You are on page 1of 21

Unifying the Concepts of Scattering and Structure Factor in Ordered and Disordered

Samples
Dingning Li1 and Kai Zhang1, ∗
1
Division of Natural and Applied Sciences, Duke Kunshan University, Kunshan, Jiangsu, 215300, China
(Dated: April 2, 2021)
Scattering methods are widely used in many research areas to analyze and resolve material struc-
tures. Given the importance, a large number of full textbooks are devoted to this topic. However,
technical details in experiments and disconnection between explanations from different perspectives
often confuse and frustrate beginner students and researchers. To create an effective learning path,
we review the core concepts about scattering and structure factor in this article in a self-contained
way. Classical examples of scattering photography and spectroscopy are calculated. Sample CPU
arXiv:2010.06126v2 [cond-mat.soft] 1 Apr 2021

and GPU codes are provided to facilitate the understanding and application of these methods.

1. INTRODUCTION structure factor S(q), which is often expressed as a func-


tion of scattering vector q. The central tasks of structural
Scattering methods, using a source of photons, elec- analysis with scattering methods are then
trons, X-rays, neutrons, etc., are powerful tools to ex-
amine microscopic structural [1] and dynamical [2] prop- • the forward problem ρ(r) → I(q): given the elec-
erties of matter, which have been successfully applied tron density distribution ρ(r) or particle positions
to study subatomic particles [3], crystals [4], liquids [5], (r1 , r2 , · · · , rN ), to predict the scattering pattern
glasses [6], surfactants [7], biomolecules [8, 9] and poly- I(q); and
mers [10]. The rule of thumb here is that the wavelength
• the inverse problem I(q) → ρ(r): given the scatter-
λ of the radiation should be comparable to the length
ing pattern I(q), to resolve the electron density dis-
scale of the structure to be observed. To detect ordering
tribution ρ(r) or particle positions (r1 , r2 , · · · , rN ).
over a range much longer than λ, methods like small-
angle scattering (SAS) are needed [11]. Another impor- In this article, we only focus on the forward problem,
tant consideration is about the contrast between scat- which could still shed light on some basic structural in-
tering signals from different elements due to underlying formation. Sometimes, the forward method may also be
physical mechanisms. Therefore, neutron scattering is used to solve ρ(r) iteratively, through a trial-and-error
often preferred for soft-matter systems, despite its lower process. That is, one keeps modifying a proposed struc-
accessibility than for X-rays. In addition, techniques like ture ρ(r) until the theoretically computed I(q) matches
resonant soft-X-ray scattering can be used to provide en- the experimentally observed one. The full solution to the
hanced resolution [12, 13]. Compared with real-space inverse problem is, however, challenged by the notorious
microscopy techniques, reciprocal-space probes like scat- “phase problem” [17].
tering methods are good at picking up periodic patterns Concepts about scattering and structure factor are
and revealing three-dimensional structures as a whole by often discussed across different disciplines including
penetrating deeply into the sample [14]. condensed-matter physics, materials science, polymer
Given the richness of material structures, a variety of physics, structure biology, etc. The same idea can take
experimental methods have been developed during the different forms in different areas, causing confusions and
last century, with the scattering being hard (high-energy) misconceptions. Graduate or advanced undergraduate
or soft (low-energy), monochromatic or polychromatic, students in need of applying these concepts to their re-
elastic or inelastic. Despite the diversity of experimen- search problems can be frustrated by the convoluted ex-
tal setups, they can largely be grouped into two cate- perimental details covered in traditional textbooks. It
gories based on how signals are collected and interpreted. is thus the purpose of this article to unify the concepts
The first category is photography of ordered samples, about scattering and structure factor, giving junior re-
which are recorded as spotted scattering signals on a searchers an effective pathway to quickly grasp the key
two-dimensional (2D) film [15]. The second category is ideas in this field without taking a whole course or read-
intensity scanning of scattering signals from disordered ing an entire textbook.
or partially ordered samples, whose one-dimensional (1D) To fulfill this task, we first elaborate the fundamen-
profile is plotted against one variable (a scalar) that char- tals about scattering (Section 2), crystallography (Sec-
acterizes the existence of periodicities in the system [16]. tion 3) and liquid-state theory (Section 4) based on the
In both types, the quantitative measurement of the sig- Fourier transform and reciprocal lattice. Using concrete
nal is scattering intensity I(q), or its normalized version, examples, we then discuss the photography of ordered
samples in Section 5-6 as well as intensity scanning of
isotropic samples in Section 7-8. Relevant CPU and GPU
∗ kai.zhang@dukekunshan.edu.cn source codes are provided online at https://github.
2

structural features of the sample, for instance, the elec-


tron density distribution ρ(r) in the case of X-ray scat-
tering by atoms.
Both the incident and the diffracted rays can be viewed
q k1 q as plane waves of the form ψk (r) = hr|ki ∝ eik·r . Accord-
ing to Fermi’s golden rule, the scattering intensity I(q)
is proportional to the square of the transition probability
2θ amplitude from state ψk0 (r) to state ψk1 (r), after inter-
k0 2π
acting with the overall scattering potential ρ(r). That
is,
λ
2 R 2
I(q) ∝ |hk0 |ρ(r)|k1 i| = drψk∗ 0 (r)ρ(r)ψk1 (r)
R 2
FIG. 1. Scattering vector q defined as the difference between ∝ dre−ik0 ·r ρ(r)eik1 ·r
the diffracted wavevector k1 and the incident wavevector k0 , R 2
both with magnitude 2π/λ during elastic scattering.
= drρ(r)eiq·r
Neglecting the coefficient of proportionality, one can
write
com/statisticalmechanics/scatter. Finally, a brief
introduction to the 2D structure factor is given in Sec- I(q) = ρ̂q ρ̂−q , (4)
tion 9, before the conclusion in Section 10. where
Z
ρ̂q = drρ(r)eiq·r (5)
2. SCATTERING
is the Fourier transform of the density distribution and
2.1. Scattering Vector ρ̂−q is its complex conjugate (Appendix A).
Unless ρ(r) has a symmetry center, ρ̂q is generally a
In a scattering experiment, the incident beam of complex number, i.e. ρ̂q = |ρ̂q |eiφq . If ρ̂q is known
wavevector k0 , after hitting the sample, is deflected from exactly, ρ(r) can in principle be reconstructed through
its straight path by a scattering angle 2θ and becomes the inverse Fourier transform Eq. (A2) [19]. However,
the diffracted beam of wavevector k1 (Fig. 1). In case in an experiment, only the scattering intensity I(q) =
of elastic [18] and monochromatic scattering (of a fixed |ρ̂q |2 eiφq e−iφq = |ρ̂q |2 is directly measurable.p
This allows
wavelength λ), |k0 | = |k1 | = 2π
λ . The change of wavevec-
us to compute the magnitude of ρ̂q by |ρ̂q | = I(q). Un-
tor, called the scattering vector, is fortunately, information about the phase angle φq is lost
during this process, which gives rise to the “phase prob-
q = k1 − k0 (1) lem” in crystallography. Special techniques [17, 20, 21]
have been developed to determine φq , which are beyond
with a magnitude the scope of this article.
In a system of N atoms or particles at positions
4π (r1 , r2 , · · · , rN ) inside a region of volume V , the den-
q = 2|k0 | sin θ = sin θ. (2)
λ sity distribution consists of the contributions from each
Let s0 = k0 /|k0 | = k0 λ/(2π) and s1 = k1 /|k1 | = particle i with a scattering potential fi (r − ri ) (i =
k1 λ/(2π) be the unit vectors of the incident and 1, 2, · · · , N ), i.e.
diffracted beam, then the scattering vector can also be N
X N
X
written as ρ(r) = fi (r − ri ) = fi (Ri ), (Ri ≡ r − ri ). (6)
i=1 i=1

q= (s1 − s0 ). (3) In this case
λ
N
fi (r − ri )eiq·r
R P
ρ̂q = dr
V i=1
N
dRi fi (Ri )eiq·Ri eiq·ri
P R
2.2. Scattering Intensity = (7)
i=1 V
N
fˆi (q)eiq·ri
P
When a detection screen is placed behind the sample =
i=1
in the path of k1 , the diffracted beam may be detected.
The strength of such signals is quantified by the scat- where
tering intensity I(q) of the ray, which changes with k1 ,
Z
or equivalently, with q. The scattering pattern, or the fˆi (q) = drfi (r)eiq·r (8)
distribution of I(q) on the screen, is determined by the V
3

