchrotron XRD

In recent years synchrotron facilities have become widely used as sources of X-rays for XRD measurements. Synchrotron radiation is
emitted by electrons that are accelerated to near light speed in a circular storage ring and the resulting X-rays are thousands to
millions of times more intense than the X-ray tubes used in the laboratory. These X-rays offer several advantages over conventional
XRD: the X-rays are tuneable to different wavelengths, they have high intensities and thus sensitivity is increased, the spectral
resolution is high, the X-rays can be focussed to very small areas (mXRD), and the analysis can be rapid. Synchrotron XRD has
allowed the analysis of much smaller and more dilute phases, and the consequent determination of very fine structures (see Gräfe
et al., 2014; Lo et al., 2018) including crystal defects and plastic deformation in minerals (Moore, 2012).
Energy-dispersive X-ray diffraction differs from conventional XRD by using polychromatic radiation as the source and is usually
operated at a fixed angle. Because there is no goniometer the method can collect diffraction patterns very rapidly. It is especially
suitable for studies in special environments (very low or high temperatures and/or pressures) or on materials that are unstable and
only exist for short periods of time. Energy dispersive XRD is almost exclusively used with synchrotron radiation but has also been
applied to mineralogy in the laboratory (e.g. Ferrell, 1971; Escárate et al., 2009).
There are several other parameters that can be derived through XRD analysis:
Amorphous content: Amorphous phases cannot be accurately measured using XRD but it is possible to measure the amounts of
amorphous material in a mixture present by comparing Rietveld analysis of the mixture to a sample with an addition of a known
amount of an internal standard.
Crystallite size: The angular spread of the reflection from a crystal plane is affected not only by the crystallinity of the mineral but
also by the size of the crystallite. The peak width b in radians (often measured as full width at half maximum, FWHM) is inversely
10 X-Ray Diffraction (XRD)

5,000

4,500

4,000

3,500

3,000
Intensity

2,500

2,000

1,500

1,000

500

0
4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 44 46 48 50

Two-Theta
150
100
50
0
–50
–100
–150

Fig. 11 An example of Rietveld analysis of a sediment sample using the programme Siroquant (Sietronics, Australia). The green trace is the original XRD trace, the
red trace is the calculated pattern and the blue trace is the residual. The quantification gave (in weight %): Quartz 35.0, Calcite 24.8, Illite 22.0, Kaolinite 8.8, Albite
5.0, K feldspar 1.6, Goethite 1.2 and others, each <1.

proportional to the crystallite size. As the average size of the crystallites decreases, so the angular spread of the reflection from the
minerals will increase. After suitable calibration, the FWHM of a reflection in a powder diffractogram can be used as a quantitative
measure of the mean crystallite size in the sample (the Scherrer equation).
Stress analysis: If a mineral is strained, the unit cell dimensions may be altered slightly, so changing the d values. Accurate
measurement of stress can be obtained by measuring these differences in the specimen before and after straining.

Clays

XRD is extremely useful for characterizing clay minerals and in fact it was instrumental in settling the early debate regarding whether
clays in soils were crystalline or amorphous materials. Very often clay minerals are too fine-grained to identify using an optical
microscope and because they are finely admixed with other phases their chemistry is difficult to determine accurately. XRD of clay-
bearing assemblages can produce excellent diffraction patterns and it is often straightforward to establish which species is present.
In addition there are further treatments to the sample that can aid identification.
Normally, clay concentrates are used for analysis so that peaks are enhanced. Concentrates are produced from the sample by
dissolving carbonates, organic materials and iron oxides and then separating the clays through particle-size separation methods.
Quartz often remains but it can actually be useful because it has a precise set of peaks that can be used as an internal standard for
accurate work. Ideally the clay samples should then be saturated with specific chloride solutions, because many of the XRD
characteristics depend on the cation held in the exchange sites; addition of these solutions to the sample can lead to ion exchange
and consequent expansion or collapse, depending on the clay and the nature of the solutions used.
In contrast to normal powder XRD identification procedures, clay patterns are often derived from oriented clay samples. This is
because their similar structures have x and y dimensions which are very close. It is the z spacing that varies and is characteristic of the
mineral species, and oriented samples enhance these basal (001, 002, etc.) reflections (patterns from the ICDD database are
typically from randomly-oriented samples, so are not very helpful for precise clay mineral identification).
For clays, the main peaks are at high d values (¼ low 2y angles); those <40 degrees are often distinctive, and sometimes those
appearing at angles <20 degrees are sufficient for routine identification.
An example showing an XRD pattern for a mixture of clays (with quartz) is shown in Fig. 10B and for a clay concentrate in
Fig. 12.
Identification of clay minerals can be further aided by various procedures, and these may be especially useful when analyzing
mixed-layer clay minerals:
Solvation with ethylene glycol: this is absorbed by smectite clays and increases the d value for the (001) reflection. So a
comparison of an air-dried trace with one saturated with ethylene glycol should reveal a decrease in the 2Y value (Fig. 12).
In addition, heating the sample to 600  C collapses the structure of some clay minerals. For instance kaolinite converts to an
amorphous material and the peaks are lost, and smectite converts to illite (Fig. 12). This procedure aids the distinction of kaolinite
from chlorite, as the latter has some peaks in identical positions to those of kaolinite.
X-Ray Diffraction (XRD) 11

