<<

Reciprocal

From Quantum Mechanics we know that symmetries have kinematic im- plications. Each symmetry predetermines its specific quantum numbers and leads to certain constraints (conservation laws/selection rules) expressed in terms of these quantum numbers. In this chapter, we consider kinematic consequences of the discrete translation symmetry. The shortest way to in- troduce the relevant quantities—the vectors of reciprocal lattice—is to con- sider the for crystal functions.

Fourier analysis for crystal functions

Let A(r) be some function featuring discrete translation symmetry:

A(r + T) = A(r) , (1) where T is any translation vector of a certain translation group GT . We want to expand A(r) in a of functions respecting translation symmetry (i.e. obeying translation symmetry with the same group GT ). Without loss of generality, we confine the coordinate r to a of the of the group GT . Hence, we work with the of functions defined within the unit cell and satisfying the periodicity condition (1). Within this vector space, we want to construct the orthogonal basis of functions that would be eigenfunctions of continuous translations. These functions are the plane waves eiG·r with the wavevector G constrained by the condition (1). Clearly, the constraint reduces to

G · T = 2π × integer (for any G and T) . (2)

From (2) we immediately see that the set of vectors G is a group, GG, with the vector addition as a group operation. The group GG can be interpreted as a group of translations and visualized with corresponding Bravais lattice referred to as reciprocal lattice. Noticing that the condition (2) is symmetric with respect to the vectors T and G, we conclude that the Bravais lattice of the group GT (referred to in this context as the direct lattice) is reciprocal with respect to the Bravais lattice of the group GG. In other words, the two lattices are mutually reciprocal. A simple way of proving the orthogonality of the plane waves, Z i(G1−G2)·r dV e = Vc δG1G2 , (3) cell

1 is to make sure that they are the eigenfunctions of a certain Hermitian (dif- ferential) operator with a discrete non-degenerate spectrum.1 For example, the operator i∇ does the job, because for the functions defined within the primitive cell and featuring the translation invariance (1) the operator ∇ is anti-Hermitian—checked by doing the inner-product integral by parts2— and i∇eiG·r = −GeiG·r, meaning that different plane waves have different eigenvalues. It is convenient to include the factor 1/Vc in the definition of the inner product of functions: Z ∗ hA|Bi = (1/Vc) A (r)B(r) dV . (4) cell In this case, the set of the functions

|Gi ≡ eiG·r (5) becomes normalized, and we arrive at the Fourier series in the form

X X iG·r A(r) = hG|Ai|Gi ≡ AG e , (6) G G where Z −iG·r AG ≡ hG|Ai = (1/Vc) e A(r) dV . (7) cell

Talking of an explicit evaluation of the Fourier coefficients AG we note that any integral over the primitive cell can be parametrized in terms of the coordinates—in a general case, non-Cartesian ones—with respect to a given primitive set (a1, a2, a3):

r = ξ1a1 + ξ2a2 + ξ3a3 , ξ1, ξ2, ξ3 ∈ [0, 1] , (8)

Z Z 1 Z 1 Z 1 (1/Vc) dV (...) = dξ1 dξ2 dξ3 (...) . (9) cell 0 0 0

The volume Vc cancels with the Jacobian of the transformation of integration variables. 1This will also guarantee the completeness of the set. 2Since the integral is over a finite region, the surface term appears, and one has to show that its contribution is zero. This is conveniently done by extending the integration over a large cluster consisting of N  1 adjacent cells. By periodicity, the bulk integral then equals to N times the integral over a single cell, while the surface integral, if non-zero, has to grow with N as N to a power smaller than 1. Since N can be arbitrarily large, this proves that the surface contribution is identically zero.