is the atomic form factor, or scattering factor, of particle


i.
If the scattering potential of each particle fi (r − ri ) is
symmetric about ri , which should be true for atoms and
most particles, fˆi (q) is real and even, i.e. its complex
conjugate, fˆi∗ (q) = fˆi (−q) = fˆi (q) (Appendix A) Under
this circumstance, the scattering intensity
N N
fˆi (q)eiq·ri fˆj (−q)e−iq·rj
P P
I(q) =
i=1 j=1
N N
fˆi (q)eiq·ri fˆj (q)e−iq·rj
P P
=
i=1 j=1
N 2 N 2
= fˆi (q) cos(q · ri ) + fˆi (q) sin(q · ri )
P P
i=1 i=1
(9) FIG. 2. Atomic form factor fˆi (q) of a sizeless point (green
dotted, Eq. 12), a uniform spherical (red solid, Eq. 14) and a
or equivalently,
Gaussian scattering center (blue dashed, Eq. 16) as a function
N P
N of q.
fˆi (q)fˆj (q)eiq·(ri −rj )
P
I(q) =
i=1 j=1
N P N
• fi (r − ri ) = ai δ(r − ri ), the scattering by each atom
fˆi (q)fˆj (q)eiq·rij
P
=
i=1 j=1 is idealized as from a sizeless point at the atomic
N P N (10) center. This model can be mapped onto the phys-
fˆi (q)fˆj (q) cos(q · rij )
P
= ical scenario of nuclear scattering or the abstract
i=1 j=1
N N P
N scenario of point-mass scattering. The scattering
fˆi2 (q) fˆi (q)fˆj (q) cos(q · rij ). strength ai of atom i generally has different values
P P
= +
i=1 i=1 j6=i for different elements, which has also been called
the atomic scattering factor, because here
Eq. (9) and Eq. (10) are mathematically equivalent be-
cause cos(q · rij ) = cos(q · ri − q · rj ) = cos(q · ri ) cos(q · fˆi (q) = ai . (12)
rj ) + sin(q · ri ) sin(q · rj ). However, in numerical compu-
tation of I(q) at a given q, Eq. (9) has a lower cost with The electron density distribution is then ρ(r) =
N
a computational complexity O(N ), while Eq. (10) is of P
ai δ(r−ri ), which, in the case of ai = 1, becomes
complexity O(N 2 ). Nevertheless, when there is an ap- i=1
propriate symmetry in the system, the expression rij in N
P
Eq. (10) allows it to be further simplified and thus to be- the particle density distribution ρ(r) = δ(r−ri ).
i=1
come more computationally efficient, as will be discussed
in later sections. • fi (r − ri ) is homogeneous and bounded within a
sphere of radius σ/2 [22],

2.3. Atomic Form Factor ai /(πσ 3 /6), |r − ri | ≤ σ/2
fi (r − ri ) = (13)
0, otherwise
For realistic scattering potentials, the atomic form fac- and
tor fˆi (q) changes with the direction and magnitude of the 4πai /(πσ 3 /6)
scattering vector q, and thus often drops as the scatter- fˆi (q) = [sin(qσ/2) − qσ/2 cos(qσ/2)]
ing angle θ increases (Fig. 2). If, however, the scattering q3
potential is spherically symmetric, i.e. fi (r) = fi (r), we 3ai
= [sin(qσ/2) − qσ/2 cos(qσ/2)].
can write (qσ/2)3
Z Z π (14)
fˆi (q) = fˆi (q) = 2π drr2 fi (r) dθ sin θeiqr cos θ
0 • fi (r − ri ) is Gaussian-like with standard deviation
2 sin(qr) σ/2,
Z
= 2π drr2 fi (r) (11)
qr !3
|r−ri |2
Z
sin(qr) 1 −
= 4π drr2 fi (r) (q 6= 0). fi (r − ri ) = ai p e 2(σ/2)2
qr 2π(σ/2)2 (15)
ai 2R2
It is useful to consider the three simple spherically sym- − σ2i
= 3 e (Ri = |r − ri |)
metric scattering potentials listed below (Fig. 2). σ (π/2)3/2
4

d∗hkl rays “reflected” by two neighboring lattice planes is

l(cos α + cos β) = l[cos α + cos(π − 2θ − α)]


s 1 − s0 = l[cos α − cos(2θ + α)]
= l sin(θ + α) sin θ
s1 = dhkl sin θ.

s0 θ λ
The rescaled scattering vector s1 − s0 = 2π q (of length
θ 2 sin θ) is parallel to the normal vector, or reciprocal vec-
l α tor d∗hkl (of length 1/dhkl ), of the lattice planes (hkl).
β θ Thus, it is sometimes convenient to express Bragg’s law
in a vector form, for instance, for the primary n = 1
θ scattering, as
dhkl s1 − s0
= d∗hkl . (18)
λ

FIG. 3. The scattering paths of two rays diffracted by two Using Eq. (3), the necessary condition to receive a strong
layers of ordered particles (black dots) with interplanar dis- signal for scattering vector q in crystals is thus
tance dhkl and scattering angle 2θ. s0 and s1 are unit vectors
of the incident and diffracted rays. When Bragg’s law is sat- q = 2πd∗hkl . (19)
isfied, the scattering vector is parallel to the normal vector
d∗hkl of the lattice planes.

3.2. The Ewald Construction


and
σ q2 2 Bragg’s law needs to be satisfied to have a strong scat-
fˆi (q) = ai e− 8 . (16) tering signal in the direction of s1 . However, this does
not mean that, given an arbitrary experimental setup,
In all numerical results shown below, we will assume Bragg’s law is guaranteed to be satisfied somewhere. In
fˆi (q) = 1, i.e. point scattering, for all particles. particular, if a monochromatic incident beam (fixed λ)
is directed onto a single crystal at an arbitrarily fixed
position (fixed θ’s and dhkl ’s), it is possible that none of
the lattice planes will be able to produce a strong scat-
3. CRYSTALLOGRAPHY
tering signal. If this happens, either λ (polychromatic)
or θ (rotate the sample or use polycrystals) has to been
We now review concepts and theories about scatter- tuned to satisfy Eq. (17-19).
ing methods used for crystal samples. The earlier the- An alternative view to check the satisfaction of Bragg’s
ory of von Laue [23] that considers diffraction of parallel law is to use Ewald’s sphere in the reciprocal space [25,
beams by arrays of atoms is skipped here. Instead, we 26]. Here, each point at vector d∗hkl represents a family
apply the more intuitive Bragg’s law that envisages crys- of parallel planes (hkl) in the direct space. When the
tallographic planes as reflective mirrors to understand orientation of the crystal sample is fixed, the relative po-
the principle, although there is no such reflection in the sition of the incident beam and reciprocal lattice points
physical sense. are also fixed. One can align the end point of the incident
wavevector k0 (in practice k0 /2π) with the origin O of
the reciprocal lattice, then draw a sphere of radius 1/λ.
3.1. Bragg’s Law The center of the sphere is found by moving from point
O by a vector displacement −k0 /2π (Fig. 4). It can be
For an incident ray of wavelength λ to generate a seen that the end point of the scattering vector q, nor-
strong constructive scattering signal by a family of crys- malized by 2π, falls on the surface of this Ewald’s sphere.
tallographic planes (hkl) of interplanar spacing dhkl (Ap- According to the vector form of Bragg’s law Eq. (19),
pendix B), the scattering angle 2θ needs to obey Bragg’s a scattering from certain lattice planes (hkl) is possible
law [24] (Fig. 3) only when the corresponding reciprocal vector point d∗hkl
falls on the surface of Ewald’s sphere. If wavelength and
nλ = 2dhkl sin θ, n = 1, 2, 3, · · · . (17) crystal orientation are not appropriately chosen, this con-
dition may not be met at all and no scattering signal is
This is because the path difference of the two scattering generated by the sample.
5

