You are on page 1of 76

Crystallography Reviews

ISSN: 0889-311X (Print) 1476-3508 (Online) Journal homepage: https://www.tandfonline.com/loi/gcry20

X-ray scattering characterisation of nanoparticles

Bridget Ingham

To cite this article: Bridget Ingham (2015) X-ray scattering characterisation of nanoparticles,
Crystallography Reviews, 21:4, 229-303, DOI: 10.1080/0889311X.2015.1024114

To link to this article: https://doi.org/10.1080/0889311X.2015.1024114

Published online: 02 Jun 2015.

Submit your article to this journal

Article views: 4106

View related articles

View Crossmark data

Citing articles: 47 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=gcry20
Crystallography Reviews, 2015
Vol. 21, No. 4, 229–303, http://dx.doi.org/10.1080/0889311X.2015.1024114

REVIEW

X-ray scattering characterisation of nanoparticles


Bridget Ingham ∗

Callaghan Innovation, P.O. Box 31310, Lower Hutt 5040, New Zealand
(Received 25 November 2014; accepted 25 February 2015)

This article provides, in tutorial style, a review of X-ray scattering methods commonly used
to characterise nanoparticles and gives numerous case studies of basic science and industrial
applications right up to the present. It is divided into two major sections, broadly covering X-
ray diffraction and small-angle X-ray scattering. Each section begins with a brief introduction
to the technique and the information that can be obtained by using it, followed by a discussion
on experimental considerations, with a particular focus on nanoparticle characterisation. The
techniques and analysis methods are demonstrated by way of examples of a wide variety of
nanoparticle materials and synthesis methods. Recent advances in related techniques such as
anomalous scattering and pair distribution function analysis are also described and discussed.
Keywords: nanoparticles; X-ray diffraction (XRD); small-angle X-ray scattering (SAXS);
synchrotron radiation; sample cells; In situ experiments; Ex situ experiments; industrial appli-
cations; catalysis studies; optical properties; nanoparticle kinetics; nanoparticle magnetism;
electrical nanowires; optoelectronic properties; high pressure studies; resonant scattering
studies; fractal and dendrimer aggregates; pair distribution function analysis

Contents PAGE

1. Introduction 230

2. X-ray diffraction 232


2.1. X-ray diffraction experimental considerations 235
2.2. Ex situ XRD studies 238
2.2.1. Ex situ XRD studies of metal nanoparticles 238
2.2.2. Ex situ XRD analysis on oxide systems 240
2.2.3. Ex situ XRD analysis on other systems 243
2.3. In situ XRD studies 245
2.3.1. In situ XRD during synthesis of metal nanoparticle systems 245
2.3.2. In situ XRD during post-synthesis processing of metal nanoparticle
systems 249
2.3.3. In situ XRD during synthesis and processing of oxide nanoparticle
systems 252
2.3.4. Synthesis and processing of other systems 255
2.4. Anomalous XRD 256
2.5. Pair distribution function analysis 258
2.5.1. Examples 263

3. Small-angle X-ray scattering 266

*Email: bridget.ingham@callaghaninnovation.govt.nz


c 2015 Callaghan Innovation
230 B. Ingham

3.1. SAXS experimental considerations 269


3.2. Ex situ SAXS studies 270
3.2.1. Ex situ SAXS studies of metal nanoparticles 270
3.2.2. Ex situ SAXS studies of other systems 272
3.3. In situ SAXS studies 273
3.3.1. In situ SAXS during synthesis of metal nanoparticle systems 273
3.3.2. In situ SAXS during post-synthesis processing of metal
nanoparticle systems 277
3.3.3. In situ SAXS during synthesis of oxide systems 279
3.4. Anomalous SAXS 280
3.5. SAXS of ordered nanoparticle assemblies 282

4. Future directions 286

5. Conclusion 287

Disclosure statement 287

Notes on contributor 287

ORCID 287

References 287

Subject index 303

1. Introduction
Nanoparticle research is a rapidly growing field, with many potential applications from medicine
to magnetism, catalysis, solar cells, and more. In the medicine field, nanoparticles are used
as drug delivery vehicles [1–7] and magnetic resonance imaging (MRI) contrast agents.[5–9]
They can be used to treat cancer via cell apoptosis [10] or via magnetic [5–7] or photo-induced
hyperthermia (localised heating).[3,11] They are used as biosensors,[4,12] in biodiagnostic
assays,[13–15] and in magnetic separation.[7] Nanoparticles are widely used in surface-enhanced
Raman scattering (SERS),[16,17] enabling detection of single molecules.[17–21] SERS can
be used for medical diagnostics [22–24] and targeting tumours.[22,23] Magnetic nanoparticles
also have applications in high-density data storage,[4,5,25–28] spintronics,[4,5] and magnetic
energy storage.[8] Many metallic and bimetallic nanoparticles have catalytic properties, which
in many cases are superior to the bulk material.[12,13,29–37] Other applications of nanoparticles
include energy conversion and storage,[37,38] solar cells,[37–39] battery materials,[36–38] and
the possibility of bottom-up fabrication of new materials.[40]
The physical and chemical properties of a nanoparticle strongly depend on its size and
shape. Semiconducting nanoparticles, also known as ‘quantum dots’, exhibit tuneable optical
and electronic properties depending on their size due to quantum confinement effects.[41,42]
The catalytic properties of metallic nanoparticles can be highly size and shape dependent [29]
due to the molecular selectivity of the terminating crystallographic planes.[30]
X-ray scattering techniques are useful for characterising the crystal and particle size and the
crystallographic phase – all of which determine the physical properties of the particle. X-ray
techniques are generally non-destructive. They enable information about the ensemble aver-
age of many particles to be obtained, in contrast to direct imaging techniques such as electron
microscopy where only a very small sample of particles are studied that may not be truly repre-
sentative of the material, for example, where there is moderate to severe polydispersity in particle
Crystallography Reviews 231

size. X-ray techniques provide direct measures of particle size and lattice dimensions, in contrast
to other indirect methods such as UV–visible spectroscopy, where the particle size is inferred
from the systematic shift in the position of the absorption peak.[43–45] X-ray diffraction (XRD)
provides information about the arrangement of atoms within a crystalline material and can detect
non-bulk structures that may be observed in very small nanoparticles, for example, structures
such as icosahedra and decahedra; or phases that are only metastable in the bulk material, which
may offer entirely new physical properties.
X-ray scattering of nanoparticles is much weaker than for bulk materials due to the small
volume of the particles and limited coherence due to their small size. To increase the signal,
samples must be measured either over long times or with high-flux sources. When comparing
X-ray sources, the important factor is not flux (photons per second) alone, but rather brilliance,
which is the flux per unit area per solid angle – in other words, the flux that is actually focused
onto the sample. Laboratory sources typically operate by ejecting electrons from a hot filament
and accelerating them onto an anode material. X-rays arise from fluorescence which is of a par-
ticular energy depending on the anode element. Standard sealed tube sources with a fixed anode
have typical brilliance values of around 108 photons s−1 mm−2 mrad−2 , while rotating anode and
microfocus anodes can be an order of magnitude higher.[46] The maximum brilliance obtainable
from these anodes is limited by the need to dissipate heat from the anode. The current state-of-
the-art laboratory X-ray sources use a molten liquid jet as the anode and can achieve brilliance
values of up to 2.6 × 1010 photons s−1 mm−2 mrad−2 .[46] This is approaching the brilliance of
a second-generation synchrotron source. Third-generation synchrotron sources typically have
brilliances of around 1016 –1020 photons s−1 mm−2 mrad−2 .[47] Synchrotron sources have an
additional benefit in that the wavelength can be tuned to a specific value, whereas laboratory
sources are limited to X-ray fluorescence lines of available anode materials. While synchrotron
radiation can be used to measure X-ray scattering of nanoparticles using short exposures (which
in some cases can be real time, allowing in situ processes to be studied), a modern laboratory
source can be quite sufficient.
While high-flux sources provide superior signal-to-noise, they can have a detrimental effect
on the sample due to radiation damage. This is particularly the case for organic molecules, such
as surfactants that may be present as stabilisers on the nanoparticles’ surfaces, or polymers being
used as a matrix into which nanoparticles may be embedded. Therefore, there is almost always
a trade-off between minimising the exposure time and maximising the signal quality. Radiation
damage can be mitigated by surrounding the sample in an inert atmosphere, performing the
measurement at cryogenic temperatures, or translating the sample during the measurement.
The following techniques are beyond the scope of this review. As mentioned earlier, electron
microscopy is a widely used technique for nanoparticle characterisation which has benefits in
being able to directly observe the particle morphology, but has limitations in that the number of
particles that can be sampled is relatively small. Like XRD, electron and neutron diffraction can
also be used to determine nanoparticle structure.[48–51] These techniques can provide comple-
mentary information to XRD due to the variation in scattering cross-section with atomic number
(and isotope) being different from that for X-rays. Neutron diffraction can also yield insights
into magnetic ordering within the structure.[48,49] Selected area electron diffraction is frequently
performed in conjunction with transmission electron microscope (TEM) studies of nanoparticles.
For determining the nanoparticle size distribution, a number of methods can be used. While
this review will focus on small-angle scattering using X-rays, the same method can be applied
using neutrons, or visible or UV light. Small-angle neutron scattering can be used to exploit the
scattering contrast between different isotopes (usually hydrogen and deuterium) to highlight var-
ious structural components through contrast matching. This has been used to determine core and
shell sizes in core–shell nanoparticles,[52] ligand shell thickness,[53] and arrangement of ligand
molecules.[54] The technique requires relatively large sample sizes and long collection times.
232 B. Ingham

Static light scattering involves measuring the scattered intensity of a visible or UV laser versus
angle. Samples must be colloidal suspensions of concentrations that are dilute enough to allow
transmission of light and to avoid particle–particle interactions. The minimum particle size acces-
sible is determined by the diffraction limit for the wavelength of light used (Dmin ∼ λ/2, ∼ 300
nm for visible light). Dynamic light scattering also uses a visible light laser to obtain particle sizes
in the 1–1000 nm range. In this technique, intensity variations with time are measured to deter-
mine the diffusion behaviour of particles due to Brownian motion, which is related to their size
(hydrodynamic radius), including any surfactant layer that may be present. Like the other scat-
tering techniques mentioned here, it is an ensemble average technique, strongly affected by the
presence of large particles or aggregates. As for static light scattering, samples must be colloidal
suspensions of relatively low concentrations to avoid artefacts arising from multiple scattering.
Extended X-ray absorption fine structure (EXAFS) is another powerful structural charac-
terisation technique that has been applied to nanoparticles, made available in recent decades
through the development of synchrotron facilities and analysis software. It involves measuring
the absorption or fluorescence signal across an absorption edge of the element of interest. Above
the absorption edge, a photoelectron is ejected from the target element. The interaction of this
ejected photoelectron with the surrounding atomic structure is manifested in the oscillatory ‘fine
structure’ in the absorption signal extending to energies a few hundred eV above the edge. Mod-
els of the local atomic environment can be constructed and refined to fit the data. In contrast
to diffraction techniques, EXAFS does not require the material to have a long-range structural
order. It can yield information about atomic distances and average co-ordination numbers in the
structure, thus providing complementary information to XRD.[55–57] For small nanoparticles,
the EXAFS signal at longer distances can be severely damped due to the finite particle size,
since atoms on the surface of the particle have lower coordination numbers than those in the
core.[58,59] EXAFS can also yield insights into the coordination environment around a target
atom in bimetallic particles, distinguishing between alloys and segregated structures.[59,60]
In this review, we will focus on X-ray scattering techniques, namely XRD and small-angle X-
ray scattering (SAXS). The review is written in tutorial style and is aimed at readers with limited
experience in nanoparticle characterisation by X-ray scattering methods. We begin by discussing
experimental considerations, then move on to describe the information offered by each technique
and how it can be extracted. Numerous examples will be given of these techniques applied to
both ex situ and in situ studies of nanoparticle systems. We will also cover advanced techniques
including anomalous (or resonant) scattering and pair distribution function (PDF) analysis.

2. X-ray diffraction
As mentioned briefly in the Introduction, XRD provides information about the arrangement of
atoms within a crystalline material. For larger particles (diameter > 10 nm), bulk polycrystalline
analysis methods are generally appropriate, since these particles are typically small crystals of
a bulk structure. For very small nanoparticles (diameter < 10 nm), many of the bulk analysis
methods fail and other techniques are required, which we will discuss in due course.
When periodic structures of a given spacing are illuminated with light of a comparable wave-
length, diffraction arises through constructive interference. X-rays have wavelengths of the
order of 1 Å and diffract from planes of atoms in a crystal according to the Bragg equation,
2d sin θ = nλ, where d is the separation between atomic planes (the ‘d-spacing’), θ is half of
the diffraction angle, n is an integer and λ is the X-ray wavelength . This equation can also be
expressed in terms of the magnitude of the scattering vector, Q = 2π/d = (4π/λ) sin θ , which is
independent of the X-ray wavelength. Analysis of the position, width, and shape of the diffrac-
tion peaks can yield information about the crystalline phases present, the crystallite size, strain,
Crystallography Reviews 233

preferential orientation (texture), shape anisotropy, and concentration of stacking faults and other
crystal defects.[61,62]
A crystal is periodic in three dimensions and can be described by symmetry relations of the
atoms within the unit cell, the smallest ‘box’ in real space which can be translated in three
dimensions to create the entire structure. Different arrangements of atoms will result in different
patterns of XRD peaks being observed. The positions of the peaks are determined by the unit
cell dimensions and angles, while the intensities of the peaks are determined by the types and
relative positions of the atoms within the unit cell. For high-symmetry structures, some peaks
may be extinguished altogether. Generally, the first step in XRD analysis is phase identification:
determining what crystalline species are present, usually by performing a simple search-match
operation on the positions of the most intense peaks observed. There are several databases avail-
able of known crystalline substances, including those maintained by the International Centre
for Diffraction Data (ICDD),[63] the Cambridge Crystallographic Data Centre (CCDC),[64]
the Fachinformationszentrum (FIZ; Karlsruhe, Germany),[65] and the US National Institute of
Standards and Technology (NIST).[66]
Phase composition. When there is a mixture of crystalline phases present, peaks from all
phases will be observed with integrated peak areas being in proportion to the relative volume
of each phase. By comparing peaks from each of the various phases and knowing (or calcu-
lating) the structure factors, the relative volumes can be obtained from the ratio of integrated
intensities.[67] The integrated intensity E for the (hkl) reflection of phase A is given by [61]

|Fhkl,A |2
Ehkl,A ∝ C0 wA LPmhkl,A Texhkl,A a exp (−2M ), (1)
VA2

where C 0 is a constant (depending on the incident intensity and other factors); wA is the mass
of phase A; L is the Lorentz factor, L = 1/(sin θ sin 2θ ) [61]; P is the polarisation factor, given
by P = P0 + (1 − P0 ) cos2 2θ , where P0 is the extent of polarisation in the incident X-ray beam.
mhkl,A is the multiplicity factor (number of equivalent reflections), Texhkl,A is the texture factor
(degree of alignment along a particular crystallographic direction), and F hkl,A is the structure fac-
tor (scattering power of the atomic planes) for the (hkl) reflection of phase A. V A is the unit cell
volume of phase A and a is the illuminated sample area.[68] The Debye–Waller factor 2M can be
calculated from the root-mean-square displacement disorder amplitude σ as 2M = (Qσ )2 , where
Q is the magnitude of the scattering vector. Both static and dynamic (vibrational) displacements
will contribute to σ . For untextured samples, the Tex factor is constant (Tex = 1). For highly
polarised incident radiation (e.g. from a synchrotron source), P ≈ 1 and L reduces to L˜(1/Q2 ).
By rearranging the equation for wA , the weight percent of each phase can be determined.
Crystallite size. Size is one of the properties of most interest to control in nanoparticle systems.
Crystallite size is manifested in the broadening of XRD peaks due to the finite extent of the
long-range order of the atoms within the particle and is commonly reported by applying the
Scherrer equation to a peak or selection of peaks.[61,62] This equation is given in terms of the
diffracting angle 2θ by D = bλ/FWHM(2θ) cos θ (where θ is given in radians and λ is the X-ray
wavelength), or in terms of Q by D = b2π/FWHM(Q). In both cases, FWHM is the full-width at
half-maximum of the peak and b is a constant having a value between 0.89 and 0.94 depending
on the function used to fit the peak. The Scherrer equation is generally applicable to crystallite
sizes between 5 and ∼ 100 nm; the lower limit being determined by the definition of crystallinity
[69–71] and the upper limit being determined by the instrument angular resolution. D is the
dimension of the crystallites assuming that they are cubic and monodisperse in size. For highly
anisotropic nanoparticles, different diffraction peaks may have very different widths and so the
interpretation of D must be treated carefully. For polydisperse systems, D is the volume-averaged
crystallite size. In many studies, D is reported as the nanoparticle size; however, this may not
234 B. Ingham

necessarily be the case. The crystallite size may be smaller than the nanoparticle size if the
particle is polycrystalline (consisting of multiple grains), or if there is a significant rearrangement
of the atoms at the surface of the particle resulting in the formation of an amorphous layer.
Polydispersity in the nanoparticle size may result in peak shifts as well as peak broaden-
ing, as surface tension causes bond contraction within the nanoparticle [72–75 and references
therein]. This effect is more pronounced for smaller particles, and means that the diffraction
peaks observed may be broadened and shifted as a result.[76]
Strain. The limited size of nanoparticles can give rise to strain, both homogeneous and inho-
mogeneous, due to surface tension and bond contraction in the vicinity of the surface. Strain, like
finite crystallite size, is manifested in an XRD pattern by the broadening of the diffraction peaks.
If the particle size is large enough, that is, crystallite size broadening is sufficiently small, these
two effects can be deconvoluted since they have different dependences on the scattering angle.
By plotting the peak width as a function of Q, a linear relation is expected according to
b2π cεrms Q
FWHM(Q) = + , (2)
D 2π
where b is the Scherrer constant and D is the crystallite size, as above; εrms is the root-mean-
square strain and c is a factor of order unity that depends on the nature of the lattice distortions
and the strain model used. A plot of FWHM(Q) versus Q then will yield an intercept of (2πb/D)
and a slope of (cεrms /2π). This method was first proposed by Williamson and Hall [77] and has
been applied to various nanoparticle systems.[78–81]
The Williamson–Hall analysis method described earlier gives values for the crystallite size
and strain that are averaged over all crystallographic directions. In some cases, this may not be
appropriate. A more rigorous method involving Fourier transforms of multiple orders of diffrac-
tion peaks (e.g. 111, 222, 333, . . . ) has been developed by Warren and Averbach [82,83] to yield
values for the crystallite size and strain for different crystallographic directions. However, the
method is not widely used: the Fourier transformation is highly dependent on instrument broad-
ening corrections and accurate background subtraction, meaning that in cases where diffraction
peaks overlap (as is common for nanoparticles), erroneous results may be obtained. A wide angu-
lar range is needed to capture the higher order reflections. For cubic crystal structures, the third
order of a reflection always overlaps with another (e.g. 100, 200, 300/221), limiting the analysis
to two reflection orders only.
Although the Williamson–Hall and Warren–Averbach analyses are often applied to nanoparti-
cle systems, care must be taken especially when the nanoparticle size is small and/or anisotropic
effects (strain and defects) are significant. These result in deviations from the expected linear
behaviour [84] and changes to the peak shapes [85] which may yield erroneous values for the
strain if misapplied. Erroneous size and strain values will also be obtained for small particles if
their size polydispersity is too large.[86]
Shape anisotropy. For sufficiently large nanoparticles where crystallite size broadening applies
according to the Scherrer equation, particle shape anisotropy can be manifested as peak broad-
ening that is dependent on the hkl indices of the diffraction peaks. For example, nanorods having
enhanced growth along the [001] direction will have sharper (00l) and broader (hk0) reflections.
These reflection series can be analysed separately as above to obtain the crystallite size in each
particular direction. Alternatively, two crystallite sizes and a high degree of crystallographic
texture can be incorporated as independent parameters in full-pattern analysis fitting.[87] More
complex models describing the crystallite shape can also be derived.[88]
Stacking faults. Crystal faults occur when the regular stacking of atomic planes in a crystallite
is interrupted (deformation faults) or reversed (twin faults). They are reasonably common in
metallic nanoparticles due to surface tension. Stacking faults are manifested in XRD patterns
as both hkl-dependent peak broadening accompanied by shifts in the peak position (deformation
Crystallography Reviews 235

faults only) and/or peak asymmetry (twin faults only). Warren [61] has derived exact solutions for
face-centred cubic (fcc), hexagonally close-packed (hcp), and body-centred cubic (bcc) metals.
Once again, the peak broadening due to faults is in addition to the broadening due to the finite
crystallite size.[89,90]
Preferential orientation. Preferential orientation, also called crystallographic texture, can
be observed in films of aligned particles. This alignment may be caused by faceting,[91–
93] shape anisotropy,[94,95] growth method,[96–100] application of magnetic fields,[101] or
epitaxy.[102–104] Nanoparticles may display epitaxial (oriented in-plane in two directions) or
fibre texture (oriented out of plane in one direction). In these cases, the scattering is not isotropic.
The direction and extent of the preferential orientation can be determined by measuring rocking
curves and/or off-axis scans.
For very small nanoparticles ( < 5–8 nm), these bulk analysis methods often fail.[69–71] For
these cases, it is more appropriate to calculate the scattering pattern using the Debye method.
[105,106] The Debye equation that describes the scattering can be written as [71,107–109]
 sin(Qrmn )
IN (Q) ∝ fm (Q)fn (Q) , (3)
n,m
Qrmn

where I N (Q) is the scattered beam per unit solid angle in the direction of the scattering vector
Q, f i (Q) is the atomic scattering factor of atom i, and rmn is the distance between atoms m and
n. Equation (3) can be further simplified for monometallic systems, since f (Q) is the same for
all atoms. The Debye method has been used to study the structures of very small nanoparticles
(diameter < 5 nm).[70,71,109–114] In this size regime, the nanoparticle structure is affected by
surface tension, resulting in size-dependent polymorphism, strain, surface reconstruction, and
in some cases partial amorphisation. As a result, the diffraction peaks are broadened and may
be shifted in position. The method is particularly useful for calculating the scattering patterns of
nanoparticles having non-bulk structures such as icosahedra and decahedra, since it requires only
a list of Cartesian coordinates of the atomic positions and does not rely on crystalline symmetry.
XRD can be used as a characterisation tool to measure the final synthesis product, or to monitor
the synthesis or manipulation of nanoparticles in situ, including in real time. In the following
sections, we will address each of these. Firstly, however, we will discuss some experimental
considerations in performing XRD experiments on nanoparticle systems.