2 Primitive set for reciprocal lattice

To find explicit expressions for the vectors of the reciprocal lattice we need to construct a primitive set. It turns out that there is a one-to-one correspondence between primitive sets of the direct and reciprocal lattices. To reveal this correspondence, let us take a primitive set (a1, a2, a3) of the direct lattice and construct the following three vectors:

c1 = a2 × a3 , c2 = a1 × a3 , c3 = a1 × a2 . (10)

By construction, the three vectors are not coplanar, and thus any vector G can be expressed, in a unique way, as their linear combination:

G = η1c1 + η2c2 + η3c3 . (11)

Multiplying both sides of (11) by aj (j = 1, 2, 3) and applying (2), we see that ηj = (2π/Vc) × integer (j = 1, 2, 3) . (12) Furthermore, the requirement (12) straightforwardly implies (2), and we conclude that (12) and (2) are equivalent. This yields a primitive set (b1, b2, b3) of the reciprocal lattice:

a2 × a3 b1 = 2π , (13) a1 · (a2 × a3)

a1 × a3 b2 = 2π , (14) a2 · (a1 × a3)

a1 × a2 b3 = 2π . (15) a3 · (a1 × a2)

Each G is then associated with the three integers (m1, m3, m3):

G = m1b1 + m2b2 + m3b3 . (16)

The 2D case readily follows from the 3D one by formally introducing a3 =z ˆ, wherez ˆ is the unit vector perpendicular to the xy plane of the primitive set (a1, a2). The relevant vectors of the reciprocal lattice are (b1, b2), since both lie in the xy plane. The vector b3 is perpendicular to the xy plane and thus can be safely neglected.

3 The symmetry of the correspondence between the primitive sets of the direct and reciprocal lattices becomes obvious by noticing that the explicit relations (13)-(15) can be written in an equivalent implicit form

ai · bj = 2πδij . (17)

With equations (17) it is easy to see a simple relation between the volume (rcpr) Vc of the primitive cell of the direct lattice and the volume Vc of the primitive cell of the reciprocal lattice:

(rcpr) d Vc = (2π) /Vc . (18)

Indeed, write (17) in the matrix form (we do it for d = 3),       a1x a1y a1z b1x b2x b3x 2π 0 0        a2x a2y a2z   b1y b2y b3y  =  0 2π 0  , a3x a3y a3z b1z b2z b3z 0 0 2π and take the determinant of both sides recalling that the determinant of a product of matrices equals to the product of determinants. For crystallographic purposes, it is convenient to work with dimension- less vectors of direct and reciprocal lattices, T˜ and G˜ , respectively, intro- duced as T = aT˜ , G = (2π/a) G˜ , (19) where a is a convenient unit of length, normally the length of one of the vectors of the primitive set. For the dimensionless lattices, the relations (2) and (17) simplify to

G˜ · T˜ = integer (for any G˜ and T˜ ) , (20)

a˜i · b˜j = δij . (21) Dimensionless primitive sets are very convenient for establishing the types of the reciprocal lattices. As an instructive example, see the calculation in Kittel’s book establishing the fact that fcc and bcc lattices are reciprocal with respect to each other.

4 Working with non-primitive sets and unit cells. Symmetry factors

In cases when the Bravais lattice features rectangular symmetries—fcc and bcc lattices being most typical and important 3D examples3—it is very convenient to use non-primitive unit cells with corresponding non-primitive sets of translation vectors. The advantage is the simplification of geometric description and nomenclature. For example, for fcc and bcc lattices we simply start with geometry and nomenclature of simple cubic lattice and then add extra lattice points converting the sc lattice into fcc or bcc one. Yet another alternative is to start with some simple lattice (say, simple cubic lattice) and erase some points to arrive at a less trivial Bravais lattice. (Recall obtaining the sodium chloride lattice from the simple cubic one by using alternating coloring.) It turns out that the two approaches—using non-primitive unit cells and erasing certain lattice points are dual in the following sense. If a direct lattice is parameterized with a non-primitive unit cell, then the reciprocal lattice can be naturally obtained as a lattice reciprocal to the Bravais lattice of non-primitive units of the direct lattice, upon applying the constraint that some points of this ‘fake’ reciprocal lattice should be removed to produce the actual reciprocal lattice; and there is a general prescription of how to calculate the coefficients—the so-called symmetry factors—explicitly telling us which points of the ‘fake’ reciprocal lattice are fake and which are real. Below we develop corresponding tools. We start with formalizing the problem. We are given a direct Bravais lattice associated with the group of translations GT consisting of translation vectors T. By introducing a non-primitive unit cell of our lattice GT (from now on we use ‘lattice GT ’ as a short version of ‘Bravais lattice associated with the group GT ’), we actually introduce a new group of translations, GT 0 , consisting of the allowed translations vectors T0 of the non-primitive unit 0 cell. Clearly, each T ∈ GT 0 also belongs to GT , that is the group GT 0 is a subgroup of GT : GT 0 ⊂ GT . (22)

