Sei sulla pagina 1di 19

CHAPTER 1

Principles of Powder Diffraction


ROBERT E. DINNEBIERa AND SIMON J. L. BILLINGEb Max-Planck-Institute for Solid State Research, Heisenbergstrasse 1, D-70569 Stuttgart, Germany; b Department of Physics and Astronomy, 4268 Biomed. Phys. Sci. Building, Michigan State University, East Lansing, MI 48824, USA
a

1.1

INTRODUCTION

This chapter presents some very basic results about the geometry of diffraction from crystals. This is developed in much greater detail in many textbooks but a concise statement of the basic concepts greatly facilitates the understanding of the advanced later chapters so we reproduce it here for the convenience of the reader. Since the results are so basic, we do not make any attempt to reference the original sources. The bibliography at the end of the chapter lists a selection of some of our favorite introductory books on powder diffraction. 1.2 FUNDAMENTALS

X-rays are electromagnetic (em) waves with a much shorter wavelength than ( 1 1010 m). The physics of visible light, typically on the order of 1 A em-waves is well understood and excellent introductions to the subject are found in every textbook on optics. Here we briey review the results most important for understanding the geometry of diffraction from crystals. Classical em-waves can be described by a sine wave that repeats periodically every 2p radians. The spatial length of each period is the wavelength l. If two identical waves are not coincident, they are said to have a phase shift with respect to each other (Figure 1.1). This is either measured as a linear shift, D on a length scale, in the units of the wavelength, or equivalently as a phase shift, df on an angular scale, such that: D dj 2p ) dj D l 2p l
1

Chapter 1

Figure 1.1

Graphical illustration of the phase shift between two sine waves of equal amplitude.

The detected intensity, I, is the square of the amplitude, A, of the sine wave. With two waves present, the resulting amplitude is not just the sum of the individual amplitudes but depends on the phase shift dj. The two extremes occur when dj 0 (constructive interference), where I (A1 + A2)2, and dj p (destructive interference), where I (A1 A2)2. In general, I [A1 +A2 exp (idj)]2. When more than two waves are present, this equation becomes: " I X
j

#2 Aj exp ijj ; 2

where the sum is over all the sine-waves present and the phases, fj are measured with respect to some origin. X-ray diffraction involves the measurement of the intensity of X-rays scattered from electrons bound to atoms. Waves scattered at atoms at different positions arrive at the detector with a relative phase shift. Therefore, the measured intensities yield information about the relative atomic positions (Figure 1.2).

Principles of Powder Diffraction

Figure 1.2

Scattering of a plane wave by a one-dimensional chain of atoms. Wave front and wave vectors of different orders are given. Dashed lines indicate directions of incident and scattered wave propagation. The labeled orders of diffraction refer to the directions where intensity maxima occur due to constructive interference of the scattered waves.

In the case of X-ray diffraction, the Fraunhofer approximation is used to calculate the detected intensities. This is a far-eld approximation, where the distance, L1, from the source to the place where scattering occurs (the sample), and then on to the detector, L2, is much larger than the separation, D, of the scatterers. This is an excellent approximation, since in this case D/L1 E D/L2 E 1010. The Fraunhofer approximation greatly simplies the mathematics. The incident X-rays form a wave such that the constant phase wave front is a plane wave. X-rays scattered by single electrons are outgoing spherical waves that again appear as plane waves in the far-eld. This allows us to express the intensity of diffracted X-rays using Equation (2). The phases jj introduced in Equation (2), and therefore the measured intensity I, depend on the position of the atoms, j, and the directions of the incoming and the scattered plane waves (Figure 1.2). Since the wave-vectors of the incident and scattered waves are known, we can infer the relative atomic positions from the detected intensities. From optics we know that diffraction only occurs if the wavelength is comparable to the separation of the scatterers. In 1912, Friedrich, Knipping and Max von Laue performed the rst X-ray diffraction experiment using single crystals of copper sulfate and zinc sulte, proving the hypothesis that X-rays are em-waves of very short wavelength, on the order of the separation of the atoms in a crystalline lattice. Four years later (1916), Debye and Scherrer reported the rst powder diffraction pattern with a procedure that is named after them. 1.3 DERIVATION OF THE BRAGG EQUATION

The easiest access to the structural information in powder diffraction is via the well-known Bragg equation (W. L. Bragg, 1912), which describes the principle

