In condensed matter physics and crystallography, the static structure factor (or structure factor for short) is a mathematical description of how a material scatters incident radiation. The structure factor is a critical tool in the interpretation of scattering patterns (interference patterns) obtained in X-ray, electron and neutron diffraction experiments.
Confusingly, there are two different mathematical expressions in use, both called 'structure factor'. One is usually written
S(q)
F
Fhk\ell
(hk\ell)
(hk\ell)
Fhk\ell
S(q)
S(q)
Fhk\ell
|Fhk\ell|2
Fhk\ell
S(q)
The static structure factor is measured without resolving the energy of scattered photons/electrons/neutrons. Energy-resolved measurements yield the dynamic structure factor.
Consider the scattering of a beam of wavelength
λ
N
styleRj,j=1,\ldots,N
q
q=
ks |
-
ko |
ks |
ko |
|
ks |
|=
|k0| |
=2\pi/λ
\theta
|ks|=
|ko| |
q=|q|={
4\pi | |
λ |
\sin(\theta/2)}
q
\Psis(q)=
N | |
\sum | |
j=1 |
fj
-iq ⋅ Rj | |
e |
For an assembly of atoms,
fj
j
The structure factor is defined as this intensity normalized by
N | |
1/\sum | |
j=1 |
2 | |
f | |
j |
If all the atoms are identical, then Equation becomes
I(q)=f2
N | |
\sum | |
j=1 |
N | |
\sum | |
k=1 |
-iq ⋅ (Rj-Rk) | |
e |
N | |
\sum | |
j=1 |
2 | |
f | |
j |
=Nf2
Another useful simplification is if the material is isotropic, like a powder or a simple liquid. In that case, the intensity depends on
q=|q|
rjk=|rj-rk|
An alternative derivation gives good insight, but uses Fourier transforms and convolution. To be general, consider a scalar (real) quantity
\phi(r)
V
style\psi(q)=\intV\phi(r)\exp(-iq ⋅ r)dr
q
style\psi(q)
N
f(r)
with
styleRj,j=1,\ldots,N
style\psi(q)=f(q) x
N | |
\sum | |
j=1 |
\exp(-iq ⋅ Rj)
This is clearly the same as Equation with all particles identical, except that here
f
q
In general, the particle positions are not fixed and the measurement takes place over a finite exposure time and with a macroscopic sample (much larger than the interparticle distance). The experimentally accessible intensity is thus an averaged one
style\langleI(q)\rangle
\langle ⋅ \rangle
In a crystal, the constitutive particles are arranged periodically, with translational symmetry forming a lattice. The crystal structure can be described as a Bravais lattice with a group of atoms, called the basis, placed at every lattice point; that is, [crystal structure] = [lattice]
\ast
q
In principle the scattering factor
S(q)
I(q)=f2\left|
N | |
\sum | |
j=1 |
-iq ⋅ Rj | |
e |
\right|2
S(q)=
1 | |
N |
\left|
N | |
\sum | |
j=1 |
-iq ⋅ Rj | |
e |
\right|2
The structure factor is then simply the squared modulus of the Fourier transform of the lattice, and shows the directions in which scattering can have non-zero intensity. At these values of
q
f
q
The units of the structure-factor amplitude depend on the incident radiation. For X-ray crystallography they are multiples of the unit of scattering by a single electron (2.82
x 10-15
10-14
The above discussion uses the wave vectors
|k|=2\pi/λ
|q|=4\pi\sin\theta/λ
|s|=1/λ
|g|=2\sin\theta/λ
2\pi
In crystallography, the basis and lattice are treated separately. For a perfect crystal the lattice gives the reciprocal lattice, which determines the positions (angles) of diffracted beams, and the basis gives the structure factor
Fhkl
where the sum is over all atoms in the unit cell,
xj,yj,zj
j
fj
j
xj,yj,zj
a,b,c
a
(hkl)
(ha*,kb*,lc*) |
(hkl)
Fhk\ell
hk\ell
(hxj+kyj+\ellzj)
h
h