to consider q’s of the form


q = 2πd∗hkl = 2π(ha∗ + kb∗ + lc∗ ),
(1̄20)
where d∗hkl represents a family of lattice planes (hkl) of
q spacing dhkl = 1/|d∗hkl | (Appendix B). The associated
k1
2π 2π b∗ structure factor can thus be denoted as Fhkl
Z
Fhkl = drρ(x, y, z)e2πi(hx+ky+lz) . (22)
k0 O a∗
2π 1 Vcell
λ
Inversely, the density distribution within each unit cell is
1 X
ρ(r) = Fhkl e−2πi(hx+ky+lz) . (23)
Vcell
hkl
m
P
For point-like scattering centers, ρ(r) = ai δ(r − ri )
i=1
and Eq. (22) reduces to
m
FIG. 4. The Ewald construction: Ewald’s sphere of radius X
1/λ (solid circle) depicts all possible scattering wavevectors Fhkl = ai e2πi(hxi +kyi +lzi ) , (24)
q under the current setup. Lattice planes with Miller in- i=1
dices (hkl) are represented by points on the reciprocal lattice
after substituting Eq. (12) and following steps in Eq. (7),
(black dots). For wavelength λ, no reciprocal lattice point is
on Ewald’s sphere implying that no scattering signal will be
where (xi , yi , zi ) are coordinates of the m atoms inside
generated at any scattering angle. If the wavelength is appro- one unit cell and are expressed as fractions of lattice vec-
priately tuned, some reciprocal lattice points can fall on the tors. The strength of Fhkl by planes (hkl) is the vector
new Ewald’s sphere (dashed circle) to satisfy Bragg’s law, for sum of each term ai e2πi(hxi +kyi +lzi ) in Eq. (24), where
instance, (1̄20). the phase angle hxi +kyi +lzi defines the direction of each
vector. For typical crystal lattices of point-like atoms of
the same type (ai = a), Fhkl can be easily computed.
3.3. Crystal Structure Factor Fhkl • Simple Cubic (SC)
m = 1 and (x1 , y1 , z1 ) = (0, 0, 0)
Bragg’s law is actually the necessary (but not suffi-
SC
cient) condition to have a strong scattering signal. Even Fhkl = ae2πi(h0+k0+l0) = a (25)
if Bragg’s law is obeyed by lattice planes (hkl), it is still
for any h, k, l.
possible that the scattering signal cancels due to special
lattice symmetries. In fact, when Bragg’s law is presented • Body-Centered Cubic (BCC)
as in Fig. 3, a simple square or oblique lattice structure m = 2, (x1 , y1 , z1 ) = (0, 0, 0) and (x2 , y2 , z2 ) =
is often used, which misses the complexity in other three- (1/2, 1/2, 1/2)
dimensional lattices. Generally, not every family of lat-
1 1 1
BCC
tice planes (hkl) can produce a constructive scattering. Fhkl = ae2πi(h0+k0+l0) + ae2πi(h 2 +k 2 +l 2 )
Because the density distribution ρ(r) is periodic in (26)
= a + aeπi(h+k+l) .
crystals, one only needs to consider the particle distribu-
tion within one unit cell. If each unit cell has a volume • Face-Centered Cubic (FCC)
Vcell and m atoms, then
m = 4, (x1 , y1 , z1 ) = (0, 0, 0), (x2 , y2 , z2 ) =
N
Z (1/2, 1/2, 0), (x3 , y3 , z3 ) = (0, 1/2, 1/2) and
ρ̂q = drρ(r)eiq·r , (20) (x4 , y4 , z4 ) = (1/2, 0, 1/2)
m
Vcell
1 1
FCC
where N/m is the number of unit cells in the N -particle Fhkl = ae2πi(h0+k0+l0) + ae2πi(h 2 +k 2 +l0)
system. In crystallography, it is customary to define ρ̂q 1 1 1
+ ae2πi(h0+k 2 +l 2 ) + ae2πi(h 2 +k0+l 2 )
1
(27)
per unit cell as the structure factor, πi(h+k) πi(k+l) πi(h+l)
= a + ae + ae + ae .
Z
Fq = drρ(r)eiq·r . (21) The Fhkl of BCC and FCC lattices completely vanishes
Vcell for certain h, k, l. The resulting reflection Miller indices
should successively be (110), (200), (211), (220), (310),
For crystals, only q’s satisfying Bragg’s law (19) can pos- (222) · · · for BCC, and (111), (200), (220), (311), (222),
sibly generate a large ρ̂q or Fq . Therefore, we only need (400) · · · for FCC crystals.
6

δθ Thus diffraction signals tend to be larger in smaller sys-


θ tems.
dhkl

.. .. 4. LIQUID-STATE THEORY

L/2
According to liquid-state theory, a static structure fac-
.. L
sin(θ + δθ)
.. tor S(q) can be used to address short-range order [27, 28]
2
and the glass transition [29] in amorphous/liquid sam-
ples [30] and more generally in nano-structured or other
L
2
sin θ structurally disordered systems [31]. In an N -particle
.. .. system, it is defined as
1 1
.. .. S(q) = N
hρ̂q ρ̂−q i = N
I(q), (29)
fˆi2 (q) fˆi2 (q)
P P
i=1 i=1

where the ensemble average h· · · i is usually taken over


configurations at thermal equilibrium [32]. Practically,
FIG. 5. Illustration of the Scherrer equation. In a finite-size
this ensemble average results from a sum over all the
crystal of thickness L, the path difference between two rays different coherence volumes in the sample, after being
reflected by a pair of planes that are L/2 apart sets the limit Fourier transformed, giving a real-space representation
of the signal broadening 2δθ. of the sample’s ensemble-averaged instantaneous local
structure.
If scattering centers are point-like, i.e. ρ(r) =
N N
3.4. Finite-Size Crystals and Bragg Peak
ai eiq·ri and
P P
ai δ(r − ri ), then ρ̂q =
Broadening i=1 i=1
* N 2 N 2 +
When Bragg’s law is satisfied by wavelength λ at an 1 X X
S(q) = N ai cos(q · ri ) + ai sin(q · ri ) .

incident angle θ, a small deviation δθ from θ only slightly P 2


changes the path difference between two rays reflected by ai i=1 i=1

a pair of neighboring planes (of spacing dhkl ), which still i=1

add constructively. If we consider two reflection planes (30)


that are 2dhkl , 3dhkl , · · · , apart, the change in path differ- For monodisperse systems (ai is the same for all parti-
ence due to δθ accumulates, and at large enough spacing, cles), S(0) = N .
becomes λ/2 such that the two rays completely cancel. In the case of ai = 1, S(q) is related to the radial
For a beam reflected by a crystallographic plane in large distribution function g(r) or the pair correlation function
crystal samples, it is always possible to find another re- h(r) = g(r) − 1 by
mote plane whose reflected beam interferes destructively,
even for very small δθ. Therefore, when other broadening
Z Z
effects are excluded, diffraction signals in large samples S(q) = 1 + ρ0 dr(g(r) − 1)eiq·ri + ρ0 dreiq·ri
at fixed λ, if there are any, should in principle be of in- Z
finitely small size (in terms of the range of θ). = 1 + ρ0 dr(g(r) − 1)eiq·ri + ρ0 (2π)3 δD (q)
For small crystal samples, it is possible that the change Z
in path difference due to δθ is much less than λ/2 such = 1 + ρ0 dr(g(r) − 1)eiq·ri + ρ0 V δq,0 (finite V )
that Bragg’s law is still approximately satisfied at θ + δθ V
and the diffraction signal is broadened by an amount Z
∼ δθ. The quantitative relationship between the broad- = 1 + ρ0 drh(r)eiq·ri + N δq,0
ening 2δθ of the signal and the linear dimension L of a V
finite-size crystal can be found by considering all pairs of
planes that are L/2 apart. When θ changes to θ + δθ, = 1 + ρ0 ĥq + N δq,0
the path difference for such a pair of planes increases by (31)
2 L2 [sin(θ + δθ) − sin θ] = L cos θδθ (Fig. 5). The diffrac-
where the global number density ρ0 = N/V and the
tion signal broadens until destructive interference occurs
Fourier transform ĥq = drh(r)eiq·ri . Note that S(q)
R
at λ2 = L cos θδθ, which gives the Scherrer equation
V
is singular or discontinuous at q = 0, i.e. lim S(q) 6=
λ 2 tan θ q→0
2δθ = = . (28) S(0) = N . Correspondingly, lim ĥq 6= ĥ0 = −1/ρ0 .
L cos θ L/dhkl q→0
7

The radial distribution function can be obtained from Therefore,


the structure factor by the inverse Fourier transform  
2π X Y D+L
1 S(q) − 1 −iq·ri (qx , qy , qz ) = , ,− (back-reflection)
Z
g(r) = 1 + dq e , (32) λ L L L
(2π)3 ρ0 (37)
q→0

where the value lim S(q) should be used at q = 0 in In contrast, one can show that, in the transmission
q→0
method,
the integration. When the system’s structure is isotropic
over the sample volume, i.e. g(r) = g(r), more convenient 2π

X Y L−D

relationships can be derived [33] (qx , qy , qz ) = , ,− (transmission).
λ L L L
Z ∞ (38)
sin(qr)
S(q) = 1 + 4πρ0 dr(g(r) − 1)r2 , (33)
+ qr
Z0 ∞
1 S(q) − 1 2 sin(qr)
g(r) = 1 + 2 dq q , (34) 5.2. Cylindrical Method
2π 0+ ρ0 qr
sin x
where lim x = 1 should be used in the integration. Compared with above two setups, the cylindrical
x→0
The limit value of S(q) as q approaches zero is related to method collects signals from all azimuthal angles φ and
the isothermal compressibility κ by [34] is thus more informative (Fig. 7). In fact, a cylindrical
film, which better preserves the natural shape of scatter-
lim S(q) = ρ0 kB T κ. (35) ing spots, can be considered as the sum of an infinitely
q→0
wide back-reflection film and an infinitely wide transmis-
sion film, on which scattering patterns farther away from
5. EXPERIMENTAL SETUPS IN the film center are more distorted.
PHOTOGRAPHY To map q onto the film, one can unfold the cylinder
into a plane with coordinates (X, Y ) = (D sin φ, Y ) with
the azimuthal angle φ ∈ (−π, π). The relationship is
In this section, we discuss some technical details
about photography methods, which collects signals of
D sin φ Y D cos φ − D
 