2.1. X-ray diffraction experimental considerations


In an X-ray scattering experiment, there are three main things to consider: the source, the sam-
ple, and the detector. In Section 1, we discussed different X-ray sources used for X-ray scattering
experiments, namely, laboratory sources and synchrotron sources. While for highly crystalline
materials the instrumental resolution of a laboratory source may be a limiting factor, for nanopar-
ticles this is generally not an issue since the particle size broadening is significantly larger than
the instrument broadening. This is illustrated in Figure 1. The benefit of using a synchrotron
source to study nanoparticle systems is the increased flux, allowing shorter measurement times,
which is particularly useful for in situ real-time studies.
Various detectors are used for XRD measurements. Broadly speaking, these may be classed as
0D (point), 1D (line or strip), or 2D (area) detectors. Point detectors include photodiodes, scin-
tillators, and photon counters and are scanned through angular space. They are most useful for
high-resolution studies, using narrow acceptance slits and/or analyser crystals. Since the broad-
ening arising from the small crystallite size of nanoparticles is significant, the angular acceptance
of the slit can be relaxed in order to increase the detected signal. Soller slits can also be used
to collimate the detected beam and allow larger beam spots to be used (e.g. grazing incidence
236 B. Ingham

Figure 1. XRD data collected for the same sample of Ru nanoparticles on a laboratory instrument
(X’Pert3 Powder Diffractometer) and a synchrotron instrument (powder diffraction beamline at the Aus-
tralian Synchrotron), demonstrating the sufficiency of modern laboratory sources for measuring diffraction
patterns from nanoparticles with adequate signal-to-noise. The laboratory measurement was recorded using
a position-sensitive line detector, 1 h total acquisition time, 2θ = 0.007°. The synchrotron data were
recorded using a Mythen strip detector, 15 min acquisition time, 2θ = 0.004°.

with the beam spilling over most of the sample). Detectors differ in terms of linearity (X-ray
flux/output signal), energy discrimination, noise, and cost. Scintillators and solid-state semicon-
ductor detectors are the most commonly used. Scintillators consist of a crystal that emits visible
light when it is struck by X-rays, connected to a photomultiplier tube that converts the visible
signal into an electrical signal. They are relatively inexpensive and do not require internal cool-
ing. The noise is generally < 1 count per second and can be linear up to 106 counts per second.
The main drawback of scintillation detectors is that the energy resolution is poor at around 50%.
Energy discrimination is important in cases where X-ray fluorescence processes may be occur-
ring. Semiconductor detectors are generally made of Si or Ge. X-rays create electron–hole pairs,
which migrate to the electrodes and are recorded as electrical pulses. The height of the pulse cor-
responds to the X-ray energy, meaning that these detectors can have good energy discrimination
(down to 2%, which is sufficient to resolve K α1 and K α2 ). They must be cooled either with liquid
nitrogen or, in the case of silicon drift detectors, an internal Peltier cooling module.
Position-sensitive detectors can be used for simultaneous data collection over a range of angles
in 2θ .[115] This allows data to be collected more rapidly, and although the angular resolution is
inferior to that of a point detector with certain slit settings, as mentioned earlier the diffraction
peak broadening from the small crystallite size of the sample is usually significant, meaning
that this is seldom an issue. Data collected using flat position-sensitive detectors need to be
corrected for positional and intensity non-linearity. Other 1D detectors are also available, such
as the Mythen detector, which spans up to 120° with a resolution of 0.004° in 2θ.[116,117]
Area detectors currently available fall into three main categories: charge coupled devices
(CCDs), image plates, and photon counters. CCDs were first developed in the 1970s. Modern
X-ray CCD detectors are scintillation devices that can offer small pixel sizes, down to 20 μm.
The maximum count rate is usually around 60,000 counts per pixel and depending on the size
of the detector the readout time can be a few seconds. The noise level of a CCD detector is
relatively high due to the amplification and reading of the electrical signal generated by the scin-
tillation and photomultiplication processes. Image plates operate using a plate made of an X-ray
storage phosphor material. When an X-ray photon strikes the plate, it excites an electron which
Crystallography Reviews 237

becomes trapped in a metastable state. The plate is read by illuminating each pixel with UV
or visible light. The trapped electrons absorb this light and return to the ground state, emitting
another optical photon of shorter wavelength (photostimulated luminescence). The intensity of
the luminescence is proportional to the X-ray dose on the pixel. While the noise level is much
lower than a CCD, the readout time is relatively long (30–80 s) and the resolution is limited by
the diffuse scattering of both the X-rays entering the material and the laser light used to read
the plate. Typical pixel sizes for image plates used for XRD are around 100–150 μm. In the last
decade or so, the Pilatus detector [118–123] has been developed by the Paul Scherrer Institute,
spanning a range of sizes up to 6 MP, having very fast readout times and virtually no noise. These
detectors have a dynamic range of up to 107 counts per second per pixel, and the pixel size is
170 μm.
The final consideration is the sample itself. Many nanoparticle samples are isotropic and
satisfy the requirements for powder diffraction. Nanoparticles may be prepared for XRD mea-
surements as thin films, either during their preparation (e.g. via sputtering,[124] inert gas
aggregation,[50,125] etc.) or, more commonly, drop-cast from a suspension onto an appropriate
low-background substrate. They may also be measured in solution, provided the concentration is
high enough to overcome the solution scattering.[126] If there is sufficient material, nanoparti-
cles can be measured as powders in capillaries. Usually, the capillary would be rotated during the
measurement to generate a powder average, especially when synchrotron X-rays are used, due
to their high collimation. Various sample mounts and in situ cells are illustrated in case studies
in the following sections.
While for randomly oriented samples a simple scan through the diffracting angle 2θ is all that
is required, samples showing preferential orientation may need to be mounted on an appropriate
goniometer to allow various reflections to be observed. Texture scans can be performed to obtain
information about the spread of orientations. These include θ − 2θ rocking scans and chi scans.
In a θ − 2θ rocking scan, the detector angle 2θ is set to where a diffraction peak is observed, and
the sample is rotated in the θ plane, that is, varying the incidence angle (Figure 2(a)). In a plot
of intensity versus θ , the width of the peak gives a measure of the spread of orientations and is
most evident for out-of-plane reflections. The angular range is limited to 0–2θ. In a chi scan, the
detector angle 2θ is again set to where a diffraction peak is observed, but this time the sample is
rotated in the plane of θ (Figure 2(b)). Area detectors can also be used to capture the entire pole
figure (a 2D representation of the intensity of a particular diffraction peak as a function of the
angular coordinates).[127]

Figure 2. Diagram showing rotation axes relative to the sample for (a) θ − 2θ rocking scan, (b) chi scan.
k and k are the incident and diffracted X-ray beam, respectively, and Q is the scattering vector.
238 B. Ingham

2.2. Ex situ XRD studies


Ex situ XRD is commonly used to measure the final product of a reaction, or samples obtained at
different times during synthesis or annealing by quenching. The primary benefit of performing
ex situ XRD measurements compared to in situ, real-time measurements, is the relaxed time lim-
itation. Obtaining high signal-to-noise ratios and/or high-resolution data can take several hours,
which in many cases is impossible during a reaction. The removal of time-dependence may mean
that a laboratory instrument may suffice. The only condition is that the sample must be stable dur-
ing the measurement, in terms of both the physical environment (temperature, atmosphere, etc.)
and the X-ray dose.
Another benefit of performing XRD measurements ex situ is that the presentation of the sample
to the X-ray beam is relatively straightforward. Nanoparticle powder samples may be mounted in
capillaries, or nanoparticle suspensions may be deposited on a low-background substrate and the
solvent allowed to evaporate. Controlling the evaporation rate of the solvent can result in highly
ordered nanoparticle arrays forming with interesting preferential orientation effects.[91,128,129]
The vast majority of XRD crystallographic texture studies are performed ex situ.
In the following sections, selected pieces of work are described that demonstrate the various
analyses that can be performed and information that can be obtained from nanoparticle sam-
ples using ex situ XRD. They are divided into sections covering metallic nanoparticles, oxide
nanoparticles, and other systems including non-oxide materials and quantum dot semiconductor
particles.

2.2.1. Ex situ XRD studies of metal nanoparticles


Metal nanoparticles are a unary chemical system, often possessing simple crystal structures, of
which the most common are fcc, bcc, and hcp. Despite this simplicity, the properties of metal
nanoparticles can vary significantly depending on their size and shape.
Faceted particles can form aligned arrays, evidenced by an enhancement of certain diffraction
peaks and suppression of others. Nanocubes can be formed from metals having a cubic crystal
structure, where {100} facets are preferred. When they are deposited on a substrate by controlling
the evaporation rate of the solvent, they may preferentially orient along this direction. A standard
symmetrical θ − 2θ scan will show an enhancement of the (200) and (400) diffraction peaks,
since the direction of the scattering vector Q is aligned with these crystal planes, normal to the
substrate surface. Such measurements have been recorded for Cu [130,131], Ag [132–134], Ni
[91], Pt,[128] and various Pt alloys.[135–137]
LaGrow et al. studied Ni nanocubes aligned on a silicon substrate by controlled evaporation of
solvent.[91] The sample exhibited fibre texture, evidenced by a difference in the intensity ratio
between diffraction peaks in in-plane and out-of-plane scans (Figure 3(a)): in both cases, the
(200) reflection was enhanced relative to all other peaks. Rocking scans were recorded in θ for
both the (200) and (111) peaks (Figure 3(b)), and in χ for both the (200) and (220) peaks (Figure
3(c)). The θ rocking scan for the (111) peak did not show any angular dependence, indicating that
at least part of the sample was isotropic, possibly a layer of particles on top of a well-aligned layer
in contact with the substrate. The (200) peak showed a prominent peak in both types of scans,
with a width at half-maximum in the θ rocking scan of around 8°. This evidence points to an
alignment of the nanocubes with the [100] direction normal to the substrate, which is consistent
with the {100} faceting of the particles. Since nickel has a fcc structure, the [100], [001], and
[010] directions are equivalent. If the [100] is normal to the substrate, one would expect to
observe an enhancement of the (220) reflection at 45°, and this was indeed observed in a χ scan
(Figure 3(c)). To explore the hypothesis that the sample consisted of a highly textured layer and
an isotropic layer, so-called L-scans (or Qz -scans) were measured on the (111) and (200) peaks
by recording the intensity as the incidence angle of the X-ray beam on the sample was changed
Crystallography Reviews 239

(a) (b)

(c) (d)

Figure 3. Various XRD scans of aligned Ni nanoparticles: (a) in-plane versus out-of-plane, (b) rock-
ing curve, (c) chi scans, and (d) L-scans with fits to a layer model. Reprinted (adapted) with permission
from LaGrow AP, Ingham B, Cheong S, Williams GVM, Dotzler C, Toney MF, Jefferson DA, Corbos EC,
Bishop PT, Cookson J, Tilley RD. Synthesis, alignment, and magnetic properties of monodisperse nickel
nanocubes. J Am Chem Soc. 2012;134:855–858 [91]. Copyright (2012) American Chemical Society.

in the vicinity of the critical angle. This has the effect of changing the X-ray penetration depth in
the sample.[138] Modelling the data yielded a textured layer having density 2.54 ± 0.04 g cm−3
and thickness 14 ± 6 nm (consistent with the size of the cubes obtained from TEM and SAXS of
12 nm), with an isotropic layer on top of density 1.67 ± 0.09 g cm−3 and thickness 21 ± 5 nm.
The difference in the densities can be explained by a more efficient packing of the nanoparticles
in the textured layer.
Bimetallic FePt nanoparticles 4 nm in size were studied by Anders et al. for potential applica-
tions in magnetic recording media.[139,140] These particles can be arranged in ordered arrays,
and if the magnetisation can be controlled, could result in a predicted potential storage density of
20 Tbits in−2 . Understanding the temperature stability and formation of the highly magnetocrys-
talline anisotropic L10 structure of FePt is therefore critical. Nanoparticle films were annealed at
various temperatures and XRD patterns recorded. The peak widths were observed to decrease
with increasing temperature, consistent with coalescence of the particles. At annealing tem-
peratures above 580°C, chemical order was observed as superstructure peaks emerged that are
normally forbidden in unordered fcc FePt. A L10 structure was obtained, having alternating
atomic layers of Fe and Pt. The unit cell became more tetragonal with increasing temperature.
Consequently, the magnetic coercivity of the material increased as the atomic order increased.
Hsieh et al. used XRD to study Ru@Pt core–shell nanoparticles.[141] Bulk Ru adopts an hcp
crystal structure, while Pt is fcc. These different structures were observed in the XRD patterns
of the monometallic particles (of diameters 8 and 3.2 nm for Ru and Pt, respectively). A series
of hcp Ru core-fcc Pt shell particles with different Pt shell thicknesses were synthesised. The
experimental XRD data showed an increasing fcc-like signal with increasing Pt shell thickness,
240 B. Ingham

and calculated XRD patterns (using analytic peak functions) showed that the best agreement to
the data was for a model that included partial alloying at the Ru–Pt interface.
A mixture of hcp and fcc structures was observed in monometallic Ni-branched particles by
LaGrow et al.[142] In this case, fcc Ni-truncated octahedral nanoparticles exhibited preferential
branched growth from the {111} facets. These branches showed a high concentration of stacking
faults in TEM and the emergence of diffraction peaks in X-ray and electron diffraction corre-
sponding to an hcp-like structure. Similar effects have also been observed in Au nanosheets
[143] and Au nanowires.[144]
Williamson–Hall and stacking fault analyses were performed on Cu/ZnO catalysts by Kan-
demir et al.[78] The materials consisted of aggregates of Cu and ZnO nanoparticles 10–13 nm in
size. The Cu nanoparticles exhibited numerous defects, including stacking faults and twin bound-
aries. This led to certain reflections showing anisotropic peak broadening. The apparent crystal
sizes, obtained from applying the Scherrer equation to each diffraction peak independently, var-
ied by up to a factor of 3. Due to the presence of stacking faults, the traditional Williamson–Hall
method needed to be modified to account for the fact that faults do not occur with equal proba-
bility in all lattice directions. The stacking fault probabilities were obtained by considering the
deviation in peak separation compared to the expected values, as derived by Warren.[61] Values
for the stacking fault density α of up to 3.7% and the twin fault density β of 7.4% were obtained.
The catalytic properties depend on the particle size, so accurate characterisation of the size when
stacking faults and other defects are present is critical for predicting the catalytic performance of
a particular material.
An example of the use of Debye function analysis is work performed by Zanchet et al. on
Au nanoparticles 2–4 nm in diameter.[71,110] Different patterns were obtained for fcc, icosahe-
dral, and decahedral particles (Figure 4).[71] The patterns were fitted to the data using variable
parameters that governed the construction of the modelled particle, and the ratio of mixtures
of different particle structures. In this way, it was determined that icosahedral structures were
favoured when the particle size was small (2 nm), while an fcc structure was favoured for larger
particles.[110] These findings are consistent with other experimental results [145] and molecular
dynamics simulations.[146] Debye modelling has also been applied to Pt and PtRu alloy nanopar-
ticles 1.5–5 nm in diameter by Baranova et al.,[109] by calculating the pattern and comparing
it to the data (no fitting was performed). Where the Scherrer equation failed to provide an ade-
quate interpretation of the data, Debye modelling enabled the researchers to obtain both size and
composition information. Cervellino et al. developed a modelling method incorporating struc-
ture, size, and strain to fit experimental data from non-crystallographic nanoparticle structures
by adding successive atomic layers to octahedral, icosahedral, and decahedral models.[147,148]

2.2.2. Ex situ XRD analysis on oxide systems


Many metal oxides can form a variety of phases, which in nanoparticles are often highly depen-
dent on the particle size. In nanoparticle systems, phases may exist which are only metastable in
bulk form (e.g. the high-pressure orthorhombic SnO2 phase formed by oxidising Sn nanoparti-
cles, observed by Chaudret et al. [149]; the metastable β-Bi2 O3 phase observed as an air-formed
oxide layer on Bi nanoparticles by Stevens et al. [150]). In addition, the physical properties
of nanoparticles depend strongly on their phase and size, for example, the magnetic proper-
ties of iron oxide nanoparticles. Lei et al. studied iron oxide particles 8–9 nm in size formed
by thermal plasma synthesis.[151] By controlling the oxygen flow rate, different iron oxide
phases were formed, including wüstite (FeO), hematite (α-Fe2 O3 ), and magnetite (Fe3 O4 ) or
maghemite (γ-Fe2 O3 ; the diffraction patterns could not be distinguished between the latter two
phases). Teng and Yang studied the effect of synthesis conditions on the formation of iron oxide
nanoparticles by thermal decomposition of a precursor solution.[152] γ-Fe2 O3 particles 3–25 nm
Crystallography Reviews 241

(a)

(b)

(c)

(d)