Again an alternative view using convolution can be helpful. Since [crystal structure] = [lattice]
\ast
l{F}
l{F}
x l{F}
\propto
x
For the body-centered cubic Bravais lattice (cI), we use the points
(0,0,0)
(\tfrac{1}{2},\tfrac{1}{2},\tfrac{1}{2})
Fhk\ell=\sumjfj
-2\pii(hxj+kyj+\ellzj) | |
e |
=f\left[1+\left(e-i\pi\right)h+k+\ell\right]=f\left[1+(-1)h+k+\ell\right]
Fhk\ell=\begin{cases}2f,&h+k+\ell=even\ 0,&h+k+\ell=odd\end{cases}
The FCC lattice is a Bravais lattice, and its Fourier transform is a body-centered cubic lattice. However to obtain
Fhk\ell
xj,yj,zj=(0,0,0)
xj,yj,zj=\left(
1 | , | |
2 |
1 | |
2 |
,0\right)
\left(0, | 1 | , |
2 |
1 | |
2 |
\right)
\left( | 1 | ,0, |
2 |
1 | |
2 |
\right)
Fhk\ell=f
4 | |
\sum | |
j=1 |
[-2\pii(hxj+kyj+\ellzj)] | |
e |
=f\left[1+e[-i+e[-i+e[-i\right]=f\left[1+(-1)h+(-1)k+(-1)h\right]
with the result
Fhk\ell=\begin{cases}4f,&h,k,\ell allevenorallodd\\ 0,&h,k,\ell mixedparity\end{cases}
The most intense diffraction peak from a material that crystallizes in the FCC structure is typically the (111). Films of FCC materials like gold tend to grow in a (111) orientation with a triangular surface symmetry. A zero diffracted intensity for a group of diffracted beams (here,
h,k,\ell
The diamond cubic crystal structure occurs for example diamond (carbon), tin, and most semiconductors. There are 8 atoms in the cubic unit cell. We can consider the structure as a simple cubic with a basis of 8 atoms, at positions
\begin{align}xj,yj,zj=&(0, 0, 0)&\left(
1 | , | |
2 |
1 | |
2 |
, 0\right) &\left(0,
1 | , | |
2 |
1 | |
2 |
\right)&\left(
1 | , 0, | |
2 |
1 | |
2 |
\right)\ &\left(
1 | , | |
4 |
1 | , | |
4 |
1 | |
4 |
\right)&\left(
3 | , | |
4 |
3 | , | |
4 |
1 | |
4 |
\right) &\left(
1 | , | |
4 |
3 | , | |
4 |
3 | |
4 |
\right)&\left(
3 | , | |
4 |
1 | , | |
4 |
3 | |
4 |
\right)\\ \end{align}
But comparing this to the FCC above, we see that it is simpler to describe the structure as FCC with a basis of two atoms at (0, 0, 0) and (1/4, 1/4, 1/4). For this basis, Equation becomes:
Fhk\ell(\rm{basis})=f
2 | |
\sum | |
j=1 |
[-2\pii(hxj+kyj+\ellzj)] | |
e |
=f\left[1+e[-i\right]=f\left[1+(-i)h\right]
And then the structure factor for the diamond cubic structure is the product of this and the structure factor for FCC above, (only including the atomic form factor once)
Fhk\ell=f\left[1+(-1)h+(-1)k+(-1)h\right] x \left[1+(-i)h\right]
with the result
|Fhk\ell|2=0
Fhk\ell=4f(1\pmi),|Fhk\ell|2=32f2
h+k+\ell=4n
Fhk\ell=4f x 2,|Fhk\ell|2=64f2
h+k+\ell ≠ 4n
|Fhk\ell|2=0
These points are encapsulated by the following equations:
Fhk\ell= \begin{cases} 8f,&h+k+\ell=4N\\ 4(1\pmi)f,&h+k+\ell=2N+1\\ 0,&h+k+\ell=4N+2\\ \end{cases}
⇒ |Fhk\ell|2= \begin{cases} 64f2,&h+k+\ell=4N\\ 32f2,&h+k+\ell=2N+1\\ 0,&h+k+\ell=4N+2\\ \end{cases}
N
The zincblende structure is similar to the diamond structure except that it is a compound of two distinct interpenetrating fcc lattices, rather than all the same element. Denoting the two elements in the compound by
A
B
Fhk\ell= \begin{cases} 4(fA+fB),&h+k+\ell=4N\\ 4(fA\pmifB),&h+k+\ell=2N+1\\ 4(fA-fB),&h+k+\ell=4N+2\\ \end{cases}
Cesium chloride is a simple cubic crystal lattice with a basis of Cs at (0,0,0) and Cl at (1/2, 1/2, 1/2) (or the other way around, it makes no difference). Equation becomes
Fhk\ell=
2 | |
\sum | |
j=1 |
fj
[-2\pii(hxj+kyj+\ellzj)] | |
e |
=\left[fCs+fCle[-i\right]=\left[fCs+fCl(-1)h\right]
We then arrive at the following result for the structure factor for scattering from a plane
(hk\ell)