I(q) = I(X, Y ) on a two-dimensional film with coor- 2π
(qx , qy , qz ) = , , . (39)
dinates (X, Y ). Three popular experimental setups are λ L L L
often used as described below, which map q onto (X, Y )
differently.
6. PHOTOGRAPHY OF SINGLE
CRYSTALLINE SAMPLES
5.1. Back-reflection and Transmission Methods
The illustration of Bragg’s law using Ewald’s sphere
In back-reflection and transmission methods, the suggests two ways to make reciprocal lattice points fall
recording film is a rectangular plane, which is placed ei- on the sphere to generate constructive scattering signals
ther before (back-reflection) or after (transmission) the from specific crystallographic planes. One is to tune the
sample as shown in Fig. 6. In both methods, it can be wavelength and the other is to change the orientation
seen that the ratio qx /qy equals X/Y . If the incident of the sample. These correspond to two experimental
wave number is |k0 | = 2π/λ, then strategies in designing photography methods for single
  crystals – the Laue method and the (monochromatic)
2π X Y rotation method.
(qx , qy ) = , , (36)
λ L L

where L2 = R2 + D2 and R2 = X 2 + Y 2 . The difference


remains in the z component qz . 6.1. Varying Wavelength at Fixed Angle – Laue
In the back-reflection method, because α = π2 − θ sat- Method
isfies cos(2α) = D/L, it follows that
In the Laue method, one fixes the orientation of the
2π sample (thus the angle θ in Bragg’s law) and changes the
qz = −2 cos α cos α
λ wavelength of incident beam over a certain range λ ∈
2π [λmin , λmax ], which is thus called “white color”.
= − [1 + cos(2α)]
λ For each pixel (X, Y ) on the film, the scattering in-
2π tensity is thus the sum of contributions from all wave-
= − (1 + D/L).
λ lengths, or equivalently, all parallel scattering vectors q,
8

(b)
(a)
(X, Y )
(X, Y )
L
L
R R
k1 q k1
q k0 2θ
α 2θ
α
(hkl) planes
k0 D
D
(hkl) planes

y
y
x x
z z

FIG. 6. Illustration of (a) back-reflection and (b) transmission methods. The scattering vector q resulted from crystallographic
planes (hkl) maps onto 2D coordinates (X, Y ) on the film, which is placed at a distance D from the sample. The incident beam
is along the z axis.

(a) D (b)

qx (X, Y )
L
q0
q Y qz
k1 φ D
0
k0 q k0
φ 2π
λ

y
x x

z −π π

FIG. 7. Illustration of cylindrical method from (a) the side view and (b) the top view. A cylindrical film is placed at a radius
D around the sample.

which can be formally written as beam is along the [001] direction and the nearest neigh-
¯ P bor distance σ is set as the unit of length. The code to
I(X, Y ) = I(q) ¯
q compute I(X, Y ) numerically implementing Eq. (40) is
2 2 # provided online. The value of D can be chosen arbitrar-
"
N N
P P ˆ ˆ
P
= fi (q) cos(q · ri ) + fi (q) sin(q · ri ) . ily, with all other lengths calculated accordingly, because

q i=1 i=1
it only leads to a scaling of the photograph. Here, we
(40) set D = 100σ for numerical convenience. If the total
This type of general equation, which computes scattering number of pixels on the film is NXY and the number
intensity from all atoms in the sample, reduces to the of wavelengths scanned is Nλ , then the computational
simple summation over atoms in the unit cell for ideal complexity using Eq. (40) is O(NXY Nλ N ).
crystals, as explained in Section 3 3.3.
We demonstrate the photography results using per-
fect SC, BCC and FCC samples (Fig. 8). The incident
9

The scattering intensity at coordinates (X, Y ) is then

¯
I(X,
" Y ) = 2 2 #
N N
fˆi (q) cos(q · ri (Ω)) + fˆi (q) sin(q · ri (Ω))
P P P

Ω i=1 i=1
(41)

where Ω represents the orientation of the sample due to


rotation. For a given sample, we apply a rotation ma-
trix about its y-axis to transform particle coordinates
into new values. The accumulated signal I(X,¯ Y ) on the
cylinder is then unfolded onto a rectangle. If a total num-
ber NΩ of rotation angles within (0, 2π) are scanned, the
computational complexity to implement Eq. (41) to pro-
duce results on NXY pixels is then O(NXY NΩ N ).

6.3. Broadening due to Finite-size Effect

So far we assume that either varying wavelength or


varying sample orientation is needed to satisfy Bragg’s
law and produce nonvanishing scattering signals on the
photograph. However, this is only true for infinitely large
systems. In our small samples with N ∼ 103 particles,
signal broadening allows us to observe certain scattering
patterns, even when the wavelength λ is fixed at one
appropriate value.
For example, in the previously mentioned SC crystals,
we can see four scattering spots in the back-reflection
method at fixed wavelength λ = 0.55σ, which correspond
to the (113) planes and equivalents (Fig. 10). When the
system size is varied from N = 73 to 303 , the size of each
FIG. 8. Back-reflection (left column) and transmission (right spot decreases. It can be√confirmed that the relationship
column) photography of SC (a-b) (N = 3375), BCC (c-d) between box size L = 3 N and spot size 2δθ roughly
(N = 4394) and FCC (e-f) (N = 5324) crystals. The range of satisfies the Scherrer equation 2δθ ∝ 1/L. An empirically
wavelength λ is 0.35-1.0σ for SC back-reflection, 0.199-0.35σ √
scaling factor 2 is needed on L to estimate the actual
for SC transmission, 0.4-1.2σ for BCC back-reflection, 0.23-
dimension of the sample perpendicular to (113) planes
0.4σ for BCC transmission, 0.5-1.42σ for FCC back-reflection
and 0.23-0.49σ for FCC transmission. The (X, Y ) coordinate
and to agree with the theoretical slope λ/ cos θ.
range [−150, 150] is now set by the grid resolution of the com-
puter code, which can be mapped onto the real length unit
on a physical film. 6.4. DNA Double Helix

One of the most successful and famous applications


6.2. Varying Angle using Fixed Wavelength – of scattering methods is the determination of the DNA
Rotation Method structure, whose X-ray photography shows a character-
istic “X”-shape pattern with horizontal stripes [35]. The
We use the conventional setup – cylindrical film – to form of the pattern can be understood analytically by
explain the rotation method for the same SC, BCC and diffraction from the 2D projected sinusoidal waves of the
FCC crystalline samples as above (Fig. 9). The wave- single or double helix [36, 37]. Here we produce the trans-
length λ of the incident beam is fixed in this method and mission photography of a single model DNA fiber with
the sample placed at the central axis of cylinder is ro- only backbone particles. Each helix has N = 70 particles
tated by a certain angle to probe possible orientations with 10 particles per turn (pitch). The parameters of the
and scattering angle 2θ for given crystallographic planes. right-handed B-DNA, 34σ for pitch and 20σ for helix di-
A full circle of 2π rotation is only necessary for noncen- ameter, are used (Fig. 11). The unit of length σ can be
trosymmetric crystals containing elements that exhibit mapped onto the real length unit Å.
anomalous dispersion; a rotation of 2π is sufficient for All the four photographs, with the fiber being single
centrosymmetric crystals. or double strand, 2D projected or 3D stereoscopic, have
10

FIG. 9. Rotation photography of SC (a) (N = 3375 and λ = 0.235σ), BCC (b) (N = 4394 and λ = 0.231σ) and FCC (c)
(N = 5324 and λ = 0.257σ) crystals.

an “X”-shape pattern at the center and are made of hor- of 3/8 pitch are present. The bright level 8 signal of 3D
izontal broken stripes (Fig. 11). The two branches of samples at X = 0 is missing for 2D structures.
the “X” pattern can be viewed as scattering signals from
the two series of parallel particles on the sinusoidal wave
(Fig. 11a). Because each pitch of the helix has 10 parti- 7. SCATTERING VECTOR q IN INTENSITY
cles, the pattern has a vertical period of 10 stripes [36]. SCANNING
The brightness and darkness along each horizontal stripe
depends sensitively on the relative position between dif- In this section, we discuss the choice of scattering
ferent particles [38]. For example, the level 4 stripe dis- vector q in the case of disordered or partially ordered
appears when two double strands with a phase difference samples that are spatially isotropic or approximately
11

7.2. Random Rotation of a Small Anisotropic


Sample

The above method of using a scattering vector q at


one direction to represent that magnitude q does not
work well for simulation samples, which are usually small
and anisotropic (single crystal instead of polycrystal). To
simulate experimental results, we can fix the direction of
the incident ray but randomly rotate the small sample
to many orientations. This is done by applying a three-
dimensional rotation matrix to the original particle coor-
dinates ri , whose rotation axis is uniformally distributed
on a sphere and rotation angle is uniformally chosen from
[0, 2π]. Then the signal I(q) in Eq. (42) can be approx-
FIG. 10. System size effect on scattering size in back- imated by accumulating intensities from all those orien-
reflection of SC samples with fixed wavelength λ = 0.55σ. tations Ω
The scattering angle 2θ √
for these four spots can be computed  2 N 2 
from tan(π − 2θ) = 85 2/100. After scaling L by a factor N
√ X X X
I(q) = cos(q · ri (Ω)) + sin(q · ri (Ω))  .

of 2, the data (red squares) agree with theoretical slope 
(dashed line) from Scherrer equation λ/ cos θ.