Chapter 1

of X-ray diffraction in terms of a reection of X-rays by sets of lattice planes. Lattice planes are crystallographic planes, characterized by the index triplet hkl, the so-called Miller indices. Parallel planes have the same indices and are equally spaced, separated by the distance dhkl. Bragg analysis treats X-rays like visible light being reected by the surface of a mirror, with the X-rays being specularly reected at the lattice planes. In contrast to the lower energy visible light, the X-rays penetrate deep inside the material where additional reections occur at thousands of consecutive parallel planes. Since all X-rays are reected in the same direction, superposition of the scattered rays occurs. From Figure 1.3 it follows that the second wave travels a longer distance PN before and NQ after reection occurs. Constructive interference occurs only if D PN + NQ is a multiple n 0, 1, 2, . . . of the wavelength l: D nl 3

In all other cases, destructive interference results since it is always possible to nd a deeper plane, p, for which the relation pD nl with n 1/2, 3/2, . . . . (i.e., perfect destructive interference) exactly holds. Thus, sharp intensity maxima emerge from the sample only at the special angles where Equation (3) holds, with no intensity in between. As can be easily seen from Figure 1.3, geometrically: D 2d sin y 4

where d is the interplanar spacing of parallel lattice planes and 2y is the diffraction angle, the angle between the incoming and outgoing X-ray beams. Combining Equations (3) and (4) we get: nl 2d sin y the famous Bragg equation. 5

Figure 1.3

Illustration of the geometry used for the simplied derivation of Braggs law.

Principles of Powder Diffraction

This simplied derivation of the Bragg equation, although leading to the correct solution, has a serious drawback. In reality the X-rays are not reected by planes but are scattered by electrons bound to the atoms. Crystal planes are not like shiny optical mirrors, but contain discrete atoms separated by regions of much lower electron intensity, and, in general, atoms in one plane will not lie exactly above atoms in the plane below. How is it then that the simplied picture shown in Figure 1.3 results in the correct result? A more general description (Bloss, 1971) shows that Equation (5) is also valid, if the atom of the lower lattice plane in Figure 1.3 is shifted by an arbitrary amount parallel to the plane (Figure 1.4). The phase shift can immediately be deduced from Figure 1.4 as: nl MN cos180 a y MN cosa y MN cosa y cosa y From any textbook on trigonometry we know that: cosa y cos a cos y sin a sin y cosa y cos a cos y sin a sin y Therefore Equation (7) becomes: nl MN 2 sin a sin y with: d MN sin a from which the already known Bragg equation follows: nl 2d sin y 10 9 8 7 6

Another equivalent, and highly useful, expression of the Bragg equation is: 6:199 12:398 Ed with lA 11 sin y E keV with the energy E of the X-rays in keV.

P N

180-(+)

Figure 1.4 Illustration of the geometry in the general case where scattering takes place
at the position of atoms in consecutive planes.

Chapter 1

To derive the Bragg equation, we used an assumption of specular reection, which is borne out by experiment. For crystalline materials, destructive interference completely destroys intensity in all directions except where Equation (5) holds. This is no longer true for disordered materials where diffracted intensity can be observed in all directions away from reciprocal lattice points, known as diffuse scattering, as discussed in Chapter 16.

1.4

THE BRAGG EQUATION IN THE RECIPROCAL LATTICE

As a prerequisite, the so-called reciprocal lattice needs to be introduced. Notably, it is not the intention of this book to reproduce basic crystallographic knowledge but, for completeness, some important formalism that recurs throughout the book is briey presented. The reciprocal lattice was invented by crystallographers as a simple and convenient representation of the physics of diffraction by a crystal. It is an extremely useful tool for describing all kinds of diffraction phenomena occurring in powder diffraction (Figure 1.5). Imagine that besides the normal crystal lattice with the lattice parameters a, b, c, a, b, g, and the volume V of the unit cell, a second lattice with lattice parameters of a*, b*, c*, a*, b*, g*, and the volume V*, and with the same origin, exists such that: a b a c b c a b a c b c 0 w a a b b c c 1 12