The two reciprocal lattices, GG and GG0 , are defined by the conditions

G · T = 2π × integrer (for any T ∈ GT ) , (23)

3In 2D, the analog is the centered rectangular lattice. And also the hexagonal lattice which in the context of the present discussion can be treated as nothing but a special case of the centered rectangular lattice.

5 0 0 0 G · T = 2π × integrer (for any T ∈ GT 0 ) . (24) The property (22) guaranties that any G satisfying (23) and thus belonging to GG belongs also to GG0 , by (24). Hence,

GG ⊂ GG0 . (25)

In plain words, Eqs. (22) and (25) say that while the lattice GT is obtained from the lattice GT 0 by adding extra points to the basis, the lattice GG is obtained from the lattice GG0 by removing certain lattice points. Not a bad deal! Instead of directly constructing reciprocal to fcc or bcc lattices, we work with corresponding simple cubic lattice GT 0 that readily yields GG0 . The latter is also a simple cubic lattice; with the straightforward relationship between the primitive sets:

0 0 2πaj bj = 0 2 (for a cubic unit cell) . (26) |aj| A price that we have to pay for the luxury is that we need to figure out which points of the lattice GG0 have to be removed to produce GG. This is conveniently done with the symmetry factors SG0 such that

( 0 1 , G ∈ GG , SG0 = 0 (27) 0 , G ∈/ GG .

An easy way to establish explicit expressions for the symmetry factors is to 0 compare Fourier coefficients FG and FG0 —corresponding to the translation groups GT and GT 0 , respectively—for some periodic function F(r + T) = F(r). The simplest F(r) is given by delta-functions centered on the sites of the Bravais lattice:

ν0 X X X 0 F(r) = δ(r − T) ≡ δ(r − T − Rν) . (28) T T0 ν=1 Here the identity represents the same function in terms of the nomenclature of the lattice GT 0 , in which case there are ν0 points in the unit cell; we enumerate them with the subscript ν, and characterize their positions by corresponding vectors Rν (for definiteness, we select R1 = 0). Doing the integrals (7), we find

ν0 0 1 0 1 X −iG ·Rν F = , F 0 = e , (29) G V G V 0 c c ν=1

6 0 where Vc and Vc are the volumes of primitive cells of the lattices GT and 0 GT 0 , respectively (note that Vc = ν0Vc). Comparing the two Fourier series,

X iG·r X iG0·r 0 e FG ≡ F(r) ≡ e FG0 , (30) G G0 we see that

ν0 ( 0 0 0 1 X −iG ·Rν 0 FG = 1/Vc , G ∈ GG , e ≡ F 0 = (31) V 0 G 0 , G0 ∈/ G , c ν=1 G and conclude that ν0 0 1 X −iG ·Rν S 0 = e . (32) G ν 0 ν=1 Let us evaluate the symmetry factors for bcc and fcc lattices, both rep- resented with the simple-cubic lattice GT 0 , for which the primitive set is (bcc) given by unit Cartesian vectors, (ˆx, y,ˆ zˆ). The vectors Rν are [ν0 = 2 and (fcc) ν0 = 4]: 1 1 1 R(bcc) = , , , (33) 2 2 2 2 1 1  1 1  1 1 R(fcc) = , , 0 , R(fcc) = , 0, , R(fcc) = 0, , . (34) 2 2 2 3 2 2 4 2 2

The lattice GG0 , reciprocal to the GT 0 lattice is sc, with the primitive set 0 (2πx,ˆ 2πy,ˆ 2πzˆ), so that G = 2π(m1xˆ + m2yˆ + m3zˆ). For the symmetry factors we then get h i (bcc) −iπ(m1+m2+m3) Sm1m2m3 = 1 + e /2 , (35) h i (fcc) −iπ(m2+m3) −iπ(m1+m3) −iπ(m1+m2) Sm1m2m3 = 1 + e + e + e /4 . (36)

(bcc) It is straightforward to see that non-zero Sm1m2m3 corresponds to a fcc (fcc) lattice, while non-zero Sm1m2m3 corresponds to a bcc lattice.