Ω i=1 i=1
(43)

isotropic. When samples are isotropic, the scattering in- Here, ri (Ω) represents the new coordinates of particle i
tensity I(q) or its normalized version, structure factor after rotation to orientation Ω.
S(q), only depends on the magnitude q of the scattering
P
Note that the sum is applied to the intensity rather
vector, and thus does not generate isolated spotty signals Ω
2
as in photography of ordered samples. The photography P PN
I(X, Y ), often of less interest in this context, should ide- than within the square like · · · . The latter
Ω i=1
ally exhibit concentric circular patterns. The intensity choice would imply a virtual system of many randomly
scanning I(q) or S(q) as a function of q is the primary orientated overlapping grains, each of size N . The po-
method used for isotropic samples. sitions of these virtual grains generated by rotation do
not reflect the absolute positions of grains in the real
polycrystal. According to the equivalency of Eq. (9) and
7.1. Vector q along a Single Direction to Represent Eq. (10), Eq. (43) is an approximation to Eq. (42) by
Magnitude q in Isotropic Systems only considering relative positions of particles within each
grain ri (Ω)−rj (Ω). Therefore, the difference between co-
In an experiment, one can vary q by observing signals ordinates of particles i and j at two different orientations,
at continuously changing scattering angle 2θ using a fixed ri (Ω) − rj (Ω0 ), does not affect the result of Eq. (43), but
incident wavelength λ. Because experimental samples are will lead to different and wrong results, if the sum is
generally large enough, a well-averaged scattering signal
P P N
2

can be detected along one particular direction at 2θ, as taken as · · · .
Ω i=1
in the powder method with a diffractometer.
For small and nearly isotropic liquids or glasses, one
For example, consider a polycrystal with M randomly
can replace random rotations of the sample by averag-
oriented crystalline grains (domains), each of N particles.
ing over many thermally equilibrated configurations. For
The scattering intensity at q computed from Eq. (9),
anisotropic systems, however, rotations are needed to
assuming fˆi (q) = 1, is sample different directions.
2 2
XM X
N XM X
N
I(q) = cos(q · rn,i ) + sin(q · rn,i ) ,

7.3. Scattering Vector q on a Lattice
n=1 i=1 n=1 i=1
(42)
An alternative and more convenient way to simu-
where rn,i is the position vector of particle i in grain n. If late experiments is to fix the sample coordinates and
M is large and crystalline grains are uniformly oriented choose q of a given q from all directions. It is often
in all directions, I(q) at the particular vector q can be suggested to select q from a 3D orthorhombic lattice,
accurate enough to represent I(q) at the magnitude q, q = ∆q(nx , ny , nz ), with integers nx , ny , nz and incre-
without averaging over all directions of q. A similar ar- ment ∆q = 2π L , where L is the linear dimension of the
gument applies to bulk liquids or glasses, in which I(q) cubic simulation box [39]. The motivation here is that L
is also well self-averaged. sets the maximum periodicity of the simulation sample
12

FIG. 11. Transmission photography of single-strand (a,c) and double-strand (b,d) 2D sinusoidal waves (a-b) and 3D DNA
helices (c-d) using λ = 1.54σ. Each helix is made of a backbone of N = 70 particles with a pitch of p = 34σ and a diameter of
20σ. There are 10 particles per pitch. The two helices in the double strand structure are offset by 3/8 pitch. Particle size in
insets is set as 5σ to enhance visibility.

that is still physically meaningful, and thus the resolution there are 1, 2, 1, 2, 2, 1, 2,√· · · q√points
√ on the lattice at
of q. The integers nx , ny , nz may be chosen to run from magnitude q/∆q = 0, 1, 2, 2, 5, 8, 9, · · · respectively
negative to positive values to sample spherically symmet- (Fig. 12). When reporting the result of S(q), one can
ric q’s, or to start from zero to sample only q’s on 1/8 of assign q’s into bins of equal size or just use the original
the sphere. At the expense of symmetry and averaging, q values visited by the lattice points. In both cases, the
the latter choice can reach a higher magnitude q with the S(q) should be the mean value averaged over all the q’s
same number of lattice points. at that q.
There are multiple q’s on this lattice that correspond If the sample is crystalline and L is an integer multi-
to the same magnitude q, from which we can compute ple of the crystallographic lattice constant a, then the q
an average S(q). The number of q’s for a given magni- lattice contains the reciprocal lattice points of the crys-
tude q tends to, but not necessarily, increase with q. For tal (subject to a 2π factor difference) [39]. If L = 5a in
example, in a 2D system with q’s on a square lattice, the above 2D example, then q points ∆q(0, 0), ∆q(5, 0),
13

or S(q) at each q is then O(Nq N ).

ny =

a In the limit of Nq → ∞, using Eq. (10), we can inte-
8 grate over all q directions and then normalize it by the
full solid angle of 4π to compute the average I(q)
7
N P
N
1
fˆi (q)fˆj (q)eiq·rij
R P
I(q) = 4π dq
6 |q|=q i=1 j=1
2π Rπ N P
N
1
fˆi (q)fˆj (q)eiqrij cos θ
R P
5 = 4π dφ sin θdθ
0 0 i=1 j=1
N
N P
4
fˆi (q)fˆj (q) qrijij
P sin(qr )
=
i=1 j=1
3
(44)
2
This is known as Debye’s scattering equation [41, 42],
which can also be viewed as the discrete version of the

∆q = 2π
L
1
Fourier transform of the radial distribution function g(r)
0
in Eq. (33). The computational complexity of Debye’s
nx = 0 1 2 3 4 5 6 7 8 method is O(N 2 ) and it becomes more efficient than
numerically sampling Nq vector q’s on a sphere when
FIG. 12. Scattering vector q = ∆q(nx , ny ) on a square lattice N < Nq .
used for a 2D system of box size L. q points at the same
magnitude q are connected by concentric quarter circles up to
q = 8∆q. If the system is a crystal of lattice constant a = L/5, 8. PHOTOGRAPHY AND INTENSITY
then four points shown in red correspond to reciprocal lattice SCANNING OF DISORDERED OR PARTIALLY
points. ORDERED SAMPLES

Although intensity scanning, I(q) or S(q) as a function


∆q(0, 5), ∆q(5, 5) correspond to reciprocal lattice points of q, generally gives more useful structural information
(0, 0), (1/a, 0), (0, 1/a), (1/a, 1/a) respectively (Fig. 12). about isotropic samples, it is sometimes interesting to
These lattice points are where Bragg’s law Eq. (19) are show the corresponding photography I(X, Y ). In fact,
obeyed. Therefore, according to the discussion in Sec- intensity scanning can be obtained from photography by
tion 3 3.3, if all atomic form factors are unity, I(q) = moving along a specific radial direction on the (X, Y )
|Fhkl |2 = N 2 and S(q) = N at each of these reciprocal film, as in the early days of the powder method [43].
lattice points. The intensity scanning result S(q) needs To generate scattering photography of isotropic sam-
to be an average over all q points at that q, some of which ples, we use the rotation or thermal averaging method of
are not reciprocal lattice points and thus have S(q) = 0. Section 7 7.2. Intensity scanning profiles are calculated
For example, at q = 5∆q of the 2D system, two points using the three methods mentioned in Section 7 7.3 and
have S(q) = N and two have S(q) = 0. The average Section 7 7.4.
S(q = 5∆q) is thus (N + N + 0 + 0)/4 = N/2 (Fig. 12).
Using lattice points to approximate q’s from all direc-
tions is problematic, when L is small and thus increment 8.1. Liquids and Glasses
∆q is large such that only a few q’s are available at each
q. The issue is more severe at small q or towards corners If a scattering photograph is taken for disordered sam-
of the cubic lattice at high q. The calculated signal S(q) ples like liquids or glasses using a fixed wavelength,
can then be quite noisy because q is not averaged enough a characteristic ring signal is expected at peak value
over all directions. q ∗ ∼ 2πσ that corresponds to the molecular size σ. This
ring is regular and clear, when the sample, like most ex-
perimental bulk samples, is large enough such that a good
7.4. Scattering Vector q on a Sphere and Debye’s average is taken within the system in the calculation of
Scattering Equation I(q). However, in a small simulation system (N = 103 -
104 ), photography of one static disordered sample gives
In order to obtain a smooth curve of S(q) that bet- spotty and noisy signals with certain traces of ring fea-
ter matches experimental results, we need to use enough tures (Fig. 13a). To enhance sharpness of the ring, one
spherically distributed q’s. To guarantee uniform dis- can either increase the size N of the sample or take
tribution of points on a sphere, we apply the Fibonacci the ensemble average of I(q) over many configurations
grid approach to randomly chose Nq scattering vectors (Fig. 13c).
q’s from a sphere of radius q [40]. Increasing Nq improves For homogeneous liquids and glasses, the static struc-
the effect of averaging. The complexity to compute I(q) ture factor S(q) = S(q) varies only with the magnitude
14

FIG. 13. Simulated transmission photography using fixed wavelength λ = 0.4σ (a,c) and structure factor S(q) (b,d) of a
static glass sample (a-b) and a thermally averaged liquid sample (c-d). Red solid rings in photography correspond to the
first and second peaks in S(q). The cross pattern at the center of each photograph is due to Fraunhofer diffraction from
the small simulation box, effectively a cubic obstacle. Three methods are used to compute S(q): with q’s on a cubic lattice
(green circles), with q’s on spheres (blue dotted line) and Debye’s scattering equation (red solid line). Insets show the radial
distribution function g(r). Both samples are N = 1000 hard spheres of diameter σ. The glass sample has one configuration at
packing fraction 0.64. The liquid sample has 1000 thermally equilibrated configurations at packing fraction π/6 = 0.5236.