This is known as the reciprocal lattice, which exists in so-called reciprocal space. As we will see, it turns out that the points in the reciprocal lattice are related to the vectors dening the crystallographic planes. There is one point in the reciprocal lattice for each crystallographic plane, hkl. For now, just consider h, k and l to be integers that index a point in the reciprocal lattice. A reciprocal lattice vector hhkl is the vector from the origin of reciprocal space to the (hkl) reciprocal lattice point:z hhkl ha kb l c ; h; k; l 2 Z The length of the reciprocal base vectors is dened according to: a xb c where the scale factor x can easily be deduced using Equation (12) as: a a xb ca xV ) x
w z

13

14

1 V

15

Vectors are in bold. The reciprocal lattice is a commonly used construct in solid state physics, but with a different normalization: a a 2p:

Principles of Powder Diffraction

Figure 1.5 leading to:

Two-dimensional monoclinic lattice and its corresponding reciprocal lattice.

a and vice versa: a

1 1 1 b c; b c a; c a b; V V V

16

1 1 1 b c ; b c a ; c a b V V V

17

The relationship between the reciprocal and the real lattice parameters is: bc sin a ; a V ac sin b ; b V ab sin g ; c V cos b cos g cos a 18 ; cos a sin b sin g cos b cos g cos a ; cos b sin b sin g cos b cos g cos a ; cos a sin b sin g p V abc 1 2 cos a cos b cos g cos2 a cos2 b cos2 g Equation (18) simplies considerably for higher symmetry crystal systems.

Chapter 1

S0 S-S0 2 S

S0

M
r

0
Figure 1.6

Illustration of the important wave and scattering vectors in the case of elastic Bragg scattering.

We now rederive Braggs law using vector notation. The wave vectors of the incoming and outgoing beams are given by s0 and s, respectively (Figure 1.6). They point in the direction of propagation of the wave and their length depends on l. For elastic scattering (no change in wavelength on scattering), s0 and s have the same length. We dene the scattering vector as: h s s0 19

which for a specular reection is always perpendicular to the scattering plane. The length of h is given by: h 2 sin y s Comparison with the formula for the Bragg equation [Equation (5)]: nl 2 sin y d we get: nl h d s 22 21 20

Setting the magnitude of s to 1/l, we get the Bragg equation in terms of the magnitude of the scattering vector h: n h 23 d This shows that diffraction occurs when the magnitude of the scattering vector h is an integral number of reciprocal lattice spacings 1/d. We dene a vector d*

Principles of Powder Diffraction

perpendicular to the lattice planes with length 1/d. Since h is perpendicular the scattering plane, this leads to: h nd 24

Diffraction can occur at different scattering angles 2y for the same crystallographic plane, giving the different orders n of diffraction. For simplicity, the number n will be incorporated in the indexing of the lattice planes, where:
dnh ;nk;nl ndhkl

25

* e.g., d* 222 2d111 and we get an alternative expression for Braggs equation:

h d hkl

26

The vector d* hkl points in a direction perpendicular to a real space lattice plane. We would like to express this vector in terms of reciprocal space basis vectors a*, b*, c*. First we dene dhkl in terms of real space basis vectors a, b, c. Referring to Figure 1.7, we can dene that: 1 1 1 OA a; OB b; OC c h k l 27

with h, k, and l being integers as required by the periodicity of the lattice. These three integers are the Miller indices that provide a unique definition for the set of parallel planes. The plane perpendicular vector dhkl originates on one plane and terminates on the next parallel plane. Therefore, OA d (OA)dcosa. From Figure 1.7 we see that, geometrically, (OA)cosa d. Substituting, we get OA d d2. Combining with Equation (27) leads to: 1 a d d2 h 28

c b
B C D

d
O A

Figure 1.7

Geometrical description of a lattice plane in terms of real space basis vectors.

10 and consequently: ha d d d ; k b 2; l c 2 d2 d d

Chapter 1

29

By definition, h, k, and l are divided by their largest common integer to be Miller indices. The vector d* hkl, from Braggs Equation (26) points in the plane normal direction parallel to d but with length 1/d. We can now write d* hkl in terms of the vector d: d hkl which gives: d hkl ha kb l c or written in terms of the reciprocal basis:
d hkl ha kb l c

d d2

30

31

32

which was obtained using:


d hkl a ha a kb a l c a h d hkl b ha b kb b l c b k

33

d hkl

c ha c k b c l c c l

Comparing Equation (32) with Equation (13) proves the identity of d* hkl and the reciprocal lattice vectorh hhkl. Braggs equation, Equation (26), can be restated as: h hhkl 34

In other words, diffraction occurs whenever the scattering vector h equals a reciprocal lattice vector hhkl. This powerful result is visualized in the useful Ewald construction that is described below. Useful equivalent variations of the Bragg equation are: jhj js s0 j and: jQj 4p sin y 2p l d 36 2 sin y 1 l d 35

The vector Q is the physicists equivalent of the crystallographers h. The physical meaning of Q is the momentum transfer on scattering and differs from the scattering vector h by a factor of 2p.