Figure 4. Calculated diffraction patterns of model nanoparticles using Debye modelling: (a) 2.8 nm fcc
particle having five shells, (b) 2.7 nm icosahedral particle having five shells, (c) 2.9 nm truncated decahe-
dron particle having five shells, and (d) ensemble average of 10 spherical fcc particles having stacking faults
along the [111] direction with a probability of 0.3. From Zanchet et al. [71]. Reproduced with permission
of the International Union of Crystallography. http://dx.doi.org/10.1107/S0021889800010888

in diameter were formed. The particles could be self-assembled into micron-sized superlattices
via rapid evaporation of solvent – a promising development for future fabrication of high-density
magnetic storage materials. Hou et al. synthesised FeO particles 14–100 nm in size in organic
solvents.[153] These particles had a Fe3 O4 coating which was more evident in the XRD pat-
terns for the smallest particle sizes. Upon heating in air, the FeO particles could be converted
to Fe3 O4 , γ-Fe2 O3 , or α-Fe2 O3 depending on the temperature and annealing time. Since the
242 B. Ingham

magnetic properties depend strongly on the phase (the FeO particles are weakly paramagnetic,
Fe3 O4 and γ-Fe2 O3 particles are superparamagnetic, while α-Fe2 O3 particles are weakly ferro-
magnetic), understanding how the phase is controlled through post-synthesis heating is critical
for tailoring particles for specific magnetic applications.
TiO2 has a number of polymorphs with different physical properties. Substoichiometric
TiO2−x has applications in photocatalysis and surface coatings through the introduction of Ti3+
ions into the structure, which introduces defects and changes the electronic band structure.
Teleki and Pratsinis studied TiO2−x nanoparticles synthesised using a diffusion flame aerosol
reactor.[154] A mixture of anatase and rutile particles were formed that had crystallite sizes
from 3 to 40 nm as determined from the XRD peak widths. The particle size depended on the
burner–nozzle distance. The difference in relative peak intensities compared to standard patterns
indicated the presence of lattice defects as a result of the oxygen substoichiometry.
Langford et al. used CeO2 particles having a log-normal size distribution with a mean of 3.5
nm and a standard deviation of 1.2 nm to demonstrate the effect of particle size distribution on
the shape of XRD peaks.[76] A significant polydispersity has the effect of lengthening the tails,
resulting in diffraction peaks being more Lorentzian in character than normal. This can be incor-
porated in whole-pattern fitting by parameterising the size distribution in the fit, which yields
physically relevant information on the size distribution, rather than simply applying analytical
functions (pseudo-Voigt) to fit the peaks. This analysis method, combined with Warren–Averbach
analysis, was also applied to a study of Fe3 O4 nanoparticles of 3–9 nm in size by Mahadevan
et al.[155] The value for the average particle size found from the Warren–Averbach analysis was
in good agreement with that obtained from dynamic light scattering. The shape of the column
length distribution derived from the Warren–Averbach analysis could be used to extract informa-
tion about the particle shape. The size results were in keeping with the observed magnetisation
behaviour of the particles.
The size and strain behaviour of CeO2 nanoparticles 4–60 nm in size were studied by Zou and
Huebner.[79] From Williamson–Hall plots, they found that the smallest nanoparticles (4 nm) did
not exhibit any internal strain (slope = 0), while larger particles (12 and 60 nm) showed a sig-
nificant strain, which was attributed to lattice mismatch at grain boundaries within the particles.
They also observed an expansion of the lattice parameter as the particle size decreased, which
was attributed to the presence of oxygen vacancies – an effect reported by others.[156,157] It
would have been preferable for the smallest particles in this study to be analysed using Debye
methods, given the difference in behaviour observed between the large and small particles.
Nanoparticles and nanostructures of the wide bandgap semiconductor ZnO have been widely
studied due to the effect of particle size on the electronic and optical properties.[158–163] ZnO
commonly has a hexagonal wurtzite crystal structure, which has the propensity for forming
anisotropic crystals, usually elongated along the c-axis to form nanorods. This shape anisotropy
is often manifested in XRD patterns of these materials by anisotropic line broadening of diffrac-
tion peaks corresponding to different crystal directions. For nanocrystals elongated along the
c-axis, the (00l) peaks will be narrower than (hk0) peaks. For larger nanoparticles, the crystallite
sizes along different directions can be obtained by single peak fitting [88,164] or by incorporating
two different crystal sizes with preferential orientation in a Rietveld refinement.[87] Zhang et al.
used XRD to follow the grain growth of ZnO nanoparticles formed by a sol–gel method.[165]
As-prepared ZnO particles 20–30 nm in size were heated at temperatures from 500°C to 1100°C.
From an Arrhenius plot of grain size versus inverse temperature, a grain growth activation energy
of 140 ± 6 kJ mol−1 was obtained. This value is less than that for the bulk; a finding consistent
with other nanoscale grain growth studies.[79,166–171]
Niederdraenk et al. included the crystallite shape as a variable together with lattice
parameters, oxygen atomic position, and stacking fault density in Debye modelling of ZnO
nanoparticles.[70,172] They found that if the crystallite shape was not accounted for in the
Crystallography Reviews 243

modelling (i.e. making the assumption that the crystal is a sphere), the fitted values for the
stacking fault densities were extremely high. However, if the crystallite shape was included, the
stacking fault density required to fit the data was negligible. Zhu et al. prepared ZnO nanoparti-
cles of isotropic shape ( ∼ 60 nm in size) deposited on carbon nanotubes and observed anisotropic
line broadening in the XRD patterns.[173] After annealing, the XRD peaks had a more uniform
peak width. These observations were attributed to the removal of stacking faults as a result of
the thermal treatment.
Texture studies of aligned Mn3 O4 nanoparticle arrays (for potential applications as magnetic
storage materials, catalysts, or batteries) were performed by Nakagawa et al.[174] The parti-
cles were 20 × 10 nm in size, elongated along the c-axis. Electron microscopy showed that
depending on whether the particles were assembled in solution or at the air–solution interface,
superstructures could be formed with either the a- or c-axis perpendicular to the substrate.
The preferred orientation of these assemblies was corroborated by XRD patterns, showing
enhancement of the diffraction peaks corresponding to those directions.

2.2.3. Ex situ XRD analysis on other systems


Nanoparticles of III–V or II–VI semiconductor materials, also known as quantum dots, have
been widely characterised using ex situ XRD. Small CdS particles of sizes between 0.7 and 10
nm exhibit a structural change from the cubic zincblende structure to the hexagonal wurtzite
structure as the particle size increases, with the crossover occurring at a size of around 4–5
nm.[175] This consequently has an effect on the band structure of the material, in addition to
particle size-related narrowing of the band gap. A similar effect was also observed in ball-milled
ZnS,[176] where the proportion of zincblende structure increased as the particle size decreased,
according to a Rietveld refinement. The wurtzite and zincblende structures can be thought of as
being related through differences in layer stacking. The stacking fault density can be included in
a Rietveld refinement as a fitted parameter.
The polytypism between zincblende and wurtzite can be exploited to produce nanostructures
of various shapes. Kim et al. synthesised GaP nanocrystals as both 8 nm zincblende spheres and
8 × 40 nm wurtzite rods.[177] Alivisatos et al. reported differences in the XRD patterns of CdSe
nanorods of 4–6 nm in diameter with varying aspect ratios.[178,179] By fitting peak widths, they
extracted relative crystal sizes in the a and c directions that were in good agreement with TEM
observations. In particular, the width of the (002) peak changed dramatically with rod length. The
nanorods exhibited regions of zincblende and wurtzite along the length of the rods, with a higher
proportion of zincblende observed for longer rods. Regions of zincblende were observed to act
as nucleation sites for the formation of branches having the wurtzite structure.[178] Spherical
CdSe nanoparticles were found to have higher stacking fault densities than nanorods.[179]
A number of ex situ XRD studies have been performed on various core–shell semiconductor
nanoparticles. These core–shell particles enable more precise tuning of the optical properties than
using either the core or shell material alone. In cases of CdSe/ZnS,[180,181] CdSe/CdS,[182] and
CdSe/ZnSe,[183] the wurtzite CdSe core peaks were observed to shift to higher 2θ angles as the
shell thickness was increased, indicating compressive strain. Baranov et al. [180] observed an
increase in the peak width, indicating that the presence of the shell may disrupt the crystallinity
of the core–shell interface. A crystalline ZnS shell was observed when the shell thickness was
greater than 2 monolayers (ML). For the CdSe/CdS nanoparticles studied by Peng et al.,[182]
the CdS shell was epitaxial and no change in peak width was observed as the shell thickness
increased. Lee et al. observed slight differences in strain along the a and c directions for CdSe
particles with ZnSe shells.[183] They also observed distinct diffraction peaks corresponding to
ZnSe for sufficiently thick shells.
244 B. Ingham

Figure 5. XRD patterns for 3.8 nm CdTe and CdTe@ZnSe nanoparticles with shell thicknesses of 2, 6,
or 9 monolayers, together with indexed bulk diffraction peaks (top and bottom). From Smith et al. [184].
Reprinted by permission from Macmillan Publishers Ltd: Nature Nanotechnology, copyright (2008).

Smith et al. studied the effect of various shell compositions and thicknesses on 1.8 and 6.2 nm
CdTe cores (Figure 5).[184] For ZnSe shells up to 9 ML thick, a systematic shift in the CdTe
zincblende peak positions was observed that was consistent with compressive strain. Additional
peaks corresponding to a wurtzite-like structure were observed for the thickest shells as a result of
a high stacking fault density (30–40%) in zincblende. These stacking faults were also responsible
for an elongation of the particle together with increased compressibility along the [111]zincblende
stacking direction.
As discussed in previous sections, Debye analysis is a useful and versatile technique for
modelling XRD patterns of very small nanoparticles. It has been applied to various II–VI semi-
conductor particles.[70,111,181,185,186] Niederdraenk et al. applied the method by way of
example to both ZnS and CdS nanoparticles approximately 2 nm in diameter.[70,185] XRD
data were fitted using a ‘single-particle modelling’ approach. The best fit to data from ZnS
particles used a zincblende crystal structure and an isotropic shape, 2.6 nm in diameter. In con-
trast, data from CdS nanoparticles could not be fitted satisfactorily using either zincblende or
wurtzite. When stacking faults were considered, acceptable fits were obtained. The particle size
obtained for the model including stacking faults was 2.1 nm, close to the value expected from
UV–visible absorption measurements, whereas the value from Rietveld refinement (using the
Scherrer equation) was 1.3 nm.[70] In an extended study, different particle shapes were also
included in the model.[185] This work highlights the failings of the Rietveld method for very
small crystallite sizes and demonstrates the usefulness of the single-particle modelling approach.
The Debye method allows the inclusion of particle shape and stacking faults, which can have
significant effects on the XRD patterns of very small particles.
XRD has also been used to study other non-oxide nanoparticles. Nickel phosphide nanopar-
ticles were studied by Moreau et al. using both XRD and EXAFS as poisoning-resistant
catalysts for fuel hydrotreatment.[187] Two different synthesis methods were used, which pro-
duced particles of similar sizes (9 and 15 nm). In the first case (where the precursor was
Crystallography Reviews 245

reacted at 230°C for 1 h), the XRD patterns revealed a single broad peak corresponding to
the most intense fcc Ni reflection. In the second case (where the precursor was reacted at
350°C for 2 h), a series of peaks corresponding to highly crystalline Ni2 P was observed.
EXAFS showed that both samples had a similar short-range structure to that expected for Ni2 P.
The Ni–Ni bond distances and coordination numbers in the first sample were consistent with
Ni2 P, not fcc Ni. The XRD data demonstrated the lack of long-range crystalline order in this
sample.
Singla et al. applied Williamson–Hall analysis to WC nanoparticles as part of a study investi-
gating the mechanical properties of this material at the nanoscale.[80] They obtained a crystallite
size of 37.1 nm and a uniform strain of 2.36 × 10−3 . The crystallite size was in good agree-
ment with the particle size obtained from TEM observations. Venkateswarlu et al. also used the
Williamson–Hall method to obtain the crystallite size and uniform strain in annealed hydrox-
yapatite nanoparticles 90 nm in size, such as those used in biomedical coatings.[81] In both
pieces of work, the effect of strain was significant; a simple application of the Scherrer equation
underestimated the true crystallite size.

2.3. In situ XRD studies


In situ XRD studies are mostly used to follow temporal changes to a crystalline material, or in
situations where to remove the sample from its environment would change its structure. These
studies may necessitate the construction of specialised environmental cells, and kinetic studies
often make use of synchrotron radiation in an effort to decrease the measurement time required.
In situ cells must be carefully designed to be compatible with both the X-ray geometry and the
environment under study.[188] Key design considerations include the following: (i) the chemical
and physical environment the cell needs to withstand, in terms of temperature, pressure, solution
composition, gas, etc.; (ii) whether a transmission or reflection geometry is required; (iii) the
X-ray path length required through the cell, which must be sufficiently large to accommodate
enough material to give a reasonable signal in the measurement time required and also minimise
diffusion-limited effects, while not being so large as to cause excessive absorption for transmis-
sion solution cells (e.g. 2 mm of water has a transmission of 60–70% for X-rays of energy 12–15
keV); (iv) whether flowing solution or gas is needed; (v) whether heating capability is required,
and over what temperature range; (vi) the angular acceptance of the cell windows, which must be
sufficiently large so as to allow detection of the necessary diffraction peaks for the X-ray energy
being used (generally not an issue for SAXS); and (vii) the window material, which should ide-
ally be as thin as possible to withstand the environment, and be made from a weakly diffracting
material. Additionally, for measurements of nanoparticles in solution, the particle concentration
must be sufficiently high to enable the nanoparticle diffraction peaks to be detected above the
solution scattering.
Kinetic processes that can be studied with in situ XRD include nanoparticle synthesis reac-
tions, chemical, or thermal annealing treatments to induce grain growth, phase transformations,
or melting, and the structural changes in nanoparticles under their operating conditions, for exam-
ple, as catalysts or sensors. In situ XRD studies are also used for detecting metastable phases
during thermal annealing, and to explore crystalline phases that may form at high pressures or
under non-ambient atmospheric conditions.
A series of selected examples are presented demonstrating the application of in situ XRD to
metal nanoparticle, oxide nanoparticle, and other nanoparticle systems.

2.3.1. In situ XRD during synthesis of metal nanoparticle systems


Cheong et al. used in situ synchrotron XRD to monitor the synthesis of Pt nanoparticles of
different shapes in a non-aqueous solution reaction [189] (Figure 6). A customised reaction
246 B. Ingham

vessel was filled with the Pt precursor solution in toluene, and sealed under a hydrogen atmo-
sphere. The sealed vessel was then mounted in a heating unit on the diffractometer and the
temperature raised to 70°C. XRD scans covering the first two reflections of Pt were recorded in
a continuous fashion. Two different concentrations of Pt precursor were used, which resulted in
different shaped nanoparticles. In both cases, Pt diffraction peaks were observed after around 75
min. The peak area was followed as a measure of the total volume of particles. For the parti-
cles formed from the low-concentration solution, the area increased in a monotonic fashion as a
function of time. The resulting particles were compact cubes and tetrahedra. However, the par-
ticles formed from the high-concentration solution formed irregular structures with high surface
areas and showed a plateau in the XRD peak area versus time plot. TEM images recorded of
samples formed under the same conditions and quenched at different times yielded insights into
the formation mechanism. These particles initially formed compact cubes, which after a certain
period of time started to be etched on the {100} faces, while continuing to grow from the corners
(the [111] direction). During this period, the total volume of material remained constant, result-
ing in the plateau observed in the XRD peak area versus time plot. After the etching process,
the nanoparticles continued to grow in a dendritic fashion until all of the precursor species were
consumed. Understanding how the high surface area particles are formed is key to controlling
the synthesis for catalytic applications.
Watt et al. used a similar reaction vessel to study Pd nanoparticle synthesis at room temperature
using in situ synchrotron XRD.[190] Spherical multiply twinned particles of around 6 nm in size
were formed when oleylamine was used as a surfactant, while highly branched particles around
85 nm in size formed when the surfactant was a 1:1 mixture of oleylamine and oleic acid. These
highly branched particles resulted due to preferential crystal growth along the [112] equivalent
directions. In the former case, in situ XRD showed a steady, nearly linear growth rate up to 45
min. However, for the latter system, three stages were observed corresponding to initial growth

(a) (b)

Figure 6. In situ XRD study of Pt nanoparticle synthesis. (a) Pt(111) XRD peak area versus time; (b)
TEM images of particles obtained from high-concentration reactions after (A, B) 75 min; (C, D) 120 min;
(E, F) 240 min; (G, H) 500 min. Insets in B, D, F, and H show the fast Fourier transform of the respective
images. Reprinted (adapted) with permission from Cheong S, Watt J, Ingham B, Toney MF, Tilley RD.
In situ and ex situ studies of platinum nanocrystals: Growth and evolution in solution. J Am Chem Soc.
2009;131:14590–14595 [189]. Copyright (2009) American Chemical Society.
Crystallography Reviews 247

of spherical particles, initial branching evidenced by an increase in the growth rate, and further
branching followed by settling of large particles out of the X-ray beam.
Cu–Ni particles of various composition ratios supported on silica for catalytic applications
were studied by Wu et al. using in situ XRD and X-ray absorption spectroscopy.[191] The par-
ticles were synthesised by an incipient wetness impregnation method, dried, and then heated
up to temperatures of 300–400°C under a reducing atmosphere. XRD and XAS measurements
were recorded during the heating step. Monometallic Cu particles were formed between 220°C
and 240°C after the Cu hydroxynitrate precursor transformed to CuO at around 200°C, which
was then reduced to Cu. Similarly, Ni hydroxynitrate decomposed to NiO at around 200–260°C,
which then reduced to Ni between 280°C and 360°C. Mixed composition systems showed the
formation of Cu particles which then appeared to catalyse the Ni reaction as the temperature was
increased. When the Ni content was high, an alloy was formed; however, when the Ni content
was low, a mixture of Cu and Ni–Cu particles were observed. The particle size was significantly
larger with higher Cu content (30 nm for monometallic Cu particles, c.f. 5–7 nm for Ni/Cu > 1).
Another example of in situ XRD being used to study the formation of bimetallic nanoparticles
is work on electrocatalytic Pt–Cu particles reported by Oezaslan et al.[192] The nanoparticles
were prepared by impregnating carbon-supported 2.4 nm Pt nanoparticles with a Cu precursor,
freeze-drying, then annealing. Ex situ XRD of the annealed material showed that three phases
were formed: a Fm3m disordered alloy, a Pm3m ordered intermetallic, and Fm3m Cu. In the
in situ XRD patterns recorded as a function of time while the temperature was held at 800°C
(Figure 7), the intensity of the Cu peak decreased, while the disordered alloy peak increased in
intensity and decreased in width (corresponding to an increase in the crystallite size). No peaks
corresponding to the desirable ordered PtCu3 phase were observed. However, under annealing
conditions where the cooling rate was controlled, superlattice peaks corresponding to ordered
PtCu3 were observed at temperatures below 600°C.
Mishra et al. reported a study of gold nanoparticles formed by co-sputtering of gold/silica thin
films, followed by ion irradiation.[193] XRD patterns collected in situ after various levels of

Figure 7. A series of XRD patterns recorded during PtCu alloy nanoparticle formation at 800°C over 9 h.
Arrows show the change in intensity with time, and numbers indicate the scan. Reprinted with permission
from Oezaslan M, Hasché F, Strasser P. In situ observation of bimetallic alloy nanoparticle formation and
growth using high-temperature XRD. Chem Mater. 2011;23:2159–2165 [192]. Copyright (2011) American
Chemical Society.
248 B. Ingham

ion fluence showed a decrease in the peak width and a slight peak shift to higher angles, corre-
sponding respectively to an increase in particle size from 4 to 9 nm and a decrease in the lattice
constant, the latter possibly as a result of reduced surface tension. In this case, the gold nanopar-
ticles grew as a result of the irradiation causing extreme local heating, resulting in melting and
resolidification. The method demonstrated good control over the nanoparticle size, which in the
case of gold is useful for tailoring its optical properties.
Peng et al. reported in situ XRD of silver nanoparticle synthesis by a hot injection method
(Figure 8).[194] They observed the formation of AgCl nanoparticles almost immediately ( < 1 s)

(a) (c)

(b) (d)

(e)

Figure 8. (a) 2D contour plot and (b) selected scans collected during the synthesis of Ag nanocubes. Posi-
tions of Ag and AgCl peaks are indicated by sticks in (a) and symbols in (b), as labelled. (c) Integrated peak
areas and (d) crystallite size obtained from particular reflections as indicated. (e) Schematic diagram of the
important processes labelled in (c, d): (I) nucleation and growth of separate AgCl and Ag particles; (II) Ost-
wald ripening of AgCl and Ag particles; (III) reduction of AgCl particles to Ag; and (IV) Rearrangement of
multiply twinned Ag particles to Ag nanocubes via Ostwald ripening. Reprinted (adapted) with permission
from Peng S, Okasinski JS, Almer JD, Ren Y, Wang L, Yang W, Sun Y. Real-time probing of the synthesis
of colloidal silver nanocubes with time-resolved high-energy synchrotron X-ray diffraction. J Phys Chem
C. 2012;116:11842–11847 [194]. Copyright (2012) American Chemical Society.
Crystallography Reviews 249

and Ag particles within 10 s. Both species coexisted for approximately 60 s before the AgCl
nanoparticles converted to Ag. This conversion was complete by 100 s. Analysis of the conver-
sion kinetics and the behaviour of the particle size with time implied that the surfaces of the
AgCl nanocubes acted as nucleation sites for Ag. Ostwald ripening – where smaller particles
are preferentially dissolved and the material absorbed onto larger particles – caused these nuclei
to increase in size and number, and once they reached a critical density, the AgCl nanocubes
self-catalytically converted into Ag.