q of the scattering vector and exhibits a major peak at rect orientation by chance to satisfy Bragg’s law. The
q ∗ ∼ 2π
σ . Using q’s on a sphere numerically or Debye’s measured intensity scanning S(q) can be used to calcu-
equation can generate well-averaged smooth S(q) curves late interplanar spacings in the crystal and, with some
for liquids or glasses (Fig. 13b,d). If only one disordered limitations, even to determine the crystal structure.
configuration is analyzed, the S(q) curve is much noisier It is difficult to produce a well randomized polycrys-
using q’s on a cubic lattice (Fig. 13b). talline sample in simulation, given the limit of system
size. Nevertheless, we can start from a small single crys-
tal sample and use random rotation or q’s from different
8.2. Polycrystalline Samples – Powder Method directions to simulate scattering signals of a polycrystal.
In particular, we use the same SC, BCC, FCC crystals
The powder method is often used to analyze polycrys- used above to generate photography and intensity scan-
tals, in which a crystalline sample is ground into pow- ning results of corresponding polycrystals (Fig. 14).
der to produce many small randomly oriented crystalline The sharp concentric rings in the photograph I(X, Y )
grains. Then, at any scattering angle 2θ where strong and the narrow peaks in S(q) correspond to scatter-
signal is expected, at least one of the grains has the cor- ing from different crystallographic planes (hkl) of the
15

FIG. 14. Simulated powder method. Transmission photography using fixed wavelength λ = 0.4σ (a,c,d) and structure factor
S(q) (b,d,f) of polycrystalline SC (a-b), BCC (c-d) and FCC (e-f) samples. Miller indices (hkl) are labelled next to each signal
peak. The photograph is produced by randomly rotating a single crystalline sample in three dimensions and taking the average
of I(q) over 5000 orientations. The small concentric circular pattern at the center of each photograph is due to Fraunhofer
diffraction from the small simulation box, effectively a circular obstacle, after being randomly rotated. Three methods are used
to compute S(q): with q’s on a cubic lattice (green vertical lines), with q’s on spheres (blue dotted line) and Debye’s scattering
equation (red solid line).


three crystals (SC, BCC, and FCC). S(q) peaks com- and N = L3 = 3375 particles. Given ∆q = 15σ ,
puted from spherically distributed q’s are lower and the S(q) peak from (100) planes is expected to occur
broader than those from cubic lattice q’s. The peak at six q points, ∆q(15, 0, 0), ∆q(0, 15, 0), ∆q(0, 0, 15),
height using cubic lattice q’s often scales with system ∆q(−15, 0, 0), ∆q(0, −15, 0), and ∆q(0, 0, −15), each has
size N . For example, the SC crystal has L = 15σ a value S(q) = 3375. However, there are other q points
16

with magnitude q = 15∆q, which correspond to integer (110) peak from 7.7σ −1 to 1.27σ −1 (Fig. 15c). This helps
solutions to n2x + n2y + n2z = 152 . In total, at q = 15, there us to identify that only the peak from the (110) planes
are 6 (15, 0, 0)-like (considering its permutation and ±), of the BCC superlattice is sharply distinguishable from
24 (12, 9, 0)-like, 24 (10, 10, 5)-like, 48 (11, 10, 2)-like, and the background signals.
48 (14, 5, 2)-like q points. Out of these 150 points, only 6
have S(q) = N while others have S(q) = 0. So the peak
height S(q = 15∆q) = 3375 × 6/150 = 135. 9. 2D STRUCTURE FACTOR

For 2D samples or 2D projection of 3D samples, it is


8.3. Mesophases – Small-Angle Method sometimes useful to express S(q) as a 2D function of
(qx , qy ) [49] or scattering angles (θx , θy ) [50]. The 2D
Mesophases are states of matter intermediate between structure factor S(qx , qy ) is related to the photography
liquids and solids found in block copolymers [44], liquid I(X, Y ) by converting coordinates (X, Y ) on the film
crystals [45], structural DNAs [46], etc., which present into components (qx , qy ) of the scattering vector using
mesoscopic ordering of length scales larger than molec- Eq. (36). For 3D structures, the component qz can be
ular size σ. To detect these long-wavelength structures, expressed as a function of qx and qy , for example, in the
small-angle X-ray scattering (SAXS) [11] or small-angle case of the transmission method (Eq. (38)),
neutron scattering (SANS) [47] methods are needed be-
cause scattering signals are expected at small q (before 2π
qz = − (1 − D/L)
the first major diffraction peak ∼ 2π σ ) thus small θ as
λ
seen from Eq. (2). A logarithmic scale axis is often set
s !
2π λ2 2 λ2 2
for S(q) in the structure factor plot because at q → 0 the =− 1− 1− qx − q (45)
λ (2π)2 (2π)2 y
signal scales with system size N [48].
Experimental mesophases are often synthesized as
s 2
2π 2π
polycrystals or many small crystalline domains randomly =− + − qx2 − qy2 .
embedded in an amorphous matrix, for which intensity λ λ
scanning S(q) at small q is used to reveal the ordering.
We illustrate the concepts of the mesophase structure Note that knowing (qx , qy ) does not uniquely determine
factor using a lamellar, a cylindrical and a BCC spheri- qz . The constant 2πλ still needs to be specified. For small-
cal configuration of domains, cut from a disordered glass angle scattering with qx , qy → 0, an approximation to set
sample of hard spheres of diameter σ. These structures qz = 0 is valid if there is no long-range periodicity along
are thus amorphous within each domain, but the domains the z direction.
form a 1D, 2D or 3D superlattice for the lamellar, cylin- We compute S(qx , qy ) for the cylindrical mesophase in
drical or spherical configuration, respectively. Fig. 15b, whose cylinder axis is aligned with the incident
In the lamellar phase, each period is of length d = 5σ ray in the z direction. We first use qz calculated from
consisting of a layer with thickness 3.5σ and a gap with Eq. (45) with λ = 0.4σ. Besides the isotropic circular
thickness 1.5σ. We find three peaks of S(q) at one time, signal corresponding to particle size σ, a characteristic
two times and three times 2π −1 hexagonal pattern with six-fold symmetry is observed at
d ≈ 1.257σ , corresponding
to the first, second and third order of Bragg diffraction small q, which results from the cylinders packed on a 2D
of the superlattice (Fig. 15a). The peak height drops triangular lattice. We can identify two sets of spots on
as q increases, and when layer thickness equals the gap the vertices of hexagons – one corresponds to the unit
thickness, peaks at even multiples of 2π cell with spacing d1 and the other corresponds to the
d disappear.
The cylindrical phase with a disk radius 1.8σ resides unit cell with spacing d2 (Fig. 16a). The second-order
on a two-dimensional triangular superlattice with lattice peak related to d1 and first-order peak related to d2 form
constant 5σ. By assigning unit cells√in two different ways a hexagon together, while the first-order peak related to
d1 is mixed with the Fraunhofer diffraction pattern at
with interplanar spacing d1 = 5 2 3 σ and d2 = 2.5σ,
−1
smaller q.
we can identify two peaks at q1 = 2π d1 ≈ 1.451σ and If we set qz = 0, the S(qx , qy ) pattern is approximately
2π −1
q2 = d2 ≈ 2.513σ (Fig. 15b). The second-order peak the same at small q, with a certain degree of enhancement
around 2q1 ≈ 2.9σ −1 is also visible (not marked). (Fig. 16b). Some higher-order peaks become visible at
The spherical phase has spheres of radius 2σ that pack larger q.
on a BCC superlattice with a lattice constant 7σ. If
each sphere domain has only one particle, the structure
factor would be the same as a normal BCC crystal apart 10. CONCLUSION
from a change of unit for q. We can obtain S(q) of this
one-particle spherical phase by rescaling the q axis of In this article, we give a comprehensive and co-
S(q) of the BCC crystal, √ which has a lattice constant herent review of core concepts about scattering meth-
2 7 3
a = 3 σ, by a factor of 2 ≈ 6.062. This moves the
√ ods used to determine the structures of ordered and
17