1600
C
o
u
n
t
s
1200

800

400

0
5 10 15 20 25

Degrees 2-Theta

Fig. 12 XRD trace of a clay concentrate in the angular range 2y ¼ 4 to 25 . The red trace is the air-dried sample; it shows a strong peak at 6.1 (d ¼ 14.6 Å).
Glycollation of the sample (blue trace) has resulted in a shift of the main peak to a lower angle (5.6 ; d ¼ 15.9 Å) indicative of a smectite clay. Heating to 600  C
(green trace) collapses the smectite structure but leaves a small illite peak at 8.8 (d ¼ 10.1 Å).

Quantification of clay mineral proportions in a mixture can present some problems due to uncertainties regarding the
crystallographic structure and properties of the clay mineral being analyzed.
The interested reader is directed to more specialist publications (e.g. Moore and Reynolds, 1997) for more detailed information
on the XRD analysis of clay minerals.

Single Crystal X-Ray Diffraction

Although reasonably accurate cell parameters and d spacings can be obtained by powder diffraction methods as outlined above,
single crystal methods are much preferred. This is because the reflections from planes can be recorded separately and more precisely,
the crystal can be aligned in specific orientations, the symmetry of the crystal is more easily appreciated and much smaller amounts
of material can be analyzed. Although single-crystal X-ray diffraction can be time-consuming and complicated, it is the common
method used for characterizing new crystalline materials: precise determinations of a unit cell, including cell dimensions and
positions of atoms within the lattice, and cation and anion coordination.
The single crystal set-up is rather similar to that of the Laue method as shown in Fig. 1. However, Laue used polychromatic
radiation and a stationary crystal. Developments of this approach use monochromatic radiation (thus fulfilling Bragg’s Law) and
moving crystals (thus exposing more planes to satisfy the Law). Initial developments from the original Laue method were the
Weissenberg and Precession cameras that introduced different modes of rotation and disposition of the photographic film for
recording the patterns. Originally, photographs were taken along crystallographic axes and indexing performed manually. Nowa-
days, data are collected with an electronic detector and the data are indexed with automatic computer programs.
The introduction of digital computers in the late 1970s led to the design of the automatic four-circle diffractometers. These
instruments use goniometers controlling three rotation axes, allowing a crystal to be aligned in any orientation in space. Once the
crystal is oriented, a fourth axis of rotation, which supports the electronic detector, is placed in the right position to collect the
diffracted beam. All these movements can be programmed in an automatic mode, with minimal operator intervention.
Crystals should be as clear and free from inclusions as possible, and a few tens or hundreds of microns in size. The position of the
crystal can be adjusted until it can be rotated about a crystallographic axis. In order to get the largest and best collection of diffraction
data, crystal samples are often maintained at a very low temperature (about 100  K, −170  C). Molybdenum (lMoKa ¼ 0.7107 Å) is
the most common target material for single-crystal diffraction work.
X-rays leave the collimator and are directed at the crystal. Diffracted rays at the correct orientation for the configuration are then
collected by the detector. Originally this would have been photographic film, but now CCD (charge-coupled device) technology is
used to convert the X-ray photons into an electrical signal that is then processed.
12 X-Ray Diffraction (XRD)

The aim is to fulfill the Bragg condition for as many crystal planes as possible by rotating the crystal around the three axes and the
detector around its own axis in order to detect the diffracted beams. One advantage of this method is that a sphere or hemisphere of
data can be collected using an incremental scan method, collecting frames in small angular increments. A complete data collection
may require anywhere between 6 and 24 h, depending on the specimen and the diffractometer. Exposure times of 10–30 s per frame
for a hemisphere of data will require total run times of 6–13 h. Older diffractometers with non-CCD detectors may require 4–5 days
for a complete collection of data.
Using these rotational axes as a coordinate system, the diffraction peaks themselves describe a lattice in three-dimensional space
that corresponds to the actual crystal lattice and the disposition of the reflections gives an indication of crystal symmetry. Then by a
series of complex iterations it is possible to relate the position of the crystal and the XRD pattern to the crystal system and the unit
cell dimensions.