2.3.2. In situ XRD during post-synthesis processing of metal nanoparticle systems


There are numerous studies where in situ XRD has been used to follow the kinetics of grain
growth as a result of thermal annealing. Sun et al. studied the thermal behaviour of 4 nm Ag
nanoparticles embedded in mesoporous silica versus deposited on a silica surface.[195] They
observed significant aggregation of the particles deposited on flat surfaces at 500°C, while for
the confined particles coarsening was only observed at temperatures above 600°C. The coales-
cence behaviour of small (1–5 nm), multiply twinned (non-fcc) gold nanoparticles was studied
using XRD by Vogel et al.[196] At temperatures around 400–420°C, the XRD pattern changed.
The Debye function analysis was used to obtain histograms of the size and proportion of octa-
hedra, decahedra, and fcc particles present. When the temperature was raised above 460°C, the
XRD peaks sharpened dramatically, indicating rapid grain growth. Ingham et al. also studied the
thermal coalescence and grain growth of 10 nm gold nanoparticles using both XRD and SAXS
as a function of time at fixed temperatures from 200°C to 350°C (Figure 9).[89] They found that
aggregation of the particles (as detected from SAXS) occurred on a much shorter timescale than
grain growth, as determined from XRD. The XRD results also showed evidence for a significant
reduction in the stacking and twin fault densities (from 0.02 to 0.005 and from 0.065 to 0.02,
respectively), determined from relative peak shifts.
XRD patterns collected during the annealing of 2–3 nm FePt nanoparticles for magnetic stor-
age applications up to 650°C by Nguyen et al. showed a number of structural changes.[197]
Above 360°C extra peaks were observed, indicating a phase transition from a fcc to face-centred

(a) (b)

(c)

Figure 9. (a) XRD time series for an Au nanoparticle sample heated at 250°C, showing sharpening and
shifting of the Au(111) and (200) peaks. (b) Deformation and twin fault densities and (c) crystallite size
obtained from the data in (a). Reprinted (adapted) with permission from Ingham B, Lim TH, Dotzler CJ,
Henning A, Toney MF, Tilley RD. How nanoparticles coalesce: an in situ study of Au nanoparticle aggre-
gation and grain growth. Chem Mater. 2011;23:3312–3317 [89]. Copyright (2011) American Chemical
Society.
250 B. Ingham

tetragonal structure. However, at temperatures well below this (180°C), the total cell volume
underwent a significant decrease, which may indicate that particle ordering began at lower tem-
peratures. The order parameter increased steadily once the temperature rose above 230°C, as did
the grain size as the surfactant was burned off. As noted earlier, control of the atomic ordering is
critical for tailoring the magnetic properties of the particles for the target application.
Champion et al. studied the sintering behaviour of 35 nm Cu nanoparticles for bottom-up
fabrication of electrical nanowires.[198] The nanoparticles were heated in a reducing atmosphere
to convert the 3.5-nm-thick Cu2 O surface layer to Cu. This was observed to begin at 90°C, with
sintering and grain growth observed from 150°C. Ingham et al. performed similar work on Cu
and Pd nanoparticles oxidised under ambient conditions,[199] and found that reduction of Cu2 O
and grain growth occurred above 200°C. No diffraction peaks were observed for Pd oxide phases,
although an oxide corresponding to one atomic monolayer was detected using X-ray absorption
spectroscopy. The effect of the oxide layer on the crystalline Pd core was evident in the XRD
patterns, with a significant reduction in strain observed after it was removed. Both pieces of
work show the importance of understanding how the air-formed oxide layers on nanoparticles
are reduced – a step that has implications for the subsequent annealing of the particles.
XRD has also been used to measure the thermal expansion properties of copper nanoparticles
on various supports and with various surface treatments by Champion et al.,[200] and of silver
nanoparticles under various atmospheric conditions by Hu et al.[201] The different behaviour
observed in the nanoparticles compared to the bulk material is attributed to the significant surface
energy of the nanoparticles. For silver nanoparticles 5 nm in size, the lattice expansion coefficient
was found to be roughly one-quarter of the bulk value.[201]
Insights into the melting behaviour of metal nanoparticles have also been obtained from in
situ XRD experiments. Theoretically, a reduction in the melting temperature with decreasing
nanoparticle size has been predicted.[202,203] Depero et al. measured Sn nanoclusters of 100 nm
in size encased in a SnO2 shell.[204] XRD reflections corresponding to metallic Sn disappeared
when the temperature was increased above 180°C (c.f. bulk melting temperature of Sn is 232°C).
Peters et al. studied the melting behaviour of 50 nm Pb crystallites in vacuum.[205] Using
XRD, they were able to observe surface melting of the crystallites from a reversible drop in inten-
sity as the temperature was oscillated just below the melting temperature. This was extended in a
later study [206] to distinguish between surface melting of large crystallites and complete melting
of small crystallites, again by oscillating the temperature. The sample consisted of particles rang-
ing from 6 to 24 nm in size, the largest of which exhibited a melting temperature suppression of
3% relative to the bulk melting temperature. The width of the diffraction peaks decreased accord-
ingly with each temperature oscillation. By successively increasing the temperature reached at
each oscillation (and thereby completely melting particles up to a certain size), they obtained
histograms of the size distribution of nanoparticles in the sample. The sample also displayed a
change in orientation of the particles: the (111) diffraction peak intensity increased by a factor
of two as the temperature increased, while the (200) peak intensity was largely unchanged. A
comparison of the diffraction pattern before and after melting at a temperature where the entire
sample is solid revealed that there was no significant coarsening or coalescence of the particles.
In contrast, Sheng et al. observed superheating of Pb nanocrystals embedded in an Al
matrix.[207] XRD peaks of Pb were observed up to 60 K above the bulk melting temperature.
Smaller particles were able to be heated to higher temperatures before melting, compared to
larger particles. The difference in behaviour between this work and that of Peters et al. [205,206]
can be attributed to the different surface energy of embedded versus supported particles.
In situ high-pressure XRD studies using synchrotron radiation were performed by Gu et al.
on Ag and Au nanoparticles.[208] A diamond anvil cell was used to apply pressure up to 30
GPa to the nanoparticle samples. The cell volume as a function of pressure (Figure 10) was
obtained by measuring the lattice parameters, and showed that the compressibility of the Ag
Crystallography Reviews 251

Figure 10. Unit cell volume as a function of applied pressure for Ag and Au nanoparticles compared
to the bulk materials. Errors in volumes are within the size of the symbols. The data for n-Au (50–100
nm), triangles, are taken from [209]. Solid lines are fits to the data using a third-order Birch–Murnaghan
equation of state (BM3 EOS) model. Dotted and dashed lines are BM3 EOS for bulk Ag [210] and Au [211],
respectively. Reprinted figure with permission from Gu QF, Krauss G, Steurer W, Gramm F, Cervellino A.
Unexpected high stiffness of Ag and Au nanoparticles. Phys Rev Lett. 2008;100:045502 [208]. Copyright
(2008) American Physical Society. http://dx.doi.org/10.1103/PhysRevLett.100.045502.

and Au nanoparticles (10 and 30 nm in diameter respectively) was significantly less than the
bulk materials; in other words, nanoparticles are stiffer than bulk material – a finding reported
by others.[212–215] This was hypothesised to be due to bond length compression due to the
presence of twin planes and stacking faults, and surface effects in small particles.
Suleiman and Pundt et al. [216–218] and Ingham et al. [219] studied the effect of hydro-
gen pressure on Pd nanoparticles less than 10 nm in diameter. Pd forms two distinct hydride
phases, with a so-called ‘miscibility gap’ between the maximum concentration of hydrogen
in the low-concentration (α) phase and the minimum concentration of hydrogen in the high-
concentration (β) phase. The increase in lattice parameter is proportional to the amount of
hydrogen in the structure,[220] offering the possibility for these nanoparticles to be used as
hydrogen gas sensors.[221,222] Both groups found a significant effect of nanoparticle size on
the lattice parameter expansion behaviour: with decreasing particle size, the α-phase could incor-
porate more hydrogen before converting to the β-phase, and the β-phase was stable at lower
hydrogen concentrations, that is, the miscibility gap decreased with decreasing particle size. A
comparison of the two bodies of work found that the nature of the surface also played a signifi-
cant role in the miscibility gap behaviour. Suleiman et al., who measured surfactant-coated and
polymer-encapsulated Pd nanoparticles 4.8 nm in size,[216] observed that the phase transition
occurred at higher hydrogen pressures and was much more pronounced for the surfactant-coated
particles. Ingham et al. measured uncoated particles formed by a physical deposition method 1.7–
6.1 nm in diameter,[219] and observed that while the behaviour of the β-phase with nanoparticle
size was the same for uncoated and coated particles, the behaviour of the α-phase was notice-
ably different, resulting in a closure of the miscibility gap at non-zero sizes for surfactant-coated
particles, while the gap remained open for all sizes of uncoated particles (see Figure 11).
Kaszkur et al. recently reported a thorough study using Debye modelling to investigate
chemisorption of various gases (including H2 , O2 , CO, and NO) on Pt nanoparticles less than
10 nm in size.[225] Adsorption of gas molecules on the nanoparticle surface changed the surface
energy, resulting in atomic reconstruction which was manifested in changes in peak position
252 B. Ingham

(a) (b)

Figure 11. (a) Lattice parameters obtained from fitting XRD peak positions for Pd clusters of three dif-
ferent sizes as a function of hydrogen pressure. (b) Lattice parameters corresponding to αmax and βmin for
the samples in (a), compared with nanograined Pd [223] and Pd nanoparticles with surfactant shells [224].
The solid and dotted trend lines are given as guides. Bulk αmax and βmin lattice parameter values are also
indicated. Reprinted (adapted) figure with permission from Ingham B, Toney MF, Hendy SC, Cox T, Fong
DD, Eastman JA, Fuoss PH, Stevens KJ, Lassesson A, Brown SA, Ryan MP. Particle size effect of hydro-
gen-induced lattice expansion of palladium nanoclusters. Phys Rev B. 2008;78:245408 [219]. Copyright
(2008) American Physical Society. http://dx.doi.org/10.1103/PhysRevB.78.245408

and intensity in the X-ray scattering patterns. The key point to note from this work is that the
shift in peak position (which would ordinarily be attributed to a change in lattice parameter)
was due only to a change in the arrangement of the surface atoms; the internal structure of the
nanoparticles was unchanged.

2.3.3. In situ XRD during synthesis and processing of oxide nanoparticle systems
As mentioned in Section 2.2.2, zinc oxide is a widely studied material for optoelectronic appli-
cations with the propensity to form a wide variety of crystal shapes. It can be formed using a
variety of techniques, ranging from physical deposition to solution methods. A number of in
situ XRD studies have been performed during synthesis of ZnO nanomaterials to understand the
growth mechanisms involved in forming the structure. Ingham et al. conducted an in situ XRD
study using synchrotron radiation to explore the origin of texture in electrochemically deposited
ZnO nanorod films (Figure 12).[96] 2D XRD patterns were collected during the deposition using
an image plate, giving a ∼ 2 min time resolution. The peak intensity followed a similar trend to
earlier work using X-ray absorption spectroscopy [226] in that there was an initial high electro-
chemical current and rapid linear growth, followed by a decline in the current and the growth rate.
This was attributed to the three-dimensional growth of the nanorods reducing to one-dimensional
growth once the rods grew large enough to come into contact with each other. The polar angu-
lar width of the peak showed little change from the earliest measurement time, indicating that
the initial nuclei are textured, even though their physical morphology from electron microscopy
appears globular.
Fall et al. also used synchrotron XRD to study the formation of ZnO by a hydrothermal
method.[227] They determined that the ZnO particles formed in solution as opposed to on the
substrate, and were isotropic in orientation. There was a limit to the lowest concentration of
Zn2+ precursor that could be used due to significant solution scattering which obscured the ZnO
diffraction peaks at low concentrations. For a concentration of 0.05 M, the growth of the ZnO
(101) peak was followed over a period of 300 min, during which time (a) the intensity rose to a
Crystallography Reviews 253

(a) (b)

(c)

Figure 12. (a) XRD scan collected at the end of a ZnO electrochemical deposition experiment showing
ZnO and Au reflections, together with the ZnO crystal geometry. The experimental parameters were 5 mM
Zn(NO3 )2 , 0.1 M CaCl2 electrolyte, deposition potential − 0.67 V (Ag/AgCl), solution temperature 65°C.
(b) ZnO (101) peak intensity (black closed squares) and width (red open circles), and (c) electrochemi-
cal current density, as a function of time during the deposition. Reprinted (adapted) with permission from
Ingham B, Illy BN, Toney MF, Howdyshell ML, Ryan MP. In situ synchrotron X-ray diffraction experi-
ments on electrochemically deposited ZnO nanostructures. J Phys Chem C. 2008;112:14863–14866 [96].
Copyright (2008) American Chemical Society.

plateau as the precursor was exhausted and (b) the peak width decreased, ultimately approach-
ing the resolution limit of the instrument geometry used. Bøjesen et al. also investigated the
hydrothermal synthesis of ZnO nanoparticles (15–30 nm in size) at different temperatures from
150°C to 250°C.[228] They observed the formation of isotropic ZnO following the appearance of
a Zn5 (OH)8 (NO3 )2 ·2H2 O intermediate. From Rietveld refinements, the apparent crystallite sizes
in different crystal directions were obtained as a function of time and showed anisotropic crystal
growth, with the c/a aspect ratio being higher for lower temperatures and reducing over time.
The effect of including twin faulting in the model was explored and was found to be significant,
with twin fault probabilities found to be ∼ 1% at the earliest times and dropping as the reaction
progressed.
Chen et al. explored the annealing behaviour of ZnO2 – an oxidising agent and precursor to
ZnO – as a function of temperature and pressure in an in situ XRD study (Figure 13).[229] ZnO2
particles with a crystallite size of 3.1 nm were obtained using a solution precipitation method.
The material has a cubic structure that is stable up to 36 GPa. At ambient pressures, ZnO2 was
stable up to 230°C, above which it rapidly converted into ZnO through the loss of oxygen.
Hydrothermal synthesis has also been used to form oxide nanoparticles. Yoon et al. used in situ
synchrotron XRD to study the formation of BaTiO3 nanoparticles 7.5 nm in size at relatively low
temperatures (50–80°C) for ferroelectric applications such as thermistors, sensors, and multi-
layer ceramic capacitors.[230] They did not observe any evidence for a TiO2 intermediate in
either XRD or Ti K-edge XANES. The BaTiO3 nanoparticles formed a tetragonal rather than
cubic structure (as normally observed for the bulk material), which may be due to size effects
or the inclusion of hydroxyl ions within the structure. Rehan et al. investigated the hydrothermal
synthesis of photocatalytic TiO2 nanoparticles using in situ energy-dispersive XRD.[231] The
titanium precursor gel was heated in acidic or alkaline conditions at either 210°C or 270°C for
up to 2 h. Under acidic conditions, rutile TiO2 was formed, while under alkaline conditions
anatase TiO2 was formed. In both cases, there was a delay before the onset of particle formation,
254 B. Ingham

Figure 13. In situ XRD patterns during ZnO2 nanoparticle synthesis at elevated temperatures (X-ray
wavelength 0.49334(5) Å). Reprinted with permission from Chen W, Lu YH, Wang M, Kroner L, Paul H,
Fecht HJ, Bednarcik J, Stahl K, Zhang ZL, Wiedwald U, Kaiser U, Ziemann P, Kikegawa T, Wu CD, Jiang
JZ. Synthesis, thermal stability and properties of ZnO2 nanoparticles. J Phys Chem C. 2009;113:1320–1324
[229]. Copyright (2009) American Chemical Society.

which was around 20 min for the formation of anatase in alkaline conditions and 35 min for the
formation of rutile in acidic conditions. The average final particle size was around 23–29 nm
in all cases, although TEM images showed a bimodal size distribution, which was attributed to
Ostwald ripening.
TiO2 anatase nanoparticles were synthesised using supercritical methods by Eltzholtz
et al.[232] In situ XRD was used to follow the development of diffraction peaks as a function of
time for the reactions performed at temperatures from 300°C to 500°C. The lattice parameters
and crystallite size were extracted as parameters from Rietveld refinements of the data collected
at each time step (0.5 s time resolution). Higher temperatures resulted in larger crystallites being
formed (up to 40 nm in size). The growth curves suggested a two-stage mechanism was respon-
sible for the formation of the particles. A significant reduction in compressive strain along the
c-axis as the particle size increased was also observed.
Catalytic NiO nanoparticles were synthesised by thermal decomposition and monitored using
ex situ and in situ XRD by Sietsma et al.[233] Mesoporous silica was impregnated with aqueous
nickel nitrate, which converted to Ni3 (NO3 )2 (OH)4 upon drying, and formed NiO upon sintering
at 450°C. Depending on the atmosphere used during the sintering step, particles with crystallite
sizes 4–25 nm were formed, with the smallest sizes observed for a 1% v/v NO/He atmosphere.
In situ XRD was conducted during thermal decomposition at 248°C under air and under 1% v/v
NO/He. The relative peak intensities of Ni3 (NO3 )2 (OH)4 and NiO were obtained as a function
of time. The reaction under air showed an incubation period of around 25 min, followed by rapid
conversion of Ni3 (NO3 )2 (OH)4 to NiO in around 5 min. In contrast, the reaction under 1% v/v
NO/He proceeded in a linear fashion, with the Ni3 (NO3 )2 (OH)4 to NiO conversion beginning
shortly after heating was initiated and being complete after 40 min.
Althues and Kaskel reported an in situ XRD study of the synthesis of ZrO2 particles in
reverse microemulsions for catalysis applications in low-temperature n-alkane isomerisation
reactions.[234] The crystallisation was followed as a function of temperature using in situ XRD.
Varying the water:surfactant ratio Rw had a profound effect on the size of the nanoparticles
formed. At low Rw values, crystalline materials were only formed at temperatures above 600°C.
These were 6.5 nm in size, increasing to 31 nm at 900°C. However, at high Rw values, crystalline
materials were observed at temperatures as low as 200°C, increasing in size to 24 nm at 600°C
and 300 nm at 900°C.
Crystallography Reviews 255

The thermal annealing behaviour of the cubic spinel ZnFe2 O4 , having magnetic and catalytic
properties, was investigated by Philip et al.[235] XRD patterns were recorded as ZnFe2 O4 par-
ticles 9.3 nm in size were heated to 1000°C. The crystallite size increased from 10 to 27 nm as
the temperature was increased from 600°C to 1000°C, after showing little change at tempera-
tures below 600°C. The lattice parameter showed a non-systematic variation with temperature
at 800°C. This was attributed to cation redistribution of Zn2+ and Fe3+ within the spinel struc-
ture, a finding supported by magnetic measurements which showed a dramatic increase in the
magnetisation after annealing at 1000°C.
Lai et al. studied the thermal annealing of magnetic Mn-doped γ-Fe2 O3 nanoparticles.[236] In
situ XRD patterns were recorded as a function of temperature. Usually, ferrimagnetic γ-Fe2 O3
transforms to antiferromagnetic α-Fe2 O3 upon heating to 500°C, which is undesirable for mag-
netic applications. However, with small amounts of Mn(III) ( < 10 wt%), the transition was
suppressed and γ-Fe2 O3 survived at temperatures exceeding 600°C with only partial transfor-
mation to the α phase. This demonstrated that the activation energy for the phase transformation
increased when Mn(III) was used as a dopant.

2.3.4. Synthesis and processing of other systems


There are a limited number of in situ XRD studies of non-metallic and non-oxide nanoparticle
systems in the literature. For most semiconducting quantum dot studies, the optical absorbance
peak is used as a reliable measure to follow the particle growth kinetics.[43–45] It is consider-
ably easier to measure the optical properties using laboratory instruments than to conduct an in
situ XRD experiment. Nevertheless, some studies investigating the formation of semiconductor
nanoparticles using in situ XRD have been performed.
Di Luccio et al. studied the nucleation of CdS nanoparticles in a polymer matrix by thermal
decomposition of a Cd thiolate precursor.[237] The precursor initially formed lamellar struc-
tures due to packing of the alkyl chains of the thiols. Chain lengths from 3 to 18 alkyl groups
were used. The lamellar structure was evident from a series of (00l) reflections. As the temper-
ature was increased above 160°C the sulphur atom was cleaved from the thiol, forming CdS
nanoparticles and disrupting the lamellar structure. In the XRD patterns (Figure 14), the (00l)
series of lamellar reflections decreased in intensity, while peaks corresponding to the CdS crys-
talline structure emerged. As in other studies,[70,175,185] these were best fitted using a mixed
zincblende/wurtzite structure. The average size of the nanoparticles was varied from 1.5 to 8 nm
by increasing the annealing temperature from 230°C to 300°C.
Zhang et al. studied the effect of water on the crystal structure of ZnS nanoparticles synthesised
in methanol.[238] As-synthesised particles gave an apparent crystallite size (from application
of the Scherrer equation) of 1.4 nm, around half the value of the particle size determined by
UV–visible absorption and TEM observations. Removing methanol by placing the particles in
a vacuum did not change their structure; however, adding 5% v/v water to methanol resulted
in a significant sharpening of the XRD peaks measured in situ, while the UV–visible absorp-
tion spectrum did not change. This indicated that the particles had become more crystalline,
while the particle size had not changed. The authors concluded that the change in surface energy
upon the binding of water was sufficient to cause a structurally disordered outer shell of the
particle to crystallise. This was corroborated with molecular dynamics simulations. XRD pat-
terns calculated from the simulated particle structures using the Debye formula agreed well with
experimental observations.
In a final example, Mainz et al. studied the annealing behaviour of Cu2 ZnSnS4 nanorods into
films using in situ energy-dispersive XRD.[239] Cu2 ZnSnS4 has a high absorption coefficient
and is a useful photovoltaic absorber material. The nanorods had a wurtzite crystal structure
and showed (hkl)-dependent peak broadening due to the anisotropic crystal shape (20 × 50 nm).
256 B. Ingham

(a) (b)

(c)

Figure 14. XRD patterns of Cd(C12 H25 )2 /topas (C12/TP) as a function of annealing temperature. (a) Low
annealing temperatures, where C12 precursor peaks (SLn) and transitory peaks corresponding to the mean
distance between growing CdS nanoparticles (Pn) are observed. TP is the topas peak. (b, c) Higher anneal-
ing temperatures showing (b) decomposition of the precursor in the low-q regime and (c) development
of crystalline CdS reflections in the high-q regime. Reprinted with permission from Di Luccio T, Laera
AM, Tapfer L, Kempter S, Kraus R, Nickel B. Controlled nucleation and growth of CdS nanoparticles in
a polymer matrix. J Phys Chem B. 2006;110:12603–12609 [237]. Copyright (2006) American Chemical
Society.