FIG. 16. 2D structure factor S(qx , qy ) of the cylindri-


cal mesophase using λ = 0.4σ with (a) qz = − 2π λ
+
q
2π 2

λ
− qx − qy and (b) qz = 0. The red solid ring marks
2 2

the broad peak corresponding to particle diameter σ at ∼ 2π σ


.
The red arrow points to the second-order peak from the unit
cell with spacing d1 . The green arrow points to the first-order
peak from the unit cell with spacing d2 .
FIG. 15. Simulated small-angle structure factor S(q) for (a)
lamellar, (b) cylindrical and (c) BCC spherical mesophases.
Three methods are used to compute S(q): with q’s on a cubic
lattice (green vertical lines), with q’s on spheres (blue dotted disordered samples. Scattering photography and in-
line) and Debye’s scattering equation (red solid line). Debye’s tensity scanning of typical examples are calculated
result of S(q) of a homogeneous glass, after vertically rescaled that can be used as benchmarks. Sample CPU
to align at q → 0, is shown for comparison (black dashed codes are provided on GitHub at https://github.com/
line). Black downward arrows mark signature peaks for each statisticalmechanics/scatter to illustrate the math-
structure. The broad peak at 7.5σ −1 corresponds to particle
ematics and algorithms. Accelerating GPU codes that
size σ. Insets show top/side views of the configurations under
consideration. Blue solid line in (c) is Debye’s result of S(q)
can reduce hours of computation to seconds are also pro-
for a BCC sphere mesophase with one particle per domain, vided for efficient simulation of scattering signals. It
obtained by rescaling the curve of a BCC crystal. should be noted that for simplicity, the intensity calcu-
lation discussed in this paper has omitted serveral im-
18

portant wavelength and/or angle-dependent factors due The inverse Fourier transform of F̂k is an integral in
to, for example, absorption, extinction, multiple scatter- the wavevector space which gives the original real-space
ing, polarization, and the Lorentz factor, as well as the function
issue of normalization of the measured intensity to an
1
Z
absolute scale. We have also omitted the effect of tem- F (r) = dkF̂k e−ik·r . (A2)
perature which generally adds a Gaussian co-factor to (2π)3
each atomic scattering factor. This expands F (r) in terms of an infinite number of pe-
riodic basis functions e−ik·r characterized by different
k’s. The coefficient or contribution of each k is just the
ACKNOWLEDGMENTS Fourier transform F̂k . In principle, the collection of all
F̂k ’s contains the entire information about the original
This work benefits from the Duke Kunshan startup function F (r) such that knowing F̂k ’s allows us to recon-
funding and resources made available at the Duke Com- struct F (r).
pute Cluster (DCC). We thank Corey O’Hern, Alex Gri- In physical systems, F (r) is often defined within a fi-
gas, Robert Hoy and Joseph Dietz for testing some of our nite volume V and the Fourier transform should be inte-
codes. We also thank Patrick Charbonneau for helpful grated over the region V
discussions. Z
F̂k = drF (r)eik·r . (A3)
V
Appendix A: Fourier Transform: Continuous and
Discrete If such a finite system is of a cubic shape with a linear
dimension L, i.e. V = L3 , then any periodicity or wave-
The Fourier transform F̂k of a function F (r) defined length λ > L is unphysical. This imposes a lower bound,
continuously in three-dimensional real space of infinite 2π/L, on the smallest wavevector to be considered. The
volume is inverse Fourier transform Eq. (A2) thus should not vary k
continuously as in an integral, but only take discrete val-
ues of k with increments (∆kx , ∆ky , ∆kz ) = ( 2π 2π 2π
Z
F̂k = drF (r)eik·r , (A1) L , L , L ).
The integral then becomes [52]
 3
where k is a wavevector used to extract the spatial pe- 1 X −ik·r 2π 1 X
riodicity of F (r) [51]. For instance, if F (r) has a pe- F (r) = F̂ k e = F̂k e−ik·r .
(2π)3 L V
k k
riodic pattern of wavelength λ along the x axis, i.e.
(A4)
F (x, y, z) = F (x + λ, y, z), then the value of F̂k is large
for the k of magnitude |k| = 2π/λ pointing in the x direc- Mathematically, for the Fourier transform Eq. (A1) to
tion, i.e. k = (2π/λ, 0, 0). Physically, if eik·r is viewed as exist, the function F (r) needs to be absolutely integrable.
a plane wave traveling in the k direction, then F̂k would If F (r) equals to some nonzero constants, or without loss
exhibit a peak value, when F (r) has wavelike properties of generality, F (r) = 1, in order to reconcile the singular-
coherent with eik·r such that they add constructively in ity, the result of the
R Fourier transform is formally written
the integral. In this sense, the Fourier transform Eq. (A1) as (2π)3 δD (k) = dreik·r , or equivalently,
quantifies the existence and the extent of periodicity cor-
1
Z
responding to k in F (r).
δD (k) = dreik·r , (A5)
In general, even if F (r) is a real function, F̂k can be (2π)3
complex. However, if F (r) is real (F ∗ (r) = F (r)) and
even (F (−r) = F (r), i.e. with a symmetry center), its where δD (x) is the (three-dimensional) singular Dirac
Fourier transform F̂k is also real and even, because the delta function (δD (0) → ∞). Usually,
R the Dirac delta
conjugate of F̂k is function with the property that dxδD (x)f (x) = f (0) is
introduced as the limiting case of a normalized Gaussian
Z Z function with its standard deviation approaching zero.
F̂k∗ = drF ∗ (r)e−ik·r = drF (r)e−ik·r = F̂−k According to the above notation, the inverse Fourier
Z ∞ Z −∞ transform of the Dirac delta function, readily reduces to
r0 =−r 1 3
δD (k)e−ik·r = e−i0·r = 1. For a system
0
R
= drF (−r)e−ik·r = − dr0 F (r0 )eik·r (2π)3 dk(2π)
−∞
Z ∞
∞ of a finite volume V , it is also customary to write
0
= dr0 F (r0 )eik·r = F̂k .
Z
−∞ dreik·r = V δk,0 , (A6)
V
Here the integration limits for the vector variable r, for-
mally denoted as ±∞, are to be understood as for each where δi,j = 1, i = j and 0, i 6= j is the Kronecker delta
of its components. function.
19

Appendix B: Direct and Reciprocal Lattices space spanned by the reciprocal lattice vectors a∗ , b∗ , c∗ ,
which are related to the direct lattice vectors by
The position vector r of particles or atoms residing on
a∗ = (b × c)/Vcell
a crystal lattice, the direct lattice, can be expressed as a
linear combination, b∗ = (c × a)/Vcell
c∗ = (a × b)/Vcell
r = xa + yb + zc,
Since a∗ is orthogonal to (b, c), b∗ is orthogonal to (a, c)
of the (direct) lattice vectors a, b, c, which are basis vec- and c∗ is orthogonal to (a, b),
tors of the unit cell with volume Vcell = a·(b×c). Gener-
ally, a, b, c may not be orthogonal to each other and thus a∗ · a = 1, a∗ · b = 0, a∗ · c = 0, etc.
x, y, z are not necessarily the projections of r in a Carte-
sian coordinate system. If particles coincide with lattice Note that, in general, a∗ , b∗ , c∗ are not orthogonal to
points, then x, y, z are integers; if partices are contained each other. Positions of reciprocal lattice points can be
inside the unit cell, their coordinates x, y, z can be frac- represented by vectors of the form
tions [53].
d∗hkl = ha∗ + kb∗ + lc∗
Particles on regular crystal lattices are situated on dif-
ferent families of parallel crystallographic planes, when where h, k, l are integers (Fig. 17b).
viewed from different angles. Such parallel planes are de- In crystallography, as the notation here implies, the
noted by three integers (hkl), the Miller indices, whose physical meaning of the reciprocal lattice is related to
reciprocals are proportional to the intercepts of the lattice planes in the direct space as follows [54]:
planes with the three axes of the direct lattice. The spac-
ing or distance, dhkl , between neighboring lattice planes • Each point with a vector d∗hkl on the reciprocal lat-
in the family (hkl) is a function of the Miller indices and tice represents a family of lattice planes with Miller
lattice parameters (Fig. 17a). In the special case of an indices (hkl);
orthorhombic lattice,
• The direction of d∗hkl is perpendicular to (or normal
1 h2 k2 l2 to) the lattice planes (hkl);
= + 2 + 2.
d2hkl a2 b c • The magnitude of d∗hkl is equal to the reciprocal of
the interplanar spacing dhkl , i.e |d∗hkl | = 1/dhkl .
The reciprocal lattice is defined mathematically in a

[1] JG Powles, “The structure of molecular liquids by neu- Joseph E Curtis, Mark Dadmun, Paul Fenimore, et al.,
tron scattering,” Advances in Physics 22, 1–56 (1973). “Neutron scattering in the biological sciences: progress
[2] WI Goldburg, “Dynamic light scattering,” American and prospects,” Acta Crystallographica Section D: Struc-
Journal of Physics 67, 1152–1160 (1999). tural Biology 74, 1129–1168 (2018).
[3] W Xiong, A Gasparian, H Gao, D Dutta, M Khandaker, [10] Ryong-Joon Roe and RJ Roe, Methods of X-ray and
N Liyanage, E Pasyuk, C Peng, X Bai, L Ye, et al., “A Neutron Scattering in Polymer Science, Vol. 739 (Oxford
small proton charge radius from an electron–proton scat- University Press, New York, 2000).
tering experiment,” Nature 575, 147–150 (2019). [11] Benjamin Chu and Benjamin S Hsiao, “Small-angle x-ray
[4] Leonid V. Azaroff, Elements of X-Ray Crystallography scattering of polymers,” Chemical Reviews 101, 1727–
(McGraw-Hill Companies, New York, 1968). 1762 (2001).
[5] Teresa Head-Gordon and Greg Hura, “Water structure [12] Jörg Fink, E Schierle, E Weschke, and J Geck, “Reso-
from scattering experiments and simulation,” Chemical nant elastic soft x-ray scattering,” Reports on Progress
Reviews 102, 2651–2670 (2002). in Physics 76, 056502 (2013).
[6] Francesco Sette, Michael H Krisch, Claudio Masciovec- [13] Feng Liu, Michael A Brady, and Cheng Wang, “Resonant
chio, Giancarlo Ruocco, and Giulio Monaco, “Dynamics soft x-ray scattering for polymer materials,” European
of glasses and glass-forming liquids studied by inelastic Polymer Journal 81, 555–568 (2016).
x-ray scattering,” Science 280, 1550–1555 (1998). [14] Subhrangsu Mukherjee, Andrew A Herzing, Donglin
[7] JB Hayter and J Penfold, “Determination of micelle Zhao, Qinghe Wu, Luping Yu, Harald Ade, Dean M De-
structure and charge by neutron small-angle scattering,” Longchamp, and Lee J Richter, “Morphological char-
Colloid and Polymer Science 261, 1022–1030 (1983). acterization of fullerene and fullerene-free organic pho-
[8] John C Kendrew, “The three-dimensional structure of tovoltaics by combined real and reciprocal space tech-
a protein molecule,” Scientific American 205, 96–111 niques,” Journal of Materials Research 32, 1921 (2017).
(1961). [15] Garry J McIntyre, “Area detectors in single-crystal neu-
[9] Rana Ashkar, Hassina Bilheux, Heliosa Bordallo, Robert tron diffraction,” Journal of Physics D: Applied Physics
Briber, David JE Callaway, Xiaolin Cheng, X-Q Chu, 48, 504002 (2015).
20