Principles of Powder Diffraction 1.5 THE EWALD CONSTRUCTION

11

The Bragg equation shows that diffraction occurs when the scattering vector equals a reciprocal lattice vector. The scattering vector depends on the geometry of the experiment whereas the reciprocal lattice is determined by the orientation and the lattice parameters of the crystalline sample. Ewalds construction combines these two concepts in an intuitive way. A sphere of radius 1/l is constructed and positioned in such a way that the Bragg equation is satised, and diffraction occurs, whenever a reciprocal lattice point coincides with the surface of the sphere (Figure 1.8). The recipe for constructing Ewalds spherey is as follows (Figure 1.8): 1. Draw the incident wave vector s0. This points in the direction of the incident beam with length 1/l. 2. Draw a sphere centered on the tail of this vector with radius 1/l. The incident wave vector s0 denes the radius of the sphere. The scattered wave vector s, also of length 1/l, points in a direction from the sample to the detector. This vector is drawn also starting from the center of the sphere and also terminates at a point on the surface. The scattering vector h s s0 completes the triangle from the tip of s to the tip of s0, both lying on the surface of the sphere. 3. Draw the reciprocal lattice with the origin lying at the tip of s0. 4. Find all the places on the surface of the sphere, where reciprocal lattice points lie. This construction places a reciprocal lattice point at one end of h. By definition, the other end of h lies on the surface of the sphere. Thus, Braggs law is only satised, when another reciprocal lattice point coincides with the surface of the sphere. Diffraction is emanating from the sample in these directions. To detect the diffracted intensity, one simply moves the detector to the right position. Any vector between two reciprocal lattice points has the potential to produce a Bragg peak. The Ewald sphere construction additionally indicates which of these possible reections satisfy experimental constraints and are therefore experimentally accessible. Changing the orientation of the crystal reorients the reciprocal lattice bringing different reciprocal lattice points on to the surface of the Ewald sphere. An ideal powder contains individual crystallites in all possible orientations with equal probability. In the Ewald construction, every reciprocal lattice point is smeared out onto the surface of a sphere centered on the origin of reciprocal space. This is illustrated in Figure 1.9. The orientation of the d* hkl vector is lost and the three-dimensional vector space is reduced to one dimension of the modulus of the vector d* hkl. These spherical shells intersect the surface of the Ewald sphere in circles. Figure 1.10 shows a two-dimensional projection. Diffracted beams emanate
y

For practical reasons, plots of the Ewald sphere are circular cuts through the sphere and the corresponding slice of reciprocal space.

12

Chapter 1

Ewald sphere

1/

s0

d*=

s-s

0
Figure 1.8
Geometrical construction of the Ewald circle. The 0 marks the origin of reciprocal space. The vectors are dened in the text.

from the sample in the directions where the thin circles from the smeared reciprocal lattice intersect the thick circle of the Ewald sphere. A few representative reected beams are indicated by the broken lines. The lowest d-spacing reections accessible in the experiment are determined by the diameter of the Ewald sphere 2/l. To increase the number of detectable reections one must decrease the incident wavelength. In the case of an energy dispersive experiment such as time-of-ight neutron powder diffraction, which makes use of a continuous distribution of wavelengths from lmin to lmax at xed angle, all smeared out cones between the two limiting Ewald spheres can be detected. In three dimensions, the circular intersection of the smeared reciprocal lattice with the Ewald sphere results in the diffracted X-rays of the reection hkl forming coaxial cones, the so-called DebyeScherrer cones (Figure 1.11). The smearing of reciprocal space in a powder experiment makes the measurement easier but results in a loss of information. Reections overlap from lattice planes whose vectors lie in different directions but which have the same d-spacing. These cannot be resolved in the measurement. Some of these overlaps are dictated by symmetry (systematic overlaps) and others are accidental. Systematic overlaps are less serious for equivalent reections [e.g., the six Bragg peaks (100), ( 100), (010), . . . from the faces of a cube] since the multiplicity is known from the symmetry. For highly crystalline samples, accidental overlaps can be reduced by making measurements