Diffraction patterns were recorded as the temperature was ramped. At around 380°C, there was a
rapid structural change from wurtzite to the desirable ordered kesterite phase. The kesterite grain
size increased rapidly from 20 to 160 nm. The final grain size was dependent on the heating rate.

2.4. Anomalous XRD


Anomalous X-ray diffraction (a-XRD) is a technique that has only relatively recently become
accessible through the construction of synchrotron radiation facilities.[62,240–242] It utilises
the tuneability of the synchrotron source to produce X-rays with energy close to an absorption
edge of an element of interest.
The atomic structure factor for XRD, Fhkl , in Equation (1) (Section 2) is a complex term
including q- and energy-dependent components. This can be written as Fhkl = f0 (q) + f  (E) +
if  (E), where f 0 (q) is the non-resonant component and f (E) and f (E) are resonant compo-
nents that arise from electronic transitions. At energies far from an absorption edge, the energy
dependence of Fhkl is minimal. However, in the vicinity of an absorption edge, it can decrease
significantly, as shown in Figure 15. This drop in |F hkl |2 can be used to determine scattering fea-
tures arising from particular elemental species, by collecting diffraction patterns at two nearby
energies where one is close to the cusp of |F hkl |2 (on the lower energy side, to avoid generating
fluorescence photons), and examining the difference.
Crystallography Reviews 257

Figure 15. Energy dependence of f 0 , f  , and f for Cu at q = 2.0 Å−1 , shown together with
|F hkl |2 , which displays a drop of around 30% at the K absorption edge (8979 eV). Values obtained
from [243,244].

Adora et al. studied the in situ deposition of platinum nanoparticles on carbon electrode sup-
ports, and used a-XRD to subtract the signal from the support, which had significant intensity
in the vicinity of the Pt peaks and was changing during the course of the deposition as a result
of changing porosity and texture.[245] a-XRD was compared to diffraction patterns obtained by
simply subtracting a scan taken of the support without any Pt particles, and the results were sim-
ilar. The a-XRD method would be preferred in cases where the change in signal from the support
is so great that a simple background subtraction is impossible.
a-XRD has proven to be a useful technique to study bimetallic nanoparticles.[246–248]
Jeng et al. studied PtRu nanoparticles at the Pt L3 -edge.[246] Diffraction patterns collected at
11.0 and 11.3 keV showed a difference in intensity but no difference in peak position, indi-
cating that the particles comprised a solid solution alloy. From the difference in intensity,
they determined that the Pt content was 63 ± 10%, which was in agreement with Vegard’s
law applied to the peak position shift relative to pure Pt and Ru. Strasser et al. studied PtCu
bimetallic core–shell catalytic particles [247] to determine the particle composition and lat-
tice parameter. The Cu content was determined by considering the drop in intensity at the Cu
K-edge, compared to the expected drop of |F hkl |2 . The results obtained from a-XRD agreed
well with those obtained by inductively coupled plasma optical emission spectroscopy measure-
ments. The lattice parameters were obtained from the peak positions and compared with those
expected from Vegard’s law. This showed a significant development of strain in the shell after
dealloying.
Connor et al. conducted an a-XRD study at multiple absorption edges to investigate Fe, Zn,
and Ga substitution in CuInS2 nanoparticles for control of the electronic band gap in solar cell
applications (Figure 16).[249] CuInS2 has a wurtzite crystal structure. The dopant species were
introduced up to 20%. Fe was found to substitute within the structure, and at higher doping levels
some CuFeS2 chalcopyrite phase was observed. Fe was found to be present to similar levels in
both wurtzite and chalcopyrite, despite wurtzite CuFeS2 being unstable. In TEM images, biphasic
nanoparticles were observed. Zn could be included in the wurtzite structure, as evidenced by a
drop in peak intensity at the Zn K-edge. a-XRD patterns of Ga-substituted CuInS2 , however, did
not show any anomalous effect at the Ga K-edge, indicating that although Ga was associated
with the nanoparticles (an overall X-ray absorption effect was observed), it did not substitute for
Cu or In at lattice sites.
258 B. Ingham

Figure 16. a-XRD effect at the Fe, Zn, and Ga edges for 10% elemental substitution in CuInS2 nanopar-
ticles shows that Fe and Zn are incorporated within the CuInS2 structure, but Ga is not. Reprinted with
permission from Connor ST, Weil BD, Misra S, Cui Y, Toney MF. Behaviours of Fe, Zn, and Ga substi-
tution in CuInS2 nanoparticles probed with anomalous X-ray diffraction. Chem Mater. 2013;25:320–325
[249]. Copyright (2013) American Chemical Society.

2.5. Pair distribution function analysis


PDF analysis,[250–253] also called total scattering analysis, is a relatively new technique made
more accessible in recent times by the increased availability of high-energy neutron and syn-
chrotron X-ray sources. It involves collecting a scattering pattern to high Q (usually Qmax ≥ 20
Å−1 ), followed by a Fourier transform procedure to yield the PDF G(r). Unlike traditional Bragg
analyses (e.g. Rietveld methods), PDF analysis utilises both the Bragg peaks and the diffuse scat-
tering components that arise from local disorder in the structure due to stacking faults, atomic
displacements, finite particle size, etc. Various analysis techniques are used to construct appro-
priate models to fit the data. PDF analysis can be used to study glasses and amorphous materials,
nanocrystalline systems, and materials with local chemical or structural disorder.[250]
A number of errors can arise in the collection of PDF data, which can produce spurious results.
(1) Termination errors arise from processing data that do not extend to a high enough Qmax ,
and are manifested both as ripples in the data at unphysically low radii and as satellite ripples
near the peaks.[254] To mitigate this, damping methods can be used (with caution), but the best
approach is to collect data with as high Q-resolution as possible to as high a value of Qmax as
possible. (2) Statistical errors arise due to increased noise at high Q. Choosing an appropriate
Qmax must therefore balance this noise with the termination errors produced. (3) Insufficient Q-
resolution affects rmax , the upper bound of the range over which oscillations are observed in
the PDF pattern. Ideally, the Q-resolution should be as high as possible, although this usually
involves a sacrifice in intensity.[251] (4) Systematic errors can arise from incorrect absorption
correction, background subtraction, and intensity normalisation, meaning that care must be taken
with these procedures.[254–256] Good signal-to-noise data are required for both the sample and
background (i.e. sample holder).[257]
X-ray PDF measurements require high-energy X-rays to achieve a high Qmax (a useful guide
is E in keV gives roughly Qmax in Å−1 ). This requirement does not necessarily limit one to
using synchrotron facilities; laboratory PDF measurements have been reported in the literature
using Mo Kα (E = 17.5 keV) and Ag Kα (E = 22.2 keV).[258] Obtaining PDF data with suffi-
cient signal-to-noise, Q-range and Q-resolution may take days using a laboratory source, even
for highly crystalline materials, although modern optics are becoming available for dedicated
laboratory PDF instruments. For weakly scattering samples, including many nanoparticle sys-
tems, this limitation may preclude the use of laboratory X-ray sources to measure PDF data and
a synchrotron source is needed. Many synchrotron facilities offer XRD beamlines suited to PDF
Crystallography Reviews 259

analysis through the combination of accessible energies and detectors available. PDF data can be
collected using a point detector, a strip detector,[257] or a 2D detector which may be an image
plate or CCD (each of which have pros and cons in terms of dynamic range, readout time, and
dark noise).[188,252] Recently, amorphous Si detectors have become available which can allow
rapid PDF data collection with measurements approaching real time.[259] PDF analysis can also
be performed using neutron scattering, which has advantages in measuring light elements due to
isotopic contrast,[251] but this is beyond the scope of this review. For nanoparticles, the Q-range
may include some of the small-angle scattering arising from the finite particle size (for diameters
less than around 10 nm). This can be incorporated into the model,[260] although such analysis
methods are still in their infancy. In an advanced application of the PDF technique, including
anomalous dispersion effects by collecting data using X-rays of energies close to an absorption
edge allows one to separate the specific contributions from a particular element of interest.[261]
This method cannot be applied to all systems, however, as there are limitations on applicable ele-
ments due to the minimum energy requirements to achieve a sufficiently high Qmax (for example,
the K-edge of first row transition metals is too low in energy to probe using this technique). In
addition, the signal-to-noise associated with the anomalous component is low. Despite this, the
technique has been successfully applied to studying different atomic geometries of bimetallic
nanoparticles, distinguishing between alloy and core–shell structures.[261]
It is preferable to use isotropic scattering data for PDF analysis, that is, obtained from powder
samples. These are usually measured as powders mounted in capillaries. The internal diameter
determines the Q-space resolution, and there may need to be a trade-off between the amount of
material required for sufficiently good signal and the required Q-resolution. Usually, the capillary
is rotated during the measurement to provide a better powder average of the sample, especially
when synchrotron X-rays are used, due to their high collimation. As for other X-ray scattering
techniques, the observed pattern is volume averaged over all particles present. To obtain mean-
ingful PDF information from nanoparticle samples, the particles should ideally be monodisperse
and have the same internal structure.
Qualitative pair-distance information can be read directly from the PDF pattern. Peak posi-
tions, especially for the first few nearest neighbours, correspond directly to interatomic distances
and are generally well defined even for poorly crystalline or amorphous materials. When the
pattern is correctly normalised, the peak amplitude corresponds to the coordination number. The
width of the peaks in the PDF pattern yields insights into both static and dynamic disorder (which
can be distinguished by examining the dependence of the peak width on the interatomic distance
r, and/or by measuring patterns as a function of temperature). The major downsides are that (a)
directional information is lost and (b) identifying the species of a neighbour in a multi-element
system is not always straightforward.
Quantitative analysis can be performed using various methods, which can generally be classed
as using either ‘small box’ or ‘big box’ models. In both cases, the PDF pattern for a structural
model is calculated and compared to the data. ‘Small-box’ modelling uses the smallest possi-
ble unit cell to describe the crystal structure – similar to the Rietveld method, with the same
parameters being refined (cell axes, angles, atomic positions, occupancies, and thermal param-
eters). The difference is that the calculated pattern is fitted to different regions of the r-space
PDF, highlighting differences in the average structure that may occur at short distances com-
pared to longer range order, where effects may average out. This method is typically used on
highly crystalline systems, such as to study chemical ordering [262] and local strain.[263] Many
nanoparticle systems can be modelled as a finite object cut out of a bulk atomic structure, by cal-
culating the bulk structure using a small-box model and then applying a shape function averaged
over all directions, which has the effect of applying a damping envelope to the data.[126,264,265]
However, in cases where there is significant disorder or structural rearrangement of the atoms
on the nanoparticle surface, such methods may fail or yield a size that is less than the true
260 B. Ingham

nanoparticle size.[126,266,267] In these cases, ‘big box’ modelling may be more appropriate.
This involves creating a model with a large number of atoms to simulate the entire structure,
such as using Monte Carlo methods.[266] Constraints are applied and parameters varied to obtain
solutions matching the experimental data. Due to the large number of parameters involved, it is
impossible to obtain a unique solution. Another method, as implemented in the program suite
DISCUS,[268,269] constructs an entire nanoparticle based on a few parameters (such as dis-
tances between atomic planes, stacking arrangements of atomic layers – including faults, and
limits on the size of the particle in three dimensions). This method allows for multi-element
systems to be modelled and can allow the creation of quite complex models, for example, var-
ious arrangement of atoms in a bimetallic nanoparticle,[270] or including a surfactant capping
layer.[264,270,271] Recently, even more complex algorithms have been proposed that build up a
structure one atom at a time. This method has been demonstrated to work for PDF data of simple
systems such as Lennard–Jones clusters and C60 [272,273] and show promise for solving larger
systems, depending on the constraints one is able to apply from a priori information.[274]
With regard to nanoparticle systems in particular, information that can be obtained from PDF
studies includes

(1) Phase. In Q-space, diffraction peaks of nanoparticles are broadened due to finite size effects.
Standard XRD methods such as Rietveld refinement may not be able to distinguish different
crystal structures, but since PDF is sensitive to the distances and coordinations of the nearest
neighbour environment, clear differences between structures can be observed (e.g. wurtzite
versus zincblende ZnSe,[275] CdSe,[112] GaN [267]; cubic versus tetragonal BaTiO3 ,[264]
(Ba,Sr)TiO3 [276]).
(2) Particle size. Due to a nanoparticle’s finite size, there can be no atomic pair correlations at
distances exceeding the particle diameter.[126,212,271] However, one needs to take care in
extracting a particle diameter directly from a PDF pattern, as there may be structural rear-
rangement at the surface, creating a disordered or amorphous layer. This will cause the PDF
oscillations to decay at shorter distances than the particle size.[126,266,267] Complemen-
tary techniques such as small-angle scattering or transmission electron microscopy (TEM)
can be used to better estimate the true particle size; noting the caveats in TEM imaging
of a small sampling volume, obtaining a number average rather than a volume average,
and the possibility of artefacts arising from surface layers that cause the particle size to be
underestimated.
(3) Particle shape. Various envelope functions can be derived for different shaped particles (e.g.
sheet, belt, rod, tube, sphere, hollow sphere) [265] and applied to a calculated PDF in the
same manner as a particle size envelope is applied.
(4) Defects. Models can be constructed to allow the inclusion of stacking faults (e.g. as reported
for nanoparticles of ZnSe [275] and CdSe [112]). Atomic displacement defects have been
detected for BaTiO3 [264] and (Ba,Sr)TiO3 [276] using PDF techniques.
(5) Non-bulk structures. Unique nanoparticle structures such as icosahedra and decahedra can be
difficult to distinguish using conventional XRD methods due to peak broadening,[110,196]
but PDF modelling shows that they can be easily distinguished, with differences arising from
around the third nearest neighbour onward.[277] Nanoparticle precursor structures of just a
few atoms can also be determined using PDF.[266,278,279]
(6) Homogeneous strain. Lattice contraction is often observed in nanoparticle systems and is
usually dependent on the particle size.[72–75] This can result in uniform changes in bond
length throughout the nanoparticle which can be easily detected using PDF.[112]
(7) Inhomogeneous strain. This is particularly seen in very small nanoparticles. Often the surface
bonds are contracted, and at smaller sizes the surface atoms make up a greater proportion of
the particle compared to the overall volume. This inhomogeneous strain is detected through
Crystallography Reviews 261

broadening of the peaks in the PDF at higher r.[112] The amplitude envelope may decay
faster than expected for the particle size, due to the atomic rearrangement of the surface.[212]
(8) Heterogeneous composition structures, for example, multi-metallic nanoparticle config-
urations. PDF can be used to distinguish between core–shell, alloyed, and segregated
structures.[261,267,280]
(9) In situ studies. The recent development of instruments for rapid acquisition PDF [259] offers
the possibility of performing real-time PDF studies of nanoparticle synthesis.[281]

(a)

(b)

Figure 17. (a) XRD data and (b) corresponding Fourier transformed PDF data for CdSe samples:
bulk (black line), 3.5 nm particles (CdSeIII, green line), 2.7 nm particles (CdSeII, red line), and 2.0
nm particles (CdSeI, blue line). Reprinted figure with permission from Masadeh AS, Božin ES, Far-
row CL, Paglia G, Juhás P, Billinge SJL, Karkamkar A, Kanatzidis MG. Quantitative size-dependent
structure and strain determination of CdSe nanoparticles using atomic pair distribution function analy-
sis. Phys Rev B. 2007;76:115413 [112]. Copyright (2007) American Physical Society. http://dx.doi.org/
10.1103/PhysRevB.76.115413.
262 B. Ingham

As mentioned in the Introduction, EXAFS is another technique that gives atomic-scale, short-
range structural information. Like PDF, it can be applied to amorphous or crystalline materials,
including nanoparticles. The main difference in the results obtained is that EXAFS is an element-
specific technique, while PDF arises from the scattering of all atoms present. Therefore while
EXAFS gives unequivocal information about the local structure around a particular atom, the
elemental species that can be studied are limited generally to Z ≥ 22 (Ti) due to the energy of
the absorption edge. PDF patterns have contributions from all the atoms in the sample and can
be easily applied to low-Z materials. EXAFS is also well established as a technique, while PDF
analysis is still largely in its infancy.

(a)

(b)

Figure 18. Fits to the PDF data in Figure 17 using (a) a wurtzite structural model and (b) a
zincblende structural model. Reprinted figure with permission from Masadeh AS, Božin ES, Farrow
CL, Paglia G, Juhás P, Billinge SJL, Karkamkar A, Kanatzidis MG. Quantitative size-dependent struc-
ture and strain determination of CdSe nanoparticles using atomic pair distribution function analysis.
Phys Rev B. 2007;76:115413 [112]. Copyright (2007) American Physical Society. http://dx.doi.org/
10.1103/PhysRevB.76.115413.
Crystallography Reviews 263

2.5.1. Examples
Masadeh et al. performed an X-ray PDF study of CdSe nanoparticles with sizes ranging from 2 to
4 nm.[112] The PDF patterns were compared with that of the bulk material. CdSe is a II–VI semi-
conductor, a family of materials that usually adopt either the zincblende or the wurtzite structure.
Both structures can be visualised as a series of stacked planes, described as ABABAB . . . for
wurtzite and ABCABC . . . for zincblende. In both structures, both elements are tetrahedrally
coordinated; however, from the second nearest neighbour onwards, the structures are different. In
conventional XRD, peak broadening due to small particle sizes can make it difficult to distinguish
between these structures, but the differences are clear in the PDF.[267] Figure 17 shows the S(Q)

(a)

(b)

Figure 19. Fits to the PDF data in Figure 17 using a wurtzite structural model with (a) no stacking faults
and (b) stacking faults allowed. Reprinted figure with permission from Masadeh AS, Božin ES, Farrow
CL, Paglia G, Juhás P, Billinge SJL, Karkamkar A, Kanatzidis MG. Quantitative size-dependent struc-
ture and strain determination of CdSe nanoparticles using atomic pair distribution function analysis. Phys
Rev B. 2007;76:115413 [112]. Copyright (2007) American Physical Society. http://dx.doi.org/10.1103/
PhysRevB.76.115413.
264 B. Ingham

data and Fourier transformed PDF G(r) for three nanoparticle samples of different sizes, along
with a bulk sample. The Q-space data (Figure 17(a)) shows that the diffraction peaks appear
in similar positions, with the peaks broadening as the particle size increases. The PDF patterns
(Figure 17(b)) all have reasonably sharp features. The features drop off in intensity at various
r values in keeping with the particle size. The PDF data were fitted using small-box modelling
methods with both zincblende and wurtzite structural models (Figure 18). The zincblende fit to
the data was poor (Figure 18(b)). While the wurtzite model provided a better fit than zincblende,
the best fit obtained was for a model allowing the inclusion of stacking faults (Figure 19). There
is little evidence in the data that the particles have an amorphous region of atomic rearrangement
at the surface. The peak position of the first nearest neighbour shifts consistently to shorter dis-
tances with decreasing particle size, compared to the bulk material. From the peak shift, a value
for the homogeneous strain in the particles of up to 0.5% can be extracted. The strain is homo-
geneous rather than heterogeneous: peaks in the PDF at longer distances are not broadened, as
would be expected for heterogeneous strain.
In another example, Welborn et al. reported a PDF study of small (1.7 nm diameter) Pt
nanoparticles either having capping ligands (amine or thiol) or encapsulated by dendrimers.[277]
A hybrid reverse Monte Carlo approach was used to analyse the data. The particles having cap-
ping ligands were found to be more disordered at their surface than the dendrimer-encapsulated
particles. This was manifested in the rapid decay of intensity at shorter distances for these sam-
ples (Figure 20). The reverse Monte Carlo method determines a particle structure by balancing
the energy obtained from DFT calculations against the goodness of fit to the experimental data.
These two aspects can be weighted to give an optimal solution. Three different particle structures
were considered: an icosahedron, a cuboctahedron, and a truncated octahedron. The truncated
octahedron gave the best results for all three samples, with clear differences emerging between
the icosahedron fits and the data in the range 6 < r < 10 Å (Figure 20). In the case of the

(a)

(b)

(c)

Figure 20. PDF data and fits to various models for (a) dendrimer-encapsulated, (b) amine-capped, and (c)
thiol-capped Pd nanoparticles 1.7 nm in diameter. Reprinted with permission from [277]. Copyright (2011)
AIP Publishing LLC.
Crystallography Reviews 265

capped particles, the number of capping ligands was also considered in the model and found
to be optimised at around 12–13.
In a final example, Petkov and Shastri performed PDF measurements near the Pt K-
edge (78.4 keV) to study the structure of bimetallic PtPd nanoparticles.[261] As discussed
in Section 2.4, the scattering factor f comprises energy-independent and dependent parts,
namely, f (q,E) = f 0(q) + f (E) + if (E). In the vicinity of an absorption edge, f (E) can deviate
significantly. The difference plot of G(r)E1 − G(r)E2 is the PDF consisting only of those atomic
pairs involving the element being probed. Figure 21 shows PDF data collected for four nanoparti-
cle systems: one pure Pt and three bimetallic PtPd samples having different atomic arrangements.
There are relatively few differences between these total PDF patterns. They were fitted using an
fcc small-box model multiplied with a particle shape function for spheres, allowing extraction
of the lattice constant and size of the coherent region. For the core–shell particles, the PDF peak
intensities decayed to zero at distances shorter than the particle size, and additional broadening of
the peaks was observed. Both these effects were due to a lack of coherence across the core–shell
interface due to lattice mismatch between Pt and Pd. The difference PDF plots at the Pt edge
(Figure 22) show clear differences between the Pd core/Pt shell and Pt core/Pd shell samples,
with oscillations persisting to longer distances in the former system, as expected. For the PtPd
alloy nanoparticle sample, the total and difference plots are virtually the same, as expected. The

Figure 21. PDF data (black symbols) and fits (red lines) for various Pt and PtPd nanoparticles as labelled.
The low-r region is shown in more detail in the inset (1 = pure Pt, 2 = Ptcore /Pdshell , 3 = PtPd alloy,
and 4 = Pdcore /Ptshell ). Reprinted figure with permission from Petkov V, Shastri SD. Element-specific
structure of materials with intrinsic disorder by high-energy resonant X-ray diffraction and differential
atomic pair-distribution functions: a study of PtPd nanosized catalysts. Phys Rev B. 2010;81:165428 [261].
Copyright (2010) American Physical Society. http://dx.doi.org/10.1103/PhysRevB.81.165428
266 B. Ingham

(a)

(b)

Figure 22. (a) Total PDFs for pure Pt (red) and pure Pd (blue). (b) Total (black symbols) and Pt dif-
ferential (red lines) PDF data collected near the Pt K-edge for various PtPd nanoparticles as labelled.
The inset in (b) shows a comparison to an fcc model (blue line), indicating that the oscillations in
the Ptcore /Pdshell experimental data at large r are related to the structure. Reprinted figure with per-
mission from Petkov V, Shastri SD. Element-specific structure of materials with intrinsic disorder by
high-energy resonant X-ray diffraction and differential atomic pair-distribution functions: A study of PtPd
nanosized catalysts. Phys Rev B. 2010; 81: 165428 [261]. Copyright (2010) American Physical Society.
http://dx.doi.org/10.1103/PhysRevB.81.165428

results show great promise for the use of this technique for other systems with absorption edges
of suitable energy.