Fhk\ell=\begin{cases}(fCs+fCl),&h+k+\ell&even\\ (fCs-fCl),&h+k+\ell&odd\end{cases}
and for scattered intensity,
|Fhk\ell|2=\begin{cases}(fCs+fCl)2,&h+k+\ell&even\\ (fCs-fCl)2,&h+k+\ell&odd\end{cases}
In an HCP crystal such as graphite, the two coordinates include the origin
\left(0,0,0\right)
\left(1/3,2/3,1/2\right)
Fhk\ell=f\left[1+e2\pi{3}+\tfrac{2k}{3}+\tfrac{\ell}{2}\right)}\right]
X\equivh/3+2k/3+\ell/2
|F|2=f2\left(1+e2\pi\right)\left(1+e-2\pi\right) =f2\left(2+e2\pi+e-2\pi\right) =f2\left(2+2\cos[2\piX]\right)=f2\left(4\cos2\left[\piX\right]\right)
|Fhk\ell|2=\begin{cases} 0,&h+2k=3Nand\ellisodd,\\ 4f2,&h+2k=3Nand\elliseven,\\ 3f2,&h+2k=3N\pm1and\ellisodd,\\ f2,&h+2k=3N\pm1and\elliseven\\ \end{cases}
The reciprocal lattice is easily constructed in one dimension: for particles on a line with a period
a
2\pi/a
S(q)
Fhk=
N | |
\sum | |
j=1 |
fj
[-2\pii(hxj+kyj)] | |
e |
However, recall that real 2-D crystals such as graphene exist in 3-D. The reciprocal lattice of a 2-D hexagonal sheet that exists in 3-D space in the
xy
z
z*
\pminfty
z
The Figure shows the construction of one vector of a 2-D reciprocal lattice and its relation to a scattering experiment.
A parallel beam, with wave vector
ki
a
ko
|ko|=|ki|
q=ko-ki
\exp(iqr)
q=2\pi/a
ko
q
Technically a perfect crystal must be infinite, so a finite size is an imperfection. Real crystals always exhibit imperfections of their order besides their finite size, and these imperfections can have profound effects on the properties of the material. André Guinier[5] proposed a widely employed distinction between imperfections that preserve the long-range order of the crystal that he called disorder of the first kind and those that destroy it called disorder of the second kind. An example of the first is thermal vibration; an example of the second is some density of dislocations.
The generally applicable structure factor
S(q)
Fhkl
For
S(q)
N
S(q)=
1 | |
N |
\left|
1-e-i | |
1-e-i |
\right|2=
1 | |
N |
\left[
\sin(Nqa/2) | |
\sin(qa/2) |
\right]2.
This function is shown in the Figure for different values of
N
q=2k\pi/a
\proptoN
\proptoN2
S(q)
S(q\to0)
S(q=2k\pi/a)=N
S(q=(2k+1)\pi/a)=1/N
1/N
N
In crystallography when
Fhkl
N
\left[
\sin(Nqa/2) | |
(qa/2) |
\right]2
S(q)
q ≈ 2k\pi/a
\ast
x
l{F}
l{F}
x l{F}
\ast{F}
\propto
x
\ast
\proptoN2
\propto1/N
\proptoN
This model for disorder in a crystal starts with the structure factor of a perfect crystal. In one-dimension for simplicity and with N planes, we then start with the expression above for a perfect finite lattice, and then this disorder only changes
S(q)
S(q)=
1 | |
N |
\left[
\sin(Nqa/2) | |
\sin(qa/2) |
\right]2\exp\left(-q2\langle\deltax2\rangle\right)
where the disorder is measured by the mean-square displacement of the positions
xj
a(j-(N-1)/2)
xj=a(j-(N-1)/2)+\deltax
\deltax
a
\deltax
\deltax
N\toinfty
N\toinfty
q2
q
q
The structure is simply reduced by a
q
In three dimensions the effect is the same, the structure is again reduced by a multiplicative factor, and this factor is often called the Debye–Waller factor. Note that the Debye–Waller factor is often ascribed to thermal motion, i.e., the
\deltax
However, fluctuations that cause the correlations between pairs of atoms to decrease as their separation increases, causes the Bragg peaks in the structure factor of a crystal to broaden. To see how this works, we consider a one-dimensional toy model: a stack of plates with mean spacing
a
To derive the model we start with the definition (in one dimension) of the