Principles of Powder Diffraction

13

Figure 1.9

Illustration of the reciprocal lattice associated with a single-crystal lattice (left) and a large number of randomly oriented crystallites (right). A real powder consists of so many grains that the dots of the reciprocal lattice form into continuous lines.

with higher resolution, or by taking data at different temperatures in an attempt to remove the overlap by differential thermal expansion of different cell parameters. To obtain the maximum amount of information, a spherical shell detector would be desirable, though currently impractical. Often, a at two-dimensional detector, either lm, image plate, or CCD is placed perpendicular to the direct beam. In this case, the DebyeScherrer cones appear as circles as shown in Figure 1.12a. For an ideal powder, the intensity around the rings is isotropic. Conventional powder diffraction measurements, e.g., BraggBrentano geometry, take one-dimensional cuts through the rings, either horizontally or vertically depending on the geometry of the diffractometer. If the full rings, or fractions of them, are detected with two-dimensional detectors the counting statistics can be improved by integrating around the rings. If the powder is non-ideal, the ring intensity is no longer uniform, as illustrated in Figure 1.12b, giving arbitrary intensities for the reections in a one-dimensional scan. To improve powder statistics, powder samples are generally rotated during measurement. However, the intensity variation around the rings can give important

14
re d > = /2 sphe g itin

Chapter 1

m Li

Prim

ary bea m

t Reflec

ed be

ams

Figure 1.10

Illustration of the region of reciprocal space that is accessible in a powder measurement. The smaller circle represents the Ewald sphere. As shown in Figure 1.9, in a powder the reciprocal lattice is rotated to sample all orientations. An equivalent operation is to rotate the Ewald sphere in all possible orientations around the origin of reciprocal space. The volume swept out (area in the gure) is the region of reciprocal space accessible in the experiment.
hkl
Refl ecte d be am

Primary beam

hkl
Ref lect ed b eam

hkl

Primary beam

hkl

Figure 1.11

Comparison between the scattered beams originating from a single crystal (top) and a powder (bottom). For the latter, some Debye Scherrer cones are drawn.

Principles of Powder Diffraction

15

Figure 1.12

DebyeScherrer rings from an ideal ne grained (a, left) and a grainy (b, right) powder sample.

information about the sample such as preferred orientation of the crystallites or texture. 1.6 TAKING DERIVATIVES OF THE BRAGG EQUATION

Several important relationships in crystallography directly follow from a derivative of the Bragg equation [Equation (5)]. First we rewrite Braggs law making the d-spacing the subject of the equation: d nl 2 sin y 37

The uncertainty of the measured lattice spacing is given by the total derivative dd, which can be written according to the chain rule as: dd leading to: dd and nally: dd dy dl d tan y l 40 nl cosy n dy dl 2 sin y sin y 2 sin y 39 @d @d dy dl @y @l 38

This equation allows us to discuss several physically important phenomena.

16

Chapter 1

Figure 1.13

Percentage error in measured d-spacing as a function of scattering angle arising from a constant angular misalignment of DY for a well aligned (0.0011), a typically aligned, (0.011) and a poorly aligned (0.051) diffractometer.

When a crystal is strained, the d-spacings vary. A macroscopic strain changes the interplanar spacing by Ddhkl, giving rise to a shift in the average position of the diffraction peak of Dy, while microscopic strains give a distribution of d-spacings Ddhkl which broaden the peak by dy. This is discussed in detail in Chapters 12 and 13. A constant angular oset due to misalignment of the diffractometer gives rise to a nonlinear error in our determination of dhkl, disproportionately affecting low angle reections (Figure 1.13). Similarly, our ability to resolve two partially overlapping reections separated by Ddhkl is limited by the nite angular resolution Dy of the diffractometer. There are many geometrical contributions to the angular resolution (e.g., angular width of the receiving slit in front of the detector). Another contribution comes from nite wavelength spread of the incident beam Dl. From Equation (40) we get the angular dispersion to be: dy tan y dl l 41