3. Small-angle X-ray scattering


SAXS can be thought of in one sense as a corollary of XRD. In the latter, X-rays scatter from
atomic-sized elements (crystal planes), whereas in the former X-rays scatter from larger objects.
The larger the object, the smaller the angular deviation of the scattering. For X-ray wavelengths
of the order of 1 Å, the small-angle scattering region is for angles of 2θ less than around 5°. (Wide
angle X-ray scattering (WAXS) is a term that covers scattering observed in the range of 2θ larger
than this.) X-rays scatter from regions of electron density contrast. This may describe particles
in a matrix (solid, liquid, gas, or vacuum), porous materials, bicontinuous emulsions, etc.[282]
In this review, we will focus on the first category. An excellent resource on the technique is
Ref. [283].
Crystallography Reviews 267

The equation describing the SAXS intensity I as a function of the magnitude of the scattering
vector q is
 ∞
I(q) ∼ V 2 (ρ)2 [f (qr)]2 n(r)S(qr) dr, (4)
0

where V is the total volume of scatterers, ρ is the electron density contrast, and the integral is
performed over all radii.
f (qr) is the form factor: an analytical function describing the scattering from a single particle of
a particular shape. Ref. [284] gives a list of equations for different form factors, including core–
shell, spherical concentric shells, ‘raspberry’-type particles, ellipsoids, cubes and rectangular
parallelepipedons, faceted particles, circular and ellipsoidal cylinders, discs, fractal aggregates,
flexible polymers, and star-shaped polymers. f (qr) for other shapes can be derived. The form
factor determines the shape of the curve in the scattering data. An assessment of the power-law
behaviour of the SAXS data in different regions and the position of turnover points in q can give
immediate insights into the size and shape of the scattering objects. Figure 23 shows calculated
SAXS patterns for a dilute population of particles of different shapes. There are three regions of
interest. At low q, the scattering intensity tends to a constant value which is related to the total
volume of particles. In the vicinity of q = π /d, where d is the particle size (the Guinier region),
there is a turnover, above which the data then follow a power law with a slope related to the
nature of the particle surface (the Porod region). Globular particles, which can be approximated
by a spherical form factor, show a single turnover point to the Porod region, while anisotropic
shapes such as rods or discs exhibit two turnover points with power-law slopes of q−1 and q−2
respectively. Porod scattering from a smooth surface has q−4 behaviour; however, if the surface is
indistinct (i.e. fractal-like) the exponent may lie closer to − 3. For monodisperse, non-interacting
particles with simple shapes, oscillations at high q are observed (Figure 23).
n(r) is the size distribution of the scatterers. The most commonly used distribution func-
tions are the Gaussian, log normal, and Schulz distributions, although it has been shown that
when the size distribution is reasonably broad the exact form does not significantly alter the
result.[285,286] The effect of including size distribution broadening is to wash out the features
of the form factor scattering (most noticeably, the high-q oscillations). Since the form factor and
the size distribution are a product in the SAXS equation, it is not possible to obtain both inde-
pendently. Assumptions must be made about one or the other, for example, for a monodisperse

Figure 23. Calculated scattering for form factors of a sphere, rod, and disc. Regions of different
power-law slopes are indicated.
268 B. Ingham

sample, n(r) is a delta function; or the shape of the particles is assumed to be a sphere, core–shell,
disc, rod, etc. (e.g. based on TEM observations).
Finally, S(qr) is the structure factor, describing inter-particle interactions. For a dilute sys-
tem of non-interacting particles, S(qr) ≡ 1. In interacting systems, the hard sphere structure
factor is often used.[287] In this model, particles have a minimum separation between their cen-
tres determined by the so-called ‘hard sphere radius’. Other structure factor functions that have
been calculated analytically include the square well,[288] sticky hard sphere, [289] and screened
Coulomb potentials.[284] A system with inter-particle interactions will show the development of
a peak at the position q = π/d, where d is the spacing between the centres of electron density. For
highly ordered systems, more structure factor peaks can emerge, resembling Bragg reflections at
small 2θ angles.
There are several different analysis approaches that can be taken. The Guinier approximation
is commonly used to estimate the particle size in terms of a radius of gyration RG , which is
independent of shape. This is performed by plotting log I(q) versus q2 . The slope of this plot
as q → 0 is − − R2G /3. The radius of gyration is the mean square distance from the centre of
gravity, according to the electron
√ density. For a sphere, the radius of gyration is related to the
physical radius R by RG = 3/5R. This is demonstrated in Figure 24.
The ‘Unified model’ developed by Beaucage [290–292] allows one to fit SAXS data from
hierarchical structures using a series of ‘unified levels’, each consisting of a Guinier region and
a power law. This gives insights into the size of the scattering region (from the Guinier com-
ponent) and the nature of the surface (from the power-law exponent). Analytically, it is more
straightforward to fit to the data than is Equation (4); however, the method does have limita-
tions. If the length scales of the scatterers are not well separated, extracting the parameters can
be difficult.[293] There are also limitations on the applicability of the model to some systems.
The model was derived from the scattering of polymers, meaning that it is best suited to systems
where the exponent p is in the range 5/3 ≤ p ≤ 3, and does not work well for p > 3, as is often
the case for nanoparticles.[294] A new model was developed by Hammouda in 2010 [295] which
uses a similar empirical function and applies a constraint that the Porod slope must match the
Guinier region smoothly. It can allow for multiple Guinier regions, such as is seen for rods or
platelets.

Figure 24. A Guinier plot of the calculated data for a sphere in Figure 23, having a radius of 100
Å. The slope of the dotted line is − 2063.2, which results in an RG of 78.7 Å (or a sphere radius of

5/3 × 78.7 = 101.6 Å).
Crystallography Reviews 269

Fitting the data using Equation (4) can also be performed and yields physically meaningful
values for the particle size distribution, shape (e.g. aspect ratio), extent of interaction, and – if
the data are correctly normalised and the electron density contrast is supplied – the total volume,
or concentration, of particles. This is implemented in software packages such as SasView [296]
and Irena.[297,298] The ATSAS package [299–301] has primarily been developed for biological
systems, but some of its routines have been used to fit nanoparticle scattering data.
Like XRD, SAXS can be used to characterise the final synthesis product, or to monitor
nanoparticles in situ, including in real time. We will address each of these in the follow-
ing sections. We begin by discussing some experimental considerations in performing SAXS
experiments on nanoparticle systems.

3.1. SAXS experimental considerations


Most SAXS instruments, both laboratory and synchrotron, cover the region of reciprocal (q)
space corresponding to objects in the size range of 1–100 nm. This makes SAXS an ideal tech-
nique for studying nanoparticle size and packing interactions. The angles, as the name of the
technique suggests, are small ( < 5°) and therefore require an X-ray beam with low divergence
in order to separate the sample scattering from the direct beam. SAXS instruments may consist
of either a point source or a line source. Synchrotron SAXS beamlines are all point sources.
The benefit of using a line source in a laboratory instrument is the increased signal, as more of
the sample can be illuminated. However, a line source limits one to studying isotropic samples,
which is acceptable for the case of spherical or near-spherical nanoparticles. Data collected using
a line source also require slit smearing corrections to be made.
The vast majority of SAXS experiments are conducted in a transmission geometry, with sam-
ples as suspensions in capillaries, packed as powders or gels into an appropriate cell, or deposited
onto thin wafers. SAXS can also be conducted in a grazing incidence geometry (i.e. GISAXS) to
determine the size and shape of nanostructured films and surfaces. GISAXS measurements yield
two-dimensional scattering corresponding to the height and width of surface nanostructures, and
generally necessitate a synchrotron source as they require a high-intensity incident beam with
very low divergence.
Another factor arising due to the small angles involved is that it is necessary to protect the
detector from the direct beam by using a beamstop. This is typically a piece of highly X-ray-
absorbing material such as lead. Beamstops have been developed that incorporate an X-ray
sensor such as a pin diode. This enables the sample absorbance to be taken into account for
intensity calibrations.
SAXS data are usually plotted and analysed on a log–log scale. The scattering is usually at
its most intense close to the direct beam, and ultimately falls away as q−4 for Porod scattering.
This means that the detector should ideally have a large dynamic range and a linear response
over as much of that range as possible for accurate collection of the high intensity at low q, while
simultaneously having low noise for accurate collection of the low intensity at higher q.
Most detectors used for SAXS experiments are 2D (e.g. CCD, multi-wire proportional counter
(e.g. Rigaku), image plate or Pilatus). CCD detectors have advantages in their small pixel size,
offering high resolution. Pilatus detectors are more expensive and have a larger pixel size, but
have far superior signal-to-noise than CCD detectors. CCD detectors are available up to 48
megapixels, while the largest Pilatus detector currently available is 6 megapixels. There is a
trade-off between q-resolution and q-range: qmin is determined by the size of the beamstop and
the pixel size (one pixel away from the edge of the beamstop) relative to the X-ray wavelength
and camera length (sample-detector distance); qmax is determined by the physical size of the
detector relative to the X-ray wavelength and camera length. Thus, the q-range can be varied
by either changing the X-ray wavelength or the distance between the sample and the detector.
270 B. Ingham

The detector can also be offset in position relative to the beam to increase the q-range, if direc-
tional information is not required (as is usually the case for isotropic samples measured in a
transmission geometry).
In an XRD experiment, the angles involved are relatively large, and air scatter is generally
not significant. However, this is not the case in a SAXS experiment, and therefore to minimise
the air scattering signal, it is desirable that as much of the X-ray beam path as possible be in
vacuum, particularly between the sample and the detector. It is often not possible or practical
to place a sample directly in vacuum. Instead, samples are loaded into a capillary or appro-
priate cell. The thickness of the sample holder and the sample concentration must be carefully
considered to optimise the scattering signal relative to the sample absorption. Another impor-
tant factor to take into account is the electron density contrast of the particles compared to the
matrix they are in. The higher the electron density contrast (which can easily be calculated and
increases with increasing atomic number), the more pronounced the scattering will be for the
same concentration of particles.
In recording SAXS data for a sample, the background subtraction is often critical. Therefore,
along with each sample, a measurement should also be taken of the sample holder and the matrix
(e.g. for nanoparticles in suspension, the background would be the same sample holder with the
solution only). It is preferable for the background scan to be performed over the same exposure
time as the sample to minimise errors from detector non-linearity, etc.
It can be of interest for some systems where the sample is a suspension of particles, to measure
a dilution series. This enables one to determine the concentration above which the particles start
to interact, that is, the concentration regime where the structure factor must be considered. The
structure factor itself can be approximated by dividing a high-concentration sample by the same
sample of lower concentration.[302]

3.2. Ex situ SAXS studies


Ex situ SAXS is frequently used in nanoparticle characterisation to confirm the average par-
ticle size determined by other methods, for example, XRD and TEM. Nanoparticles may be
measured in their final state as suspensions, encapsulated in a polymer matrix, or deposited as
a thin film (either on a thin substrate and measured in transmission, or measured in a grazing
incidence geometry). However, many literature reports contain little further analysis beyond the
particle size, and some fail to mention the model assumptions used, such as the size distribution
function chosen.
The following examples are literature reports where ex situ SAXS was used in a more
extensive way, either as the focus of the study or where more in-depth analysis was performed.

3.2.1. Ex situ SAXS studies of metal nanoparticles


Stevens at al. performed a comprehensive study on various loadings of Pt nanoparticles on carbon
supports, from 0 to 80 wt% Pt, for use as fuel cell electrodes.[303] For this application, it is
desirable to optimise the catalytic activity while minimising the amount of precious metal used.
SAXS patterns were fitted using a spherical form factor and log-normal size distribution, together
with a Porod term to describe the scattering from the carbon support. The size distribution was
reasonably narrow for small particle sizes (observed at low loadings), and became increasingly
broad as the loading and consequently the particle size increased. Application of the Scherrer
equation to the width of XRD peaks recorded for the same samples gave similar values for
small particle sizes (although the authors noted that for particle sizes below 2–3 nm the width
of the XRD peaks could not be accurately determined), and gave significantly larger values for
Crystallography Reviews 271

higher loadings, which may be due to the polydispersity. This shows the value of using multiple
techniques to characterise the particle size or size distribution.
Lin et al. obtained SAXS patterns of fractal-like aggregate networks of Pt nanoparticles syn-
thesised by a polyol process.[304] These networks have high-surface-area-to-volume ratios,
enabling high catalytic activities to be obtained from a small amount of Pt. The particles formed
short rods that aggregated together. The size distribution (chosen to be a Schulz function) was
relatively broad. The fractal dimension was found to increase over time upon dilution, evidenced
by the SAXS curves having more negative power-law slopes. Pavlopoulou et al. also investi-
gated the formation of a Pt nanoparticle fractal network.[305] They undertook detailed analysis
of the slopes in various regions of the SAXS curves obtained after the various synthesis steps,
since the slope is related to the fractal dimension of the Pt network. The results for two synthesis
methods were compared to the initial microgel network. It was found that in the low pH syn-
thesis, a distorted network of Pt nanoparticles was formed due to homogeneous impregnation of
the microgel by the Pt precursor, while in the high pH synthesis, the Pt distribution was more
heterogeneous and resulted in a non-distorted compressed network.
Another SAXS study of Pt, Au, and Cu nanoparticles formed within a dendrimer polymer
network was reported by Gröhn et al. (Figure 25).[306] The particles were found to have a narrow
dispersity, characterised by the presence of form factor oscillations at high q, and systematically
increased in size with different generations of dendrimer. At higher dendrimer concentrations
( > 10%), a structure factor peak emerged, due to the formation of networks within which the
nanoparticles were encapsulated in close proximity to one another.
In a final example, Ceolín et al. used SAXS to characterise Cu nanoparticles formed within
apoferritin capsules.[307] They used the PDF curve, obtained by the Fourier transformation of
the SAXS curve, to analyse the change in electron density as a result of Cu precursor incor-
poration within the core–shell-like apoferritin protein structure, and subsequent nanoparticle
formation, to gain understanding into the formation mechanism of these particles.

Figure 25. SAXS patterns of polymer networks of dendrimers of different generations (G7–G10), mass
fraction of 1% dendrimer, loaded with H2 PtCl6 . The curves have been translated on the intensity scale for
ease of viewing. From Gröhn et al. [306].
272 B. Ingham

3.2.2. Ex situ SAXS studies of other systems


Biswas et al. studied the formation of ZnO nanorods via solution synthesis, in the presence of
capping molecules.[308] Samples were obtained after different reaction times and measured ex
situ. Nanorods formed in the presence of a capping layer of poly(vinyl pyrollidone) (PVP) were
longer than those formed in the absence of PVP. The SAXS patterns of the capped nanorods
were broadened due to polydispersity; however, the modelling performed by the authors did not
account for this. Nevertheless, the average rod diameter and length obtained from the SAXS
measurements were in good agreement with values obtained from TEM images at the various
times studied, increasing from 8 × 16 nm to 12 × 50 nm over 24 h.
Champagnon et al. used SAXS together with high-resolution TEM and inelastic Raman
scattering to determine the size of Cd(S, Se) quantum dots formed in glasses by thermal
annealing.[309] Guinier analysis was applied to the SAXS data and yielded particle sizes of

Figure 26. Top: Guinier plots for 42 Å diameter CdSe nanoparticles coated with ZnS in Zn/Cd ratios of
(b) 0.42, (c) 0.85, (d) 3.1, and (e) 5.3. The inset shows data for (d, e) at small q values. Bottom: Porod
plots for the same samples together with (a) bare 42 Å diameter CdSe nanoparticles. The curves are fits
to the data with dispersities of (a) 0.11, (b) 0.14, (c) 0.16, and (d) 0.25. Reprinted with permission from
Dabbousi BO, Rodriguez-Viejo J, Mikulec FV, Heine JR, Mattoussi H, Ober R, Jensen KF, Bawendi MG.
(CdSe)ZnS core–shell quantum dots: Synthesis and characterisation of a size series of highly lumines-
cent nanocrystallites. J Phys Chem B. 1997;101:9463–1475 [181]. Copyright (1997) American Chemical
Society.
Crystallography Reviews 273

5–17 nm for the various samples studied, which were consistent with those obtained by other
methods.
Dabbousi et al. reported SAXS data together with Guinier and Porod analyses of bare and ZnS-
coated CdSe nanoparticles dispersed in solution (Figure 26).[181] As the ZnS shell thickness
was increased, the particles became more polydisperse in size, evidenced by broadening of the
features in the Porod plot and having the effect of broadening the features in the ultraviolet part
of the optical absorption spectrum. The results were related to XRD and TEM analyses, and
photoluminescence resulting from the change in particle size.
Kagan et al. also reported SAXS data for CdSe quantum dots.[310] Samples of quantum dots
dispersed in poly(vinyl butyral) were used as dilute systems from which the form factors of
the particles could be obtained. The particles had a narrow size distribution. These data were
compared to SAXS obtained from the same particles prepared in densely packed films. The
resulting curves were a superposition of a structure factor peak on the form factor scattering
observed for the dilute samples.
A GISAXS study of germanium quantum dots grown on Si(111) substrates by molecular beam
epitaxy was reported by Metzger et al.[311] GISAXS was observed as ‘wings’ on both sides of
the reflected beam, which varied in intensity as the sample was rotated. The variations were
explained by the triangular shape of the islands. This information could not be obtained from
transmission SAXS measurements. Similar intensity oscillations resulting from the particle shape
were observed in the diffuse scattering of the diffraction peaks.

3.3. In situ SAXS studies


As for in situ XRD, in situ SAXS is useful for following changes in particle size during kinetic
processes such as synthesis, annealing, and other processing treatments. As stated earlier, cells
for in situ measurements must be compatible with the environment and the X-ray geometry.[188]
Kapton® and other polymers are generally not suitable for windows for SAXS cells due to strong
scattering features from the polymer structure that often occur in similar positions to the scat-
tering from the nanoparticles under study. Thin films of these materials may be used, or other
materials such as mica, quartz, or silicon nitride, which have very low scattering that is reason-
ably constant over the q range where the scattering signal from nanoparticles is observed. In situ
measurements can also be performed in quartz capillaries. To avoid beam-induced fouling of the
cell windows, it may be possible in some instances to create a ‘windowless’ cell using a liquid jet
[312] or levitating droplet.[313] SAXS experiments have also been performed on nanoparticles
formed in flames.[314,315]
Time-resolved studies have been performed for nanoparticles in a variety of solutions and
using various apparatus. Generally, these fall into two categories: mixing experiments, where
the timescale is determined by the length of tubing from the mixing point of the reagents to
the measurement point and the flow rate; and bulk solution experiments. The former is useful
for studying very short experiment times after mixing, down to the millisecond regime. The
measurement time can be much longer than the timescale of the reaction measurement, meaning
that such experiments can be performed using laboratory instruments. The major drawback is that
depending on the flow rate and the measurement time required, large solution volumes may be
generated. For bulk solution experiments, the same solution is measured in a continuous fashion.
High-flux synchrotron radiation can be used to reduce the measurement time. This is convenient
for reactions that take a few minutes to several hours.