Å for atomic systems, nm for nano-systems, and µm for


(a) colloidal systems.
[23] Donald A. McQuarrie and John D. Simon, Physical
Chemistry: A Molecular Approach (University Science
d120 Books;, New York, 1997).
[24] Lawrence Bragg, “X-ray crystallography,” Scientific
American 219, 58–74 (1968).
[25] Christopher Hammond, The Basics of Crystallography
b and Diffraction (Oxford Science Publications, Oxford,
2001).
[26] Leonard J. Barbour, “Ewaldsphere: an interactive ap-
a (120) proach to teaching the ewald sphere construction,” Jour-
nal of Applied Crystallography 51, 1734–1738 (2018).
(b) [27] CD Thomas and NS Gingrich, “Presentation of the con-
cept of liquid structure,” American Journal of Physics 9,
10–13 (1941).
[28] AZMS Rahman, KS Singwi, and A Sjölander, “Theory
of slow neutron scattering by liquids. i,” Physical Review
d∗120 126, 986 (1962).
(210) [29] Liesbeth Janssen, “Mode-coupling theory of the glass
b∗ (110)
transition: A primer,” Frontiers in Physics 6, 97 (2018).
[30] Henry E Fischer, Adrian C Barnes, and Philip S Salmon,
“Neutron and x-ray diffraction studies of liquids and
(1̄1̄0) (100) glasses,” Reports on Progress in Physics 69, 233 (2006).
[31] Simon J. L. Billinge, “21st century challenges: Structure
a∗ beyond crystals,” IUCr NEWSLETTER 27 (2019).
[32] J. Hansen and I. R. McDonald, Theory of Simple Liquids
with Applications to Soft Matter (Academic Press, New
York, 2013).
[33] David A Keen, “A comparison of various commonly
used correlation functions for describing total scatter-
ing,” Journal of Applied Crystallography 34, 172–177
FIG. 17. (a) Direct and (b) reciprocal lattices. The (120)
(2001).
planes (dashed lines) with interplanar distance d120 in direct
[34] J. Barrat and J. Hansen, Basic Concepts for Simple
space correspond to the point denoted by the vector d∗120 in
and Complex Liquids (Cambridge University Press, New
the reciprocal space. d∗120 is normal to the (120) planes and
York, 2003).
|d∗120 | = 1/d120 .
[35] Rosalind E. Franklin, “Molecular configuration in sodium
thymonucleate,” Nature 171, 740–741 (1953).
[36] C Kittel, “X-ray diffraction from helices: Structure of
[16] Greg L Hura, Angeli L Menon, Michal Hammel, Robert P dna,” American Journal of Physics 36, 610–616 (1968).
Rambo, Farris L Poole Ii, Susan E Tsutakawa, Francis E [37] J Thompson, G Braun, D Tierney, L Wessels,
Jenney Jr, Scott Classen, Kenneth A Frankel, Robert C H Schmitzer, B Rossa, HP Wagner, and W Dultz, “Ros-
Hopkins, et al., “Robust, high-throughput solution struc- alind franklin’s x-ray photo of dna as an undergradu-
tural analyses by small angle x-ray scattering (saxs),” ate optical diffraction experiment,” American Journal of
Nature Methods 6, 606–612 (2009). Physics 86, 95–104 (2018).
[17] Herbert A Hauptman, “The phase problem of x-ray crys- [38] Amand A Lucas and Philippe Lambin, “Diffraction by
tallography,” Reports on Progress in Physics 54, 1427 dna, carbon nanotubes and other helical nanostructures,”
(1991). Reports on Progress in Physics 68, 1181 (2005).
[18] It should be noted that the diffraction experiment gen- [39] M. P. Allen and D. J. Tildesley, Computer Simulation of
erally detects both elastic and inelastic scattering, where Liquids (Oxford University Press, New York, 1987).
the latter results from dynamic processes in the sample. [40] Edward B Saff and Amo BJ Kuijlaars, “Distributing
To measure just elastic scattering, an energy analyser many points on a sphere,” The Mathematical Intelli-
should be placed between the sample and detector. gencer 19, 5–11 (1997).
[19] Patrick Argos, “Protein crystallography in a molecular [41] Noel William Thomas, “A new approach to calculating
biophysics course,” American Journal of Physics 45, 31– powder diffraction patterns based on the debye scatter-
37 (1977). ing equation,” Acta Crystallographica Section A: Foun-
[20] Robert W Harrison, “Phase problem in crystallography,” dations of Crystallography 66, 64–77 (2010).
J. Opt. Soc. Am. A 10, 1046–1055 (1993). [42] Luca Gelisio and Paolo Scardi, “100 years of debye’s
[21] Garry Taylor, “The phase problem,” Acta Crystallo- scattering equation,” Acta Crystallographica Section A:
graphica Section D: Biological Crystallography 59, 1881– Foundations and Advances 72, 608–620 (2016).
1890 (2003). [43] MU Cohen, “Precision lattice constants from x-ray pow-
[22] Throughout the paper, we use σ as the unit of length and der photographs,” Review of Scientific Instruments 6, 68–
1/σ as the unit of wavevector. We choose σ to be particle 74 (1935).
diameter, which can be mapped onto the length scale of
21

[44] SHINICHI Sakurai, S Okamoto, TEXSUYA Kawamura, W540–W544 (2010).


and T Hashimoto, “Small-angle x-ray scattering study of [49] U. Tutsch, B. Wolf, S. Wessel, L. Postulka, Y. Tsui,
lamellar microdomains in a block copolymer,” Journal of H.O. Jeschke, I. Opahle, T. Saha-Dasgupta, R. Valenti,
Applied Crystallography 24, 679–684 (1991). A. Bruhl, and K. Removic-Langer, “Evidence of a
[45] D John Mitchell, Gordon JT Tiddy, Loraine Waring, field-induced berezinskii-kosterlitz-thouless scenario in a
Theresa Bostock, and Malcolm P McDonald, “Phase two-dimensional spin-dimer system,” Nature Communi-
behaviour of polyoxyethylene surfactants with water. cations 5, 5169 (2014).
mesophase structures and partial miscibility (cloud [50] Byeongdu Lee, Insun Park, Jinhwan Yoon, Soojin Park,
points),” Journal of the Chemical Society, Faraday Jehan Kim, Kwang-Woo Kim, Taihyun Chang, and
Transactions 1: Physical Chemistry in Condensed Phases Moonhor Ree, “Structural analysis of block copolymer
79, 975–1000 (1983). thin films with grazing incidence small-angle x-ray scat-
[46] Ye Tian, Julien R Lhermitte, Lin Bai, Thi Vo, tering,” Macromolecules 38, 4311–4323 (2005).
Huolin L Xin, Huilin Li, Ruipeng Li, Masafumi Fukuto, [51] MJ Lighthill, An Introduction to Fourier Analysis and
Kevin G Yager, Jason S Kahn, et al., “Ordered three- Generalized Functions (Cambridge University Press,
dimensional nanomaterials using dna-prescribed and Cambridge, 1958).
valence-controlled material voxels,” Nature Materials , [52] Paul M Chaikin, Tom C Lubensky, and Thomas A Wit-
1–8 (2020). ten, Principles of Condensed Matter Physics (Cambridge
[47] Randal W Richards and JL Thomason, “Small-angle neu- University Press, Cambridge, 1995).
tron scattering study of block copolymer morphology,” [53] Donald E Sands, Introduction to Crystallography (Dover
Macromolecules 16, 982–992 (1983). Publications, Inc., New York, 1993).
[48] Dina Schneidman-Duhovny, Michal Hammel, and An- [54] Nan-Xian Chen, “An elementary method for introducing
drej Sali, “Foxs: a web server for rapid computation the concept of reciprocal lattice,” American Journal of
and fitting of saxs profiles,” Nucleic Acids Research 38, Physics 54, 1000–1002 (1986).

You might also like