Bragg diffraction: Basic principles

Consider a problem of elastic perturbative scattering of a quantum par- ticle from a crystal. We remind that elastic means that the state of the crystal does not change after the scattering. Perturbative means that the

7 probability of the process is much smaller than unity, so that we can rely on the perturbation theory (Fermi Golden Rule). For our theoretical pur- poses, the specific type of the particle does not matter. In practice, very important are X-ray quanta (photons), neutrons, and electrons, with (de Broglie) wavelength comparable to the crystal lattice period. Let the initial state of the quantum particle be the plane wave with the wavevector k. We are interested in the kinematic structure of the probability amplitude Mkk0 of finding the particle in the plane-wave state with the momentum k0 as a result of the scattering event. By kinematic structure we mean selection rules imposed on k and k0 by the periodicity of the crystal. These selection rules, known as Bragg conditions, deal only with the direction of k0 (as well as the direction of k), but not its absolute value, since k0 = k by elasticity of the process. The expression for the probability amplitude is (r being the position of the particle)

Mkk0 = h final state | Uint | initial state i , (37) where | initial state i = |ki | crystal i , |ki ≡ eik·r , (38) 0 | final state i = |k0i | crystal i , |k0i ≡ eik ·r , (39) R d and Uint ≡ U(r) d r is the Hamiltonian of interaction between the particle and the crystal; the ket | crystal i stands for the wavefunction of the crystal (the same in the initial and final states). Hence, Mkk0 can be written as the integral over the crystal, Z −iq·r Mkk0 = e A(r) dV , (40) crystal with q = k0 − k , (41) and A(r) = h crystal | U(r) | crystal i . (42) In terms of the general kinematic constraints we are revealing, the details of the form of A(r) are irrelevant, the only crucial property being the transla- tion symmetry, Eq. (1), that A(r) has to respect. Nevertheless, it is worth mentioning that for the X-ray scattering, A(r) amounts to the expectation value of the density of electrons at the point r times the coupling constant.

8 In view of the periodicity of A(r), the integral (40) over the crystal reduces to the sum of the integral over the primitive cell with exponential pre-factors: (crystal) X −iq·Tn Mkk0 = Sq e , (43) n Z −iq·r Sq = e A(r) dV . (44) cell n=0

From (43) we see that Mkk0 is sharply peaked when

q ≡ k0 − k = G (Bragg condition) . (45)

Indeed, in this case, all the exponentials interfere constructively and

Mkk0 = SG Ncells , (46) where Ncells is the (macroscopically large) number of primitive cells in the crystal. If q is not macroscopically close to one of the vectors G, many exponentials in (43) interfere destructively and the value of Mkk0 is much smaller than (46). For the wavevectors k and k0 satisfying the Bragg con- dition (45), the scattering amplitude is linearly proportional to Ncells. In accordance with the Golden Rule, the intensity of scattering—that is the probability density of having a scattering event with given k and k0—scales 2 as the square of the scattering amplitude, and thus is ∝ Ncells; this is the manifestation of enhancement of quantum-mechanical elementary processes by constructive interference of the amplitudes.

Alternative forms and interpretations of Bragg conditions. Ewald sphere and Brillouin zones

By conservation of energy we have k0 = k, and the complete set of equations expressing the kinematic constraints of the Bragg diffraction read

k0 = k , k0 − k = G (Bragg conditions, basic version) . (47)

We have one and one vector constraints. Mathematically, this is equivalent to a system of four scalar constraints. The constraints can be re-written in the following two important forms

k · G = G2/2 , k0 = k − G (Bragg conditions, version A) , (48)