S(q)=
1 | |
N |
N | |
\sum | |
j,k=1 |
-iq(xj-xk) | |
e |
To start with we will consider, for simplicity an infinite crystal, i.e.,
N\toinfty
m
-m
m
N
S(q)=1+2
infty | |
\sum | |
m=1 |
infty | |
\int | |
-infty |
{\rmd}(\Deltax)pm(\Deltax)\cos\left(q\Deltax\right)
where
pm(\Deltax)
\Deltax
m
p1(\Deltax)=
1 | |||||||||
|
\exp\left[-\left(\Delta
2)\right] | |
x-a\right) | |
2 |
and we also assume that the fluctuations between a plane and its neighbour, and between this neighbour and the next plane, are independent. Then
p2(\Deltax)
p1(\Deltax)
pm(\Deltax)=
1 | ||||||||
|
\exp\left[-\left(\Delta
2)\right] | |
x-ma\right) | |
2 |
The sum in
S(q)
infty | |
S(q)=1+2\sum | |
m=1 |
rm \cos\left(mqa\right)
for
2/2] | |
r=\exp[-q | |
2 |
infty | |
\sum | |
m=1 |
[r\exp(iqa)]m
S(q)= | 1-r2 |
1+r2-2r\cos(qa) |
This has peaks at maxima
qp=2n\pi/a
\cos(qPa)=1
S(q | ≈ | ||||
|
4 | = | |||||
|
a2 | ||||||||||||
|
i.e., the height of successive peaks drop off as the order of the peak (and so
q
q\sigma2\ll1
r\simeq
2/2 | |
1-q | |
2 |
\cos(qa)\simeq1-(\Deltaq)2a2/2
\Deltaq=q-qP
S(q) ≈ | S(qP) | ||||||
|
≈
S(qP) | ||||||||||||
|
which is a Lorentzian or Cauchy function, of FWHM
2/a=4\pi | |
q | |
2 |
2n
2/a | |
2/a) |
q
Finally, the product of the peak height and the FWHM is constant and equals
4/a
q\sigma2\ll1
n
\sigma2/a\ll1
For a one-dimensional crystal of size
N
| ||||
S(q)=1+2\sum | ||||
m=1 |
\right)rm\cos\left(mqa\right)
where the factor in parentheses comes from the fact the sum is over nearest-neighbour pairs (
m=1
m=2
N
N-1
N-2
In contrast with crystals, liquids have no long-range order (in particular, there is no regular lattice), so the structure factor does not exhibit sharp peaks. They do however show a certain degree of short-range order, depending on their density and on the strength of the interaction between particles. Liquids are isotropic, so that, after the averaging operation in Equation, the structure factor only depends on the absolute magnitude of the scattering vector
q=\left|q\right|
j=k
One can obtain an alternative expression for
S(q)
g(r)
In the limiting case of no interaction, the system is an ideal gas and the structure factor is completely featureless:
S(q)=1
Rj
Rk
\langle\exp[-iq(Rj-Rk)]\rangle=\langle\exp(-iqRj)\rangle\langle\exp(iqRk)\rangle=0
Even for interacting particles, at high scattering vector the structure factor goes to 1. This result follows from Equation, since
S(q)-1
g(r)
q
In the low-
q
\chiT
\limqS(q)=\rhokBT\chiT=kBT\left(
\partial\rho | |
\partialp |
\right)
In the hard sphere model, the particles are described as impenetrable spheres with radius
R
r\geq2R
V(r)=\begin{cases} infty&forr<2R,\\ 0&forr\geq2R. \end{cases}
This model has an analytical solution[9] in the Percus–Yevick approximation. Although highly simplified, it provides a good description for systems ranging from liquid metals[10] to colloidal suspensions.[11] In an illustration, the structure factor for a hard-sphere fluid is shown in the Figure, for volume fractions
\Phi
In polymer systems, the general definition holds; the elementary constituents are now the monomers making up the chains. However, the structure factor being a measure of the correlation between particle positions, one can reasonably expect that this correlation will be different for monomers belonging to the same chain or to different chains.
Let us assume that the volume
V
Nc
Np
NcNp=N
Np
\alpha,\beta
j,k
\alpha=\beta
\alpha ≠ \beta
S1(q)
Fhkl