3.3.1. In situ SAXS during synthesis of metal nanoparticle systems


Several in situ SAXS studies have been performed on metal nanoparticles made of heavy ele-
ments such as Au and Pt. Due to their high electron density contrast, ρ (see Equation (4)),
dilute suspensions can be measured, such as the early stages of particle nucleation.
274 B. Ingham

A simple and versatile experimental setup, which was used by Henkel et al. to study
Au nanorod formation,[316] consists of a large reaction vessel from which the solution was
continually pumped through a capillary. X-rays were passed through the capillary and SAXS pat-
terns obtained. The reaction took around 30 min to reach completion. From these measurements,
the size and total volume of particles were obtained
 as a function of time. The total volume in this
case was obtained from the invariant Q = q2 I(q)dq. In a similar reaction, Aux Cu1−x nanorods
were produced that were generally smaller in size and the reaction progressed in a shorter time. In
both cases, a change in the growth mechanism from 1D to 3D growth was observed, presumably
due to exhaustion of the precursor.
Harata et al. used a UV reaction vessel to perform in situ SAXS during photoreduction of Ag
[317] and Pd and Rh [318] nanoparticles in solution. In the Ag study,[317] various precursor
concentrations were studied. The data were fitted to the SAXS equation (Equation (4)), with a
Schultz–Zimm size distribution function. Three growth stages were ascertained. The first stage
corresponded to an incubation period where particles were rapidly nucleating via an autocatalytic
reaction. In the last stage, Ostwald ripening was the dominant process. In the intermediate stage,
both processes were occurring and competing. For the Pd and Rh study,[318] it was found that
the Pd particles grew by nucleation followed by Ostwald ripening, but the Rh particles grew
through autocatalysis, resulting in a broad log-normal size distribution. The relative effect of
these processes in the Pd system was dependent on the precursor concentration.
LaGrow et al. studied the formation of Ni nanoparticles in solution in a sealed reaction cell
as a function of the nickel precursor:surfactant ratio.[319] They observed growth of the particles
to maximum sizes that decreased as the precursor:surfactant ratio increased. At a 1:1 ratio, the
particles had a narrow dispersion. In a number of cases, the scattering curves could be fitted with
a structure factor of unity (i.e. dilute model) at early times, but after a critical point in the reaction
a prominent structure factor peak emerged (Figure 27(a)). The data were best fitted using a sticky
hard sphere structure factor,[289] the rationale being that a critical number of particles reached
a size above which they became ferromagnetic, which then induced magnetic attraction of the
particles in solution. After the structure factor peak appeared, the size distribution broadened
(Figure 27(b)), which was attributed to the particles inside the magnetically attracted aggregates
ceasing to grow due to local depletion of the precursor, while those on the outside of the aggre-
gates continued to grow. Curves obtained of the total volume of particles (obtained by fitting
Equation (4)) as a function of time were consistent with a two-stage nucleation and autocatalytic
growth model (Figure 27(c)).[320]
Abécassis et al. performed SAXS studies investigating the formation of Au nanoparticles by
reduction of AuCl3 by BH− 4 in toluene using a fast mixing stopped-flow cell setup.[321,322] The
experiments were correlated with UV–visible spectroscopy [321] and XANES [322] to yield
information about the conversion of Au3+ to Au0 , on a timescale with 100 ms resolution. Two
different surfactant ligands were used: one having an amine functional group and the other a
carboxylic acid. The amine produced smaller nanoparticles (1.2 nm radius) that ceased growing
after 3 s, while the acid produced 3.7 nm radius particles and took 12 s to stabilise. A model
comprising nucleation and growth, specifically excluding aggregation, was used to fit the data.
Takesue et al. reported an investigation on Ag nanoparticle formation where the time resolu-
tion was a mere 0.18 ms,[323] obtained using a mixing setup with high flow rate and a focused
X-ray beam. Three distinct periods of the formation mechanism were identified. Ag13 clusters
were found to be an instrumental intermediate in the formation of the final particles.
The growth of Au nanoparticles was measured on a similar timescale by Polte et al. using a
laboratory instrument setup and a microfluidic mixing device (Figure 28(a)).[324] They observed
the formation of Au nuclei consisting of a few tens of atoms within 200 ms, which then coalesced
at later times (Figure 28(b)). The microfluidic mixer could be used to measure SAXS patterns for
reaction times up to 140 s. For longer times, aliquots of the reaction solution were obtained and
Crystallography Reviews 275

(a)

(b)

(c)

Figure 27. (a) Background-subtracted SAXS data showing development of form factor scattering fol-
lowed by emergence of a correlation peak, (b) size distribution, (c) total volume of particles fitted with
a two-stage autocatalytic growth model. Reprinted (adapted) with permission from LaGrow AP, Ingham
B, Toney MF, Tilley RD. Effect of surfactant concentration and aggregation on the growth kinetics of
nickel nanoparticles. J Phys Chem C. 2013;117:16709–16718 [319]. Copyright (2013) American Chemical
Society.

measured. A similar mixing device was used by the same authors for an in situ SAXS study of
Ag nanoparticle formation.[312] In that case, the growth mechanism is more complex. To cover
the full timescale over which all of the changes occurred, aliquots were taken at longer times (up
to 100 min after mixing). A number of critical points were noted, namely a dramatic increase
in the volume fraction of particles, and a sudden onset of rapid coalescence of particles. A third
experimental setup was also used: a liquid jet, which delivered solution from a reaction vessel
to the X-ray beam. This had the advantage of removing the capillary where the X-ray beam was
sampling the solution, eliminating the problem of beam-induced fouling due to the aggregation
of particles on the capillary walls.
276 B. Ingham

(a)

(b)

Figure 28. (a) Schematic diagram of a SAXS flow cell incorporating a mixing device. (b) SAXS data
collected using the flow cell in (a) during Au nanoparticle synthesis. Red lines indicate fits to the data
using a Schulz–Zimm distribution. Reprinted (adapted) with permission from Polte J, Erler R, Thüne-
mann AF, Sokolov S, Ahner TT, Rademann K, Emmerling F, Kraehnert R. Nucleation and growth of gold
nanoparticles studied via in situ small angle X-ray scattering at millisecond time resolution. ACS Nano.
2020;4:1076–1082 [324]. Copyright (2010) American Chemical Society.

Other SAXS studies of in situ nanoparticle formation in non-solution environments have


also been reported. Carrado et al. investigated Pt nanocluster formation within polymer-clay
composites.[325] The composites were loaded with a Pt(II) salt precursor solution and deposited
as films, before being mounted in a customised sample holder and heated to temperatures
up to 200°C while SAXS patterns were collected. The data were fitted using the Unified fit
model [290–292] to obtain values for the radius of gyration of the Pt particles as a function of
temperature, which was about 5 nm and relatively independent of temperature.
Crystallography Reviews 277

Sakamoto et al. studied Pd nanoparticle formation within a block copolymer gel.[326] The
Pd(II)-loaded gel was mounted in a sample holder and heated rapidly to 80–100°C, while SAXS
patterns were recorded as a function of time. A hard sphere structure factor was required to fit the
data, indicating that the Pd nanoparticles formed in close proximity to one another. The kinetics
of the total volume versus time were also analysed and related to the growth mechanism.
An in situ study of Ag nanoparticles forming on an HOPG electrode in an electrochemical
cell was presented by Ustarroz et al.[327] The electrode formed one of the windows of the
transmission cell. SAXS patterns were collected as a function of time while the electrochemical
potential was pulsed to induce nanoparticle formation and subsequent growth. Size distributions
were calculated from the data (assuming functions for the structure factor and form factor) and
showed an increase in the primary particle size with time, with possible agglomeration of the
particles.
In a final example, Polte et al. studied the growth of Au nanoparticles within a levitating
droplet using simultaneous SAXS and XANES.[313] The droplet acted as a ‘wall-free reaction
vessel’ and its position could be controlled to ± 20 µm. Evaporation of solvent (water) was
countered using piezoelectric nozzles to replace the solvent volume by feeding additional sol-
vent to the droplet. SAXS patterns were fitted to Equation (4) to obtain the particle radius as
a function of time. The total particle volume was obtained from the invariant Q. The XANES
spectra showed the reduction of Au(III) to Au(0).

3.3.2. In situ SAXS during post-synthesis processing of metal nanoparticle systems


In addition to detecting particle growth during real-time synthesis, in situ SAXS can be used to
monitor changes in particle size during post-synthesis treatment, such as thermal annealing or
electrochemical processing.
As mentioned in Section 2.3.2, Ingham et al. performed a combined XRD and SAXS study
investigating the thermal coalescence of Au nanoparticles as a function of time at fixed temper-
atures between 200°C and 350°C.[89] Particles were drop-cast onto mica windows and heated
while SAXS patterns were collected. The initial SAXS scans showed a prominent correlation
peak, corresponding to hexagonal close packing of the nanoparticles, which had a reasonably
narrow size distribution with a mean size of 10 nm. Over time, the correlation peak disappeared
and was replaced by a broader SAXS feature, which moved to lower q (Figure 29). This was due
to the formation of aggregates. Aggregation of the particles at 250°C was complete in around half
the time of the grain growth of the particles and reduction of stacking and twin fault densities as
determined by XRD.
Winans et al. reported a GISAXS study on Pt nanoparticles during thermal annealing on silicon
wafers up to 400°C.[328] Horizontal and vertical slices of the data were extracted and analysed
using Guinier plots (ln(I) versus q2 ). The vertical slices, which corresponded to the height of
the clusters on the surface, did not show any change with temperature. The horizontal slices,
however, showed a change for temperatures above 320°C, corresponding to an increase in the
lateral size of the clusters. These results indicated that the particles were agglomerating and
forming oblate spheroids with the same height as the original particles.
Smith et al. studied the coarsening of Pt nanoparticle catalysts of two different loadings on car-
bon supports during repeated electrocatalytic cycling.[329] The initial particle size distribution
was reasonably broad and was fitted to a Schulz distribution. For 20% loading, the initial mean
particle size was 2.2 nm and increased rapidly to 3.2 nm over 10 h of potential cycling (between
0.4 and 1.4 V (versus Ag/AgCl) at 10 mV s−1 ), where it plateaued before rising again. For a
sample with 40% loading, the initial mean particle size was 3.2 nm and steadily increased to 4.2
nm over 15 h of potential cycling. Examination of the particle size during cycling (Figure 30)
revealed a general increasing trend in particle size, with partial reversibility. The maximum in
278 B. Ingham

(a)

(b)

(c)

Figure 29. Left: Selected SAXS data at various times as indicated (solid lines) and fits (dashed lines) for
an Au nanoparticle sample heated at 250°C. Right: TEM images of Au nanoparticles at various stages of
coalescence. Reprinted (adapted) with permission from Ingham B, Lim TH, Dotzler CJ, Henning A, Toney
MF, Tilley RD. How nanoparticles coalesce: an in situ study of Au nanoparticle aggregation and grain
growth. Chem Mater. 2011;23:3312–3317 [89]. Copyright (2011) American Chemical Society.

Figure 30. Particle size obtained from SAXS analysis (black line) and applied potential (grey bars) dur-
ing potential cycling of 20 wt% Pt catalyst particles on XC-72 Vulcan carbon. Reprinted with permission
from Smith MC, Gilbert JA, Mawdsley JR, Seifert S, Myers DJ. In situ small-angle X-ray scattering obser-
vation of Pt catalyst particle growth during potential cycling. J Am Chem Soc. 2008;130:8112–8113 [329].
Copyright (2008) American Chemical Society.

particle size occurred just after the voltage maximum, which the authors attributed to either for-
mation of an oxide layer on the surface of the particles, or deposition of solution phase Pt due to
electrochemically induced Ostwald ripening.
Yu et al. also studied catalytic Pt nanoparticles during electrochemical cycling.[330] As-
prepared particles 2.0 nm in size were annealed at various temperatures to produce particles with
different initial mean diameters for the electrochemical studies, with the aim being to identify the
major growth mechanisms. No change in particle size was observed when the maximum cycling
potential was 0.96 V/RHE, but there was a significant change when the maximum potential was
Crystallography Reviews 279

Figure 31. SAXS patterns for a sample of 2.5% polyacrylamide gel with 3.5 × 10−3 g cm−3 magnetic
particles, with various magnetic fields B applied in the horizontal direction. Reprinted figure with permis-
sion from Teixeira AV, Morfin I, Ehrburger-Dolle F, Rochas C, Geissler E, Licinio P, Panine P. Scattering
from dilute ferrofluid suspensions in soft polymer gels. Phys Rev E. 2003;67:021504 [331]. Copyright
(2003) American Physical Society. http://dx.doi.org/10.1103/PhysRevE.67.021504

1.16 V/RHE. This provided conclusive evidence that particle growth was due to electrochemical
Ostwald ripening through the dissolution of Pt species.
A final interesting example of the use of in situ SAXS is a study reported by Teixeira et al.
on ferrofluid suspensions in a magnetic field.[331] The ferrofluid consisted of particles approxi-
mately 5 nm in size, dispersed in a soft polymer gel. These particles formed fractal-like aggregate
chains. When a magnetic field was applied, the chains aligned along the direction of the field,
and the scattering pattern was compressed in that direction (Figure 31). Due to the elasticity of
the gel, the field required to induce this alignment was higher for higher polymer concentrations.

3.3.3. In situ SAXS during synthesis of oxide systems


Pulcinelli et al. have reported a number of in situ SAXS studies of ZnO nanoparticle synthesis via
hydrolysis.[332,333] In their earlier work, they reported SAXS data that showed the development
of objects of two different sizes, which were attributed to the primary particles and aggregates
of these particles that formed at later times.[332] In more recent work, they combined in situ
SAXS with UV–visible and X-ray absorption spectroscopies to follow the nanoparticle growth
and aggregation processes in more detail.[333] The formation of crystalline ZnO was observed
with the growth of a peak in the Fourier transformed X-ray absorption spectrum corresponding to
the first Zn–Zn coordination shell, which was absent in scans of the precursor solution. The UV–
visible spectra, which were collected simultaneously, relate to the electronic structure. These are
affected by the particle size, showing a steady shift of the absorption edge to higher wavelengths
with time, indicating a narrowing of the semiconducting band gap. After 500 min, the baseline
absorbance started to increase dramatically due to light scattering from large objects; in this case
aggregates of ZnO particles. The growth kinetics of the particle size as determined from UV–
visible spectroscopy followed an oriented attachment model.[334,335] A combined simultaneous
SAXS and UV–visible spectroscopy study was performed in a separate experiment. The UV–
visible spectra showed development of the same features on the same timescale as the XAS/UV–
visible experiment. The SAXS patterns could initially be described as a Guinier region at low q,
where the intensity plateaued, and a Porod region at high q, where the intensity followed a q−4
power law, indicating that the particles were dense with smooth surfaces (i.e. non-fractal). Over
time, the turnover point corresponding to Rg moved to lower q, and the feature broadened, so
that after 600 min a power law at low q was observed with a slope of − 1.76. This indicates that
fractal objects were formed, made up of the smaller dense particles.
Boissiere et al. performed in situ SAXS during spray drying of SiO2 particles.[336] The dried
particles were spherical mesostructured assemblies of well-ordered particles approximately 3
nm in diameter, evidenced in the SAXS patterns by a strong correlation peak. This peak showed
280 B. Ingham

a non-linear temperature dependence, shifting to higher q-values as the temperature increased


above 120°C. This shift may be due to a phase transition, and/or shrinkage of the assembled
structures.
Silica particles produced via flame pyrolysis have been studied using in situ SAXS by
Beaucage et al.[314,315] In their first report,[314] a diffusion flame apparatus was positioned
such that the X-ray beam passed through the flame, and SAXS patterns from the particles formed
were detected as a function of position. Larger particles were observed at greater heights above
the nozzle (corresponding to longer times in the flame). The data were fitted using the Unified
model.[290–292] A plot of the particle size versus residence time yielded a value for the diffu-
sion coefficient of the particle growth. A subsequent study [315] involved expanding the q-range
to lower values using ultra-small-angle X-ray scattering (USAXS) in order to extract information
about both the primary particles and mass-fractal aggregates of these that formed, including pri-
mary particle size and standard deviation, mass-fractal dimension, number concentration, silica
volume fraction, aggregate size, and number of primary particles per aggregate.

3.4. Anomalous SAXS


As with a-XRD (Section 2.4), anomalous SAXS (ASAXS) [337,338] is a technique that pro-
vides element-specific scattering information. It requires an X-ray beam that can be tuned to
energies close to the absorption edge of an element of interest; in other words, a synchrotron
source must be used. The electron density contrast, ρ, depends on the atomic structure factor
Fhkl = f0 (q) + f  (E) + if  (E). The difference between two scans at energies near the absorp-
tion edge of a particular element will display only the scattering components arising from
structures containing that element. This is an effective way of removing a background scat-
tering component, which may itself be nanostructured and give rise to a significant SAXS
signal.[246,338–343] ASAXS can also be used to obtain compositional information [338] and
the relative distribution of different elements, for example, distinguishing between core–shell
and alloyed bimetallic nanoparticles.[343–345]
Haubold et al. investigated Pt catalytic particles on carbon supports, using ASAXS to remove
the background scattering from the support material.[339] Both reduced and oxidised particles
were examined in this way, showing a change in size consistent with the formation of a sur-
face oxide layer. In a later publication,[340] the same authors reported a study on supported Pt
particles of different loadings, which resulted in different sized particles. In the case of the 60
wt% Pt/C sample, the turnover corresponding to the particle size was obscured by the support
scattering and could only be determined from the anomalous component (Figure 32).
In another study, the same authors reported the use of ASAXS to distinguish the scattering
from Pt nanoparticles from the background scattering contribution from organic crosslinking
molecules.[346] While the mean particle radius did not change over time, the mass fraction
increased as more particles were formed from the precursor. The authors were able to con-
clude that a monodisperse population of Pt nanoparticles was formed. They also determined
the decomposition rate of the precursor species.
A study by Vad et al. applied ASAXS to three-dimensional Pt nanoparticle networks.[347] The
anomalous components of the SAXS curves showed the same features as the total scattering: a
q−4 (Porod) slope at low q, and a correlation peak at q = 0.2–0.3 Å−1 depending on the linking
molecule used. This study demonstrated the use of ASAXS in distinguishing between matrix
inhomogeneities (e.g. pores) and particles present in the matrix.
Bóta et al. reported the use of ASAXS to study CdS nanoparticle formation in multilamellar
vesicles.[342] SAXS patterns collected at different energies showed a rise in intensity at low q
and a peak at q = 0.1 Å−1 (Figure 33). The intensity of the low q feature varied with energy, but
the peak did not, indicating that the low q feature was due to the CdS particle scattering while
Crystallography Reviews 281

Figure 32. SAXS data collected for three Pt/C electrodes with loadings as indicated in their reduced
state at 0.25 V (versus Ag/AgCl), at two X-ray energies: 10,350 eV (filled circles) and 11,546 eV
(open circles). The difference is shown with open squares and fitted to a log-normal particle size
distribution (solid lines). From Haubold HG, Wang XH, Goerigk G, Schilling W. In situ anoma-
lous small-angle X-ray scattering investigation of carbon-supported electrocatalysts. J Appl Crystallogr.
1997;30:653–658 [340]. Reproduced with permission of the International Union of Crystallography.
http://dx.doi.org/10.1107/S0021889897002422
282 B. Ingham

Figure 33. ASAXS of dipalmitoyl-sn-glycero-3-phosphatidylcholine (DPPC)/water vesicles with CdS


nanoparticles, measured at energies close to the Cd K-edge (E1 = 26120 eV, E2 = 26653 eV, E3 = 26711
eV). Reprinted with permission from Bóta A, Varga Z, Goerigk G. Biological systems as nanoreactors:
Anomalous small-angle scattering study of the CdS nanoparticle formation in multilamellar vesicles. J
Phys Chem B. 2007;111:1911–1915 [342]. Copyright (2007) American Chemical Society.

the peak arose from the vesicle layers. The size of the CdS nanoparticles was determined to be
6.9 nm.
ASAXS was used by Benedetti et al. to study Au, Pd, and AuPd alloy particles [343] and
by Haubold et al. to study Pt, Rh, and PtRh alloy particles [344] as catalysts on active carbon
supports. In both cases, the signal from the particles was distinguished from the nanostructured
background. In addition, by performing measurements at both edges for the alloy samples, the
authors were able to determine that a homogeneous alloy was formed, as the anomalous signals
gave similar results.
Yu et al. reported an ASAXS study of electrocatalytic PtCu alloy particles.[345] Measurements
were recorded near the Pt L3 and Cu K edges. Annealing the particles had the effect of increasing
the particle size. The effect of electrochemical treatment to preferentially dissolve some of the
Cu was explored. The nanoparticles annealed at 950°C showed very little difference between the
ASAXS scans at the two absorption edges, indicating that the particles were homogeneous in
composition. However, after dealloying, the anomalous scattering obtained at the Pt edge shifted
to higher values of q, while the anomalous scattering obtained at the Cu edge shifted to lower
values of q. This implied that the Cu-containing part of the particle after dealloying was smaller;
the particles developed a Pt-rich shell.
ASAXS can also be performed in a grazing incidence geometry. Lee et al. reported such a
study on Pt nanoparticles.[348] Obtaining the anomalous component of horizontal and vertical
cuts of the 2D GISAXS data enabled the authors to distinguish between the scattering due to the
surface roughness of the substrate, and the scattering due to the Pt particles.