9 k0 · G = G2/2 , k = k0 − G (Bragg conditions, version B) . (49) (These alternative forms of Bragg conditions are easily derived by taking squares of k0 = k − G and k = k0 − G, respectively. Note that the sign of G does not matter since −G ∈ GG.) Geometrically, any scalar constraint on a three-dimensional vector yields a surface (a system of surfaces) in corresponding vector space. The scalar constraint k0 = k means that the end of the vector k0 has to be in the sphere of the radius k in the wavevector space. To visualize Eqs. (47), we draw this sphere together with the properly positioned reciprocal lattice, as shown in K-Fig. 8. This geometric construction is known as Ewald construction, in the context of which the sphere k0 = k is referred to as Ewald sphere. With the Ewald construction we clearly see that Bragg diffraction requires a fine-tuning: It happens only if, in addition to the origin of the reciprocal lattice that belongs to the Ewald sphere by construction, there is yet another reciprocal lattice point on the sphere. The necessity of fine-tuning is also seen with the first equation of Eqs. (48), imposing a scalar constraint on the vector k. To reveal the surfaces standing behind this constraint, rewrite it as

k · eˆG = G/2 , (50) wheree ˆG = G/G is the unit vector in the direction of G. We see that the ends of the vectors k satisfying the Bragg conditions live in special planes that are perpendicular bisectors of the reciprocal lattice vectors (for an illustration, see K-Fig. 9). There are infinitely many such planes.4 And these divide the wavevector space into infinitely many pieces called Brillouin zones. Of crucial importance is the first —the one contain- ing the origin of the reciprocal lattice. By construction it is nothing but the Wigner-Seitz cell of the reciprocal lattice. Now note that if we know the Wigner-Seitz cell, we can immediately restore the whole lattice by re- flecting the origin against the boundaries, and then applying the procedure iteratively. Having restored the reciprocal lattice, we can pick up one of its primitive sets of translation vectors and construct corresponding set of the direct lattice. We thus arrive at the fundamental conclusion that Bragg scattering allows one to measure the Bravais lattice of the . There are two equivalent ways of measuring the shape of the first Bril- louin zone. The above-discussed one deals with the condition (50) for the incident wave. Equation (49), identical to (48) up to replacement k ↔ k0,

4Recall that those planes are at the heart of the Voronoi-Wigner-Seitz construction!

10 tells us that the same goal can be achieved by working with k0, in which case one reveals the first Brillouin zone of Bragg-scattered waves. For example, think of the experimental protocol when the incident particle beam is colli- mated but not monochromatic, while the detector measures the magnitude of k0 for a fixed direction. This way one obtains a discrete sequence of points of the reciprocal space, each point corresponding to the intersection of the chosen direction axes with the boundary of each Brillouin zone.

Structure factor and atomic form-factors

In the Bragg diffraction context, the periodic function A(r) often comes in the form of atomic decomposition, X A(r) = fj(r − rj) , (51) j

5 where rj is the position of the j-th atom in the infinite space. Clearly, there is only a finite number of different functions fj(r), forming sub-lattices of the global lattice; it is sufficient to consider just one of them (and then add similar contributions from the others). For a single sub-lattice we have fj(r) ≡ f(r), and that is what will be assumed from now on. In this case, along with the Bragg interference responsible for the Bragg diffraction, there can take place additional interference within a primitive cell. This additional interference can either enhance or suppress the scatter- ing amplitude, depending on G. Below we develop corresponding theory. Given that diffraction peaks are macroscopically sharp while the function Sq defined by Eq. (44) is a smooth function of q, we naturally confine our consideration with q = G. In the Bragg spectroscopy, SG is called the . We have (the utility of the integration over the whole crystal volume V will become clear later) Z Z −iG·r 1 −iG·r SG = e A(r) dV ≡ e A(r) dV , (52) cell n=0 Ncells V and comparing this to (7) we see that the structure factor is nothing but the Fourier coefficient of A (up to the normalization of the latter which is a matter of convention): SG = Vc AG . (53)

5To clearly see the origin of Eq. (51), think, e.g., of a neutron interacting with each P (j) nucleus individually: Uint(r) = j Uint (r − rj ).