This is plotted in Figure 1.14, which shows that the resolution due to a nite spread in l is decreasing at higher angles. In a real experiment the angle dependence of the resolution function can be complicated. In traditional modeling programs the Bragg-peak line shapes are modeled using empirical line-shape functions. More recently, approaches have been developed that explicitly account for the different physical processes that result in the line

Principles of Powder Diffraction

17

Figure 1.14

Angular dependence of the intrinsic peak width (resolution function) of the diffractometer due to the wavelength spread between Cu-Ka1 and Cu-Ka2 (about 12 eV).

shapes. This is called the fundamental parameters approach and is described in Chapters 5, 6 and 13. 1.7 BRAGGS LAW FOR FINITE SIZE CRYSTALLITES

Assuming an innite stack of lattice planes, Braggs equation gives the position of delta-function Bragg peaks. Finite size crystallites give rise to Bragg peaks of nite width. This size broadening is described by the Scherrer equation. We now reproduce the simple derivation following Klug and Alexander (1974; see Bibliography). Figure 1.15 shows the path length difference versus the depth of the lattice plane. When the angle between the incoming beam and the lattice plane Y is different by an amount e from the Bragg condition, it is always possible to nd a lattice plane inside the crystal where the extra path is D l/2 producing destructive interference. For a thick crystal this is true for arbitrarily small e, which explains the sharp Bragg reections. For a crystal with nite dimensions, for small e the plane for which D n 1 2l holds will not be reached. In this case there is not perfect cancellation of the intensity away from the Bragg condition, thus leading to an intensity distribution over some small angular range. We can use this idea to estimate the broadening of a Bragg reection due to size effects. The thickness of a crystallite in the direction perpendicular to p (hkl) planes of separation dhkl (Figure 1.15) is: Lhkl pdhkl 42

18
= n( + )

Chapter 1

= 0

2
2

1 2 3 4

p-1 p

Figure 1.15

Path length difference of the scattered ray versus the depth of the lattice plane in the crystal.

The additional beam path between consecutive lattice planes at the angle y + e is: D 2d siny e 2d sin y cos e cos y sin e nl cos e sin e2d cos y E nl sin e2d cos y The corresponding phase difference is then: dj 2p D 4p 4ped cos y 2pn ed cos y l l l 44

43

The phase difference between the top and the bottom layer, p is then: dj p 4ped cos y 4pLhkl e cos y l l 45

Rearranging Equation (45) leads to: e ldj 4pLhkl cos y 46

which gives an expression for the misalignment angle in terms of the crystallite size Lhkl and the phase difference dj between the reections between the top and the bottom plane. Clearly, the scattered intensity is at a maximum for dj 0 (e 0). With increasing e the intensity decreases giving rise to a peak of nite width. Perfect cancellation of the top and bottom waves occurs a phase difference of dj p at which point e l/(4Lhkl cosy). On a measured 2y-scale the

Principles of Powder Diffraction measured angular width between these points is: bhkl 4e l Lhkl cos y

19

47

giving us some measure of the peak width in radians due to the nite particle size. A full treatment taking into account the correct form for the intensity distribution gives: bhkl Kl Lhkl cos y 48

with a scale factor of K 0.89 for perfect spheres. In general K depends on the shape of the grains (e.g., K is 0.94 for cubic shaped grains) but is always close to unity. This equation is not valid for crystallitesz that are too large or too small. With large crystallites the peak width is governed by the coherence of the incident beam and not by particle size. For nano-scale crystallites, Braggs law fails and needs to be replaced by the Debye equation (see Chapter 16). BIBLIOGRAPHY 1. V. K. Pecharsky and P. Y. Zavalij, Fundamentals of Powder Diffraction and Structural Characterization of Materials, Kluwer Academic Publishers, Boston, 2003. 2. D. L. Bish, J. E. Post (Eds.), Modern Powder Diffraction, Reviews in Mineralogy Volume 20, Mineralogical Society of America, Chantilly VA, 1989. 3. H. P. Klug and L. E. Alexander, X-ray Diffraction Procedures, 2nd edn., John Wiley & Sons, New York, 1974.

Strictly speaking, the term crystallite size here refers to the dimension of a coherently scattering domain. Only in a perfect crystal, is this the grain size.

Potrebbero piacerti anche