Conclusion

XRD has come a remarkably long way since its introduction just over 100 years ago, although the same questions are still being
tackled. It is now possible to make measurements at high and low temperatures through the attachments of heating and cooling
stages and thus monitor the variations of crystal parameters with temperature. And XRD equipment is now available in a variety of
user-friendly forms, including smaller benchtop models and even portable equipment. One has even been landed on Mars. CheMin
(Chemistry and Mineralogy) is an XRD instrument that has been exploring the surface of Gale crater on Mars and analysis of Martian
soil has revealed the presence of several minerals, including feldspar, pyroxene and olivine.

References
Escárate P, Bailo D, Guesalaga A, and Rossi Albertini V (2009) Energy dispersive X-ray diffraction spectroscopy for rapid estimation of calcite in copper ores. Minerals Engineering
22: 566–571.
Ferrell RE (1971) Applicability of energy-dispersive X-ray powder diffractometry to determinative mineralogy. American Mineralogist 56: 1822–1831.
Gräfe M, Klauber C, Gan B, and Tappero RV (2014) Synchrotron X-ray microdiffraction (mXRD) in minerals and environmental research. Powder Diffraction 29(S1): S64–S72.
ICDD (1993) Mineral Powder Diffraction File Databook (Sets 1–42). International Centre for Diffraction Data.
Klein C (2002) Mineral Science, 22nd edn John Wiley and Sons.
Lo BTW, Ye L, and Tsang SCE (2018) The contribution of synchrotron X-ray powder diffraction to modern zeolite applications: A mini-review and prospects. Chem 4(8): 1778–1808.
Moore DM and Reynolds RC (1997) X-ray Diffraction and the Identification and Analysis of Clay Minerals, 2nd edn Oxford: Oxford University Press.
Moore M (2012) White-beam X-ray topography. Crystallography Reviews 18: 207–235.
Rietveld HM (1969) A profile refinement method for nuclear and magnetic structures. Journal of Applied Crystallography 2: 65–71.
Ross KC, Petrus JA, and McDonald A (2014) An empirical assessment of two-dimensional (2D) Debye–Scherrer-type image-plate X-ray diffraction data collapsed into a 1D
diffractogram. Powder Diffraction 29(04): 337–345.
Zussman J (1967) X-ray diffraction. In: Zussman J (ed.) Physical Methods in Determinative Mineralogy, pp. 261–334. Academic Press.

Further Reading
Bish DL and Post JE (1989) Modern Powder Diffraction: Reviews in Mineralogy. vol. 20. Mineralogical Society of America.
Brindley GW and Brown G (1980) Crystal Structures of Clay Minerals and Their Identification. Monograph 5, Mineralogical Society.
Cullity BD (1978) Elements of X-ray Diffraction, 2nd edn Menlo Park, California: Addison-Wesley Publishing Co.
Cullity BD and Stock SR (2001) Elements of X-ray Diffraction, 3rd edn Prentice Hall.
Hammond C (2001) The Basics of Crystallography and Diffraction, 2nd edn Oxford University Press.
Jenkins R and Snyder RL (1996) Introduction to Powder Diffractometry. New York: John Wiley and Sons.
Klug HP and Alexander LE (1974) X-ray Diffraction Procedures, 2nd edn New York: John Wiley and Sons.

Relevant Websites

https://serc.carleton.edu/research_education/geochemsheets/techniques/SXD.html—Carleton University, Ottawa, Canada.


http://ww1.iucr.org/iucr-top/comm/cteach/pamphlets/2/2.html—An Introduction to the Scope, Potential and Applications of X-ray Analysis By Michael Laing.
https://www.xtal.iqfr.csic.es/Cristalografia/index-en.html—An excellent introduction to diffractometer methods and instrumentation.
doitpoms.ac.uk—Teaching & Learning Packages on the subject of crystallography and X-ray diffraction from the University of Cambridge

You might also like