11 Now we plug (51), with fj = f, into the second integral of (52) and— understanding that the integration over the whole crystal volume of an essen- tially localized function f can be upgraded to the infinite-space integration— obtain fG X S = e−iG·rj , (54) G N cells j with Z −iG·r fG = e f(r) dV (over all the space !) . (55)

The quantity fG is called atomic form-factor. The expression (54) is not yet the final answer. It can be further simplified by taking into account the translational symmetry. Let us take j ≡ (n, ν), where n labels the cell (corresponding to the basis translated by Tn) and ν enumerates the atoms within the cell. (At this point it is worth noting that the treatment of this section does not imply that the cell is necessarily primitive.) Then

rj = Tn + Rν , (56) where Rν is the position of the ν-th atom in the basis cell, and

ν0 ν0 ν0 X −iG·rj X X −iG·(Tn+Rν ) X X −iG·Rν X −iG·Rν e = e = e = Ncells e . j n ν=1 n ν=1 ν=1

This brings us to the final answer

ν0 X −iG·Rν SG = fG e . (57) ν=1 We cannot but notice that the form of Eq. (57) is reminiscent of that of the symmetry coefficients, Eq. (32). Apart from the pre-factor fG, the most important qualitative distinction is that SG’s do not have to be zero for certain G’s. Indeed, if we are dealing with a primitive cell containing more than one atom (of the same sort or of different sorts), we will still have ν0 > 1 and thus summation over positions in (57), but no extra symmetries leading to exact cancellations of the terms for certain special G’s. Additivity of the structure factor with respect to different sorts of atoms allows us to immediately generalize Eq. (57) to the case when there are different fj’s: ν0 X (ν) −iG·Rν SG = fG e , (58) ν=1

12 the label ν standing now not only for the position of the atom in the primitive cell, but also reflecting the sort of the atom.

Width of Bragg peaks

Our next goal is to understand the width of the Bragg peaks. We are mainly interested in the scaling of the peak width with the system size. The details of the form of the peak depend on the particular overall shape of the crystal and thus are not very instructive. We select the most con- venient shape, which is the one prompted by the shape of the primitive parallelepiped. In this case (choosing the cell n = 0 to be in the corner of the crystal), (crystal) L1−1 L2−1 L3−1 X (...) ≡ X X X (...) , (59) n n1=0 n2=0 n3=0 where Lj is the number of primitive cells per translational dimension j (L1L2L3 = Ncells), which is very convenient for evaluating the sum in (43) by reducing it to the product of three similar and simple factors,

(crystal) X −iq·Tn e = Q(χ1,L1) Q(χ2,L2) Q(χ3,L3)(χj = q · aj) , (60) n that are nothing but partial sums of geometric series:

L−1 1 − e−iχL sin(χL/2) Q(χ, L) = X e−iχn = = e−iχ(L−1)/2 . (61) 1 − e−iχ sin(χ/2) n=0 The exponential pre-factor is irrelevant since the probability density for the scattering is given by the square of the absolute value of the scattering amplitude:

2 2 |Mkk0 | = |Sq| I(χ1,L1) I(χ2,L2) I(χ3,L3) , (62)

sin2(χL/2) I(χ, L) = |Q(χ, L)|2 = . (63) sin2(χ/2) The function I(χ, L) is a 2π-periodic function of χ, with very sharp (at L  1) peaks of the amplitude L2 and the width ∆χ ∼ 2π/L, see Fig. 1. Recalling that χj = q · aj, we conclude that, in the wavevector units, the width of the peak is on the order of the inverse linear system size.

13 Figure 1: The function I(χ, L), Eq. (63), in the vicinity of one of its narrow (at L  1) peaks. The amplitude at the maximum is I = L2, the width of the peak is ∆χ ∼ 2π/L.

Note that there are also satellite peaks of the same width and progres- sively decreasing amplitude. The shape of the function I(χ, L) in the vicinity of the peak is readily revealed by expanding the sine in the denominator: sin2 χ/2 ≈ χ2/4 at |χ|  1:

sin2(χL/2) I(χ, L) = |Q(χ, L)|2 = 4 . (64) χ2 With this formula we see that the amplitude of the m-th satellite peak decreases with m as 1/m2.

14