3.5. SAXS of ordered nanoparticle assemblies


Colloidal crystals, or superlattices, are highly ordered arrangements of monodisperse
nanoparticles.[349] Various crystal structures have been generated via self-assembly of nanopar-
ticles of different materials and sizes, having similar symmetry relations as atomic crystals
[129,350,351] and other tiling motifs.[352] Superlattices have applications in the bottom-up
Crystallography Reviews 283

fabrication of electronic and magnetic devices, with functionality imparted by the nanoparticle
systems chosen. Formation of superlattices as thin films or as self-supporting crystals [129,353–
355] or membranes [356] is critically dependent on controlling the interaction strength between
the nanoparticles through the selection of capping molecules and carefully controlling the solvent
evaporation rate.[129,353,354]
These superlattice structures are periodic on a nanometre length scale and, in addition to
diffraction peaks arising from the atomic planes within the individual nanoparticles, can also
exhibit diffraction peaks from the highly ordered arrangement of the nanoparticles at small
angles. The benefit of using X-ray scattering techniques to probe the structure, over other tech-
niques such as electron microscopy, is that X-ray techniques provide information about the
average long-range crystalline order of the superlattice throughout the entire sample.
In one of the earliest reports on the formation of nanoparticle superlattices, Murray et al.
formed fcc arrangements of CdSe quantum dots as films and as colloidal crystals, using par-
ticles of different sizes and varying the surfactant chain length to control the inter-particle
spacing.[355] The X-ray scattering curves showed a series of peaks in the SAXS region
corresponding to the fcc particle packing, as well as peaks in the conventional XRD region cor-
responding to the atomic CdSe crystal structure (Figure 34). The most intense (111) superlattice
peak shifted consistently in position with particle size and spacing.

Figure 34. XRD patterns from CdSe nanoparticle superlattices: (a) colloidal crystals of particles 48 Å
in diameter and (b) a thin film array of particles 48 Å in diameter with the superlattice oriented in the
111SL direction. The intensity in the SAXS region is plotted on a log scale, and in the WAXS region
on a linear scale. Inset: the (111)SL reflection for trioctylphosphine oxide-capped particles with diameters
of (c) 63, (d) 54, (e) 47, (f) 39, and (g) 35 Å; and for 48 Å particles capped with (h) hexadecyl phos-
phate, (i) trioctylphosphine oxide, and (j) tributyl phosphine oxide to vary the inter-particle spacing. From
Murray CB, Kagan CR, Bawendi MG. Self-organisation of CdSe nanocrystallites into three-dimensional
quantum dot superlattices. Science. 1995;270:1335–1338 [355]. Reprinted with permission from AAAS.
http://www.sciencemag.org/content/270/5240/1335
284 B. Ingham

The dependence of the sharp peaks observed in the SAXS region on the extent of the nanoparti-
cle ordering was demonstrated in work by Korgel et al. [357] and Murray et al.[353] Korgel et al.
published a series of SAXS patterns of an fcc superlattice of Ag nanoparticles as a function of
annealing temperature.[357] Above 140°C, the sharp superlattice peaks rapidly decayed in inten-
sity due to desorption of the capping ligands, leading to disruption of the nanoparticle ordering
and eventually resulting in aggregation of the Ag nanoparticles. Murray et al. reported SAXS pat-
terns of films of Ni nanoparticles deposited from different solutions at different rates.[353] This
had a profound effect on the long-range ordering of the nanoparticles. For quickly deposited
films, only form factor scattering was observed. However, for films of particles dried slowly
from higher boiling point solvents, the dominant features were sharp peaks related to the spacing
of the nanoparticles. Similar results were obtained by Heitsch et al. [358] who studied the extent
of order in FePt monolayers deposited by various techniques (Langmuir–Blodgett, Langmuir–
Schäfer, and spin-coating) using GISAXS, with the Langmuir–Blodgett-deposited film having
the highest degree of ordering and the spin-coated film having the lowest.
Ordered arrays of Au nanoparticles linked with DNA have been assembled by Park et al. [359]
and by Nykypanchuk et al. [360]. Park et al. showed SAXS patterns for both fcc and bcc Au-
DNA structures.[359] The structure was controlled through the selectivity of the DNA linker
molecules. Nykypanchuk et al. investigated the effect of temperature on the self-assembly pro-
cess of DNA-capped Au nanoparticles using SAXS.[360] After raising the temperature above a
point where the nanoparticles dissociate from their initially disordered structures, highly ordered
assemblies formed, as evidenced by the appearance of higher order diffraction peaks in the SAXS
region. The propensity of the system to form highly ordered structures was dependent on the
chain length of the DNA linker molecules.
A number of GISAXS experiments have been reported on nanoparticle superlattice films,
giving insights into the ordering both in-plane and out-of-plane. As mentioned earlier, Heitsch
et al. studied the effect of the film deposition technique on the extent of order for FePt
nanoparticles.[358] Disch et al. reported GISAXS experiments on self-assembled iron oxide-
truncated nanocubes (Figure 35).[361] Samples assembled through rapid evaporation of solvent
showed broad features, while those assembled through slow evaporation of solvent (especially
in the presence of an applied magnetic field) showed additional diffraction spots corresponding
to a body-centred tetragonal packing of the nanoparticles.
Siffalovic et al. also studied the self-assembly of iron oxide (Fe3 O4 ) nanoparticles.[362] They
performed GISAXS measurements of the dried sample, showing that the particles (6.0 ± 0.4 nm
in size) formed ordered arrays with inter-particle separation of 7.7 nm. The ordering was weak
past the first nearest neighbours, as evidenced by the lack of features past the first maxima in
either the horizontal or the vertical direction. The authors also conducted an in situ GISAXS
study during solvent evaporation, to determine whether the self-assembly of the nanoparticles
took place at the substrate–drop interface, or at the drop–air interface. The X-ray beam was
directed at each of these interfaces in two separate experiments. Satellite reflections indicating the
presence of an ordered layer of nanoparticles were only observed at the substrate–drop interface.
The only signal observed from the drop surface was diffuse scattering from the solution.
The Korgel group reported a GISAXS study on a binary Au/Fe2 O3 nanoparticle superlat-
tice system.[363] Due to the difference in size between the Au and Fe2 O3 nanoparticles used,
the superlattice had an orthorhombic Cmmm structure. The sample had regions containing dis-
locations, evidenced by the presence of additional peaks, and in some instances, spotty rings,
corresponding to crystalline regions with other orientations.
A final example, also by the Korgel group, is a GISAXS study investigating the thermal
stability of a Si/Au nanoparticle superlattice (Figure 36).[364] Superlattices formed from the
individual Si and Au particle systems had the fcc and bcc structure, respectively. When the two
particle systems were combined, an orthorhombic structure was observed. Upon heating, the
Crystallography Reviews 285

(a) (b)

(c)

Figure 35. GISAXS patterns of iron oxide nanocube arrays, self-assembled under (a) rapid evaporation,
(b) slow evaporation, and (c) slow evaporation in the presence of a perpendicular magnetic field. In (c)
body-centred tetragonal reflections are indicated by triangles (up triangles are with reflection of the X-ray
beam at the substrate, while down triangles are without). Scale bars in the micrographs (insets) are 100
nm. Lateral scattering at Qz = 0.34 nm−1 (intensity on a log scale) is shown beneath each GISAXS pat-
tern. From Disch S, Wetterskog E, Hermann RP, Salazar-Alvarez G, Busch P, Brückel T, Bergström L,
Kamali S. Shape induced symmetry in self-assembled mesocrystals of iron oxide nanocubes. Nano Lett.
2011;11:1651–1656 [361]. Used with permission. http://pubs.acs.org/doi/pdf/10.1021/nl200126v

structure was maintained. The out-of-plane lattice parameter increased as the temperature was
raised, until at 200°C where it abruptly decreased. GISAXS patterns after this transition showed
significant intensity at qx = 0, corresponding to migration of the Au nanoparticles out of the
superlattice and their subsequent coalescence into a thin layer.
286 B. Ingham

(a) (b) (c)

(d) (e) (f)

Figure 36. GISAXS patterns of a simple hexagonal AB2 -type binary superlattice of Si and Au nanocrys-
tals heated to (a) 20°C, (b) 50°C, (c) 95°C, (d) 152°C, and (e) 199°C, compared with (f) after heating
at 200°C for 10 min. Reprinted with permission from Yu Y, Bosoy CA, Smilgies DM, Korgel BA. Self-
-assembly and thermal stability of binary superlattices of gold and silicon nanocrystals. J Phys Chem Lett.
2013;4:3677–3682 [364]. Copyright (2013) American Chemical Society.

4. Future directions
The recent development of X-ray laser facilities around the world opens many possibilities for
studying dynamic processes in a wide variety of systems, including nanoparticles. An X-ray free
electron laser (FEL) is a highly coherent, pulsed X-ray source.[365,366] Currently, operational
X-ray FELs are:
• FLASH (Germany), operational since 2005, λ = 4.2–45 nm [367,368]
• LCLS (USA), operational since 2009, λ = 1–40 Å [369,370]
• SACLA (Japan), operational since 2011, λ = 0.6–3 Å [371,372]
• FERMI (Italy), operational since 2011, λ = 4–100 nm [373,374]
Other X-ray FELs currently under construction include:
• SwissFEL (Switzerland), user operations anticipated 2016, λ = 1–70 Å [375,376]
• European XFEL (Germany), user operations anticipated 2017, λ = 0.5–60 Å [377]
• PAL-XFEL (Korea), user operations anticipated 2017, λ = 0.6–60 Å [378,379]
Due to the small spot size and high intensity of the X-rays (up to 1013 photons per pulse,
having 10–100 fs duration), the sample exposed to the beam is usually destroyed. However, suf-
ficient information about the sample can be obtained prior to its disintegration. It is possible
to measure individual molecules, and by extension, individual nanoparticles. Coherent diffrac-
tion imaging has been used to investigate the size distributions of and elemental distributions
within Ag nanocubes and Au/Ag nanoboxes 100–200 nm in size,[380] Co@SiO2 core–shell par-
ticles 10–30 nm in size,[381] and ellipsoidal iron oxide particles 50 × 250 nm.[382] Hundreds
of individual particles are measured and the images processed through Fourier transformation
to reconstruct the original nanoparticle. Another recent study at the LCLS used X-ray photon
Crystallography Reviews 287

correlation spectroscopy to measure the relaxation dynamics of Au nanoparticles suspended in


polymer melts.[383] While these techniques differ markedly from the X-ray scattering tech-
niques previously discussed here, they offer great potential by enabling one to obtain a complete
set of structural information about an individual nanoparticle, rather than an ensemble average.

5. Conclusion
This review has provided a theoretical and practical overview of X-ray scattering techniques,
namely XRD and SAXS, applied to nanoparticles. Descriptions of the theory of each technique
have been illustrated by a wide range of literature examples, showing how measurements of
particle size, crystallinity, defect structures, and superlattice ordering can be obtained using non-
destructive, laboratory-accessible methods. Recent advances in the techniques have also been
discussed, including a-XRD and SAXS, and the PDF technique. Anomalous (or resonant) scat-
tering has great potential for differentiating between different elemental species; it can be used
to distinguish between the nanoparticles of interest and a nanostructured support or to determine
the structure of multi-metallic particles. PDF analysis is pushing the boundaries of structural
characterisation of materials that may have limited crystallinity. The advantages of using high-
flux synchrotron X-ray sources have been demonstrated through in situ investigations into the
synthesis and processing of a wide variety of nanoparticle systems. Such investigations yield
insights into the kinetic processes involved, which are often different from bulk mechanisms due
to the small size of nanoparticles and the importance of surface effects.

Disclosure statement
No potential conflict of interest was reported by the author.

Notes on contributor
Dr Bridget Ingham is a Senior Research Scientist at Callaghan Innovation (for-
merly Industrial Research Ltd.). Following her PhD (Physics, Victoria University
of Wellington, 2005), she spent two years as a post-doc in the Department of
Materials at Imperial College London and the Stanford Synchrotron Radiation
Lightsource, where she developed her current expertise in the use of synchrotron
techniques for investigating nanomaterials, particularly X-ray diffraction, small-
angle X-ray scattering, and X-ray absorption spectroscopy.

ORCID
Bridget Ingham http://orcid.org/0000-0003-4642-8064

References
[1] Wang AZ, Langer R, Farokhzad OC. Nanoparticle delivery of cancer drugs. Annu Rev Med.
2012;63:185–198.
[2] Singh R, Lillard JW Jr. Nanoparticle-based targeted drug delivery. Exp Mol Pathol. 2009;86:215–
223.
[3] Dreaden EC, Mackey MA, Huang X, Kang B, El-Sayed MA. Beating cancer in multiple ways using
nanogold. Chem Soc Rev. 2011;40:3391–3404.
[4] Krishnan KM, Pakhomov AB, Bao Y, Blomqvist P, Chun Y, Gonzales M, Griffin K, Ji X, Roberts
BK. Nanomagnetism and spin electronics: materials, microstructure and novel properties. J Mater
Sci. 2006;41:793–815.
288 B. Ingham

[5] Singamaneni S, Bliznyuk VN, Binek C, Tsymbal EY. Magnetic nanoparticles: recent advances in
synthesis, self-assembly and applications. J Mater Chem. 2011;21:16819–16845.
[6] Berry CC, Curtis ASG. Functionalisation of magnetic nanoparticles for applications in biomedicine.
J Phys D: Appl Phys. 2003;36:R198–R206.
[7] Ito A, Shinkai M, Honda H, Kobayashi T. Medical application of functionalized magnetic nanopar-
ticles. J Biosci Bioeng. 2005;100:1–11.
[8] Frey NA, Peng S, Cheng K, Sun S. Magnetic nanoparticles: synthesis, functionalization, and
applications in bioimaging and magnetic energy storage. Chem Soc Rev. 2009;38:2532–2542.
[9] Gallo J, Long NJ, Aboagye EO. Magnetic nanoparticles as contrast agents in the diagnosis and
treatment of cancer. Chem Soc Rev. 2013;42:7816–7833.
[10] Austin LA, Kang B, Yen CW, El-Sayed MA. Plasmonic imaging of human oral cancer cell com-
munities during programmed cell death by nuclear-targeting silver nanoparticles. J Am Chem Soc.
2011;133:17594–17597.
[11] Huang X, Jain PK, El-Sayed IH, El-Sayed MA. Gold nanoparticles: interesting optical properties and
recent applications in cancer diagnostics and therapy. Nanomedicine (London). 2007;2:681–693.
[12] Daniel MC, Astruc D. Gold nanoparticles: assembly, supramolecular chemistry, quantum-size-
related properties, and applications toward biology, catalysis, and nanotechnology. Chem Rev.
2004;104:293–346.
[13] Ferrando R, Jellinek J, Johnston R. Nanoalloys: from theory to applications of alloy clusters and
nanoparticles. Chem Rev. 2008;108:845–910.
[14] Rosi NL, Mirkin CA. Nanostructures in biodiagnostics. Chem Rev. 2005;105:1547–1562.
[15] Dreaden EC, Alkilany AM, Huang X, Murphy CJ, El-Sayed MA. The golden age: gold nanoparticles
for biomedicine. Chem Soc Rev. 2012;41:2740–2779.
[16] Kneipp K, Moskovits M, Kneipp H, editors. Surface-enhanced Raman scattering: physics and
applications. Topics in applied physics. Vol. 103. Berlin: Springer; 2006.
[17] Guerrini L, Graham D. Molecularly-mediated assemblies of plasmonic nanoparticles for surface-
enhanced Raman spectroscopy applications. Chem Soc Rev. 2012;41:7085–7107.
[18] Larmour IA, Graham D. Surface enhanced optical spectroscopies for bioanalysis. Analyst.
2011;136:3831–3853.
[19] Pieczonka NPW, Aroca RF. Single molecule analysis by surface-enhanced Raman scattering. Chem
Soc Rev. 2008;37:946–954.
[20] Kneipp K, Wang Y, Kneipp H, Perelman LT, Itzkan I, Dasari RR, Feld MS. Single molecule detection
using surface-enhanced Raman scattering (SERS). Phys Rev Lett. 1997;78:1667–1670.
[21] Nie S, Emory SR. Probing single molecules and single nanoparticles by surface-enhanced Raman
scattering. Science. 1997;275:1102–1106.
[22] Qian XM, Nie SM. Single-molecules and single-nanoparticle SERS: from fundamental mechanisms
to biomedical applications. Chem Soc Rev. 2008;37:912–920.
[23] Qian XM, Peng XH, Ansari DO, Yin-Goen Q, Chen GZ, Shin DM, Yang L, Young AN, Wang
MD, Nie SM. In vivo tumor targeting and spectroscopic detection with surface-enhanced Raman
nanoparticle tags. Nature Biotechnol. 2007;26:83–90.
[24] Sha MY, Xu H, Penn SG, Cromer R. SERS nanoparticles: A new optical detection modality for
cancer diagnostics. Nanomedicine (London). 2007;2:725–734.
[25] Ross CA. Patterned magnetic recording media. Annu Rev Mater Res. 2001;31:203–235.
[26] Wang JP. FePt magnetic nanoparticles and their assembly for future magnetic media. Proc IEEE.
2008;96:1847–1863.
[27] Moser A, Takano K, Margulies DT, Albrecht M, Sonobe Y, Ikeda Y, Sun S, Fullerton EE. Magnetic
recording: advancing into the future. J Phys D: Appl Phys. 2002;35:R157–R167.
[28] Terris BD, Thomson T. Nanofabricated and self-assembled magnetic structures as data storage
media. J Phys D: Appl Phys. 2005;38:R199–R222.
[29] Bell AT. The impact of nanoscience on heterogeneous catalysis. Science. 2003;299:1688–1691.
[30] Cheong S, Watt JD, Tilley RD. Shape control of platinum and palladium nanoparticles for catalysis.
Nanoscale. 2010;2:2045–2053.
[31] Haruta M, Daté M. Advances in the catalysis of Au nanoparticles. Appl Catal A. 2001;222:427–437.
[32] Astruc D, Lu F, Aranzaes JR. Nanoparticles as recyclable catalysts: the frontier between homoge-
neous and heterogeneous catalysis. Angew Chem Int Ed. 2005;44:7852–7872.
[33] Wieckowski A, Savinova ER, Vayenas CG, editors. Catalysis and electrocatalysis at nanoparticle
surfaces. New York: Marcel Dekker; 2003.
[34] Lewis LN. Chemical catalysis by colloids and clusters. Chem Rev. 1993;93:2693–2730.
[35] Sardar R, Funston AM, Mulvaney P, Murray RW. Gold nanoparticles: past, present, and future.
Langmuir. 2009;25:13840–13851.
Crystallography Reviews 289

[36] Zhou ZY, Tian N, Li JT, Broadwell I, Sun SG. Nanomaterials of high surface energy with exceptional
properties in catalysis and energy storage. Chem Soc Rev. 2011;40:4167–4185.
[37] Li Y, Somorjai GA. Nanoscale advances in catalysis and energy applications. Nano Lett.
2010;10:2289–2295.
[38] Zhang Q, Uchaker E, Candelaria SL, Cao G. Nanomaterials for energy conversion and storage. Chem
Soc Rev. 2013;42:3127–3171.
[39] Standridge SD, Schatz GC, Hupp JT. Distance dependence of plasmon-enhanced photocurrent in
dye-sensitized solar cells. J Am Chem Soc. 2009;131:8407–8409.
[40] Xu L, Ma W, Wang L, Xu C, Kuang H, Kotov NA. Nanoparticle assemblies: dimensional
transformation of nanomaterials and scalability. Chem Soc Rev. 2013;42:3114–3126.
[41] Wang Y, Herron N. Nanometer-sized semiconductor clusters: materials synthesis, quantum size
effects, and photophysical properties. J Phys Chem. 1991;95:525–532.
[42] Alivisatos AP. Semiconductor clusters, nanocrystals, and quantum dots. Science. 1996;271:933–937.
[43] Qu L, Peng X. Control of photoluminescence properties of CdSe nanocrystals in growth. J Am Chem
Soc. 2002;124:2049–2055.
[44] Park J, Lee KH, Galloway JF, Searson PC. Synthesis of cadmium selenide quantum dots from a non-
coordinating solvent: growth kinetics and particle size distribution. J Phys Chem C. 2008;112:17849–
17854.
[45] Tuinenga C, Jasinski J, Iwamoto T, Chikan V. In situ observation of heterogeneous growth of CdSe
quantum dots: effect of indium doping on the growth kinetics. ACS Nano. 2008;2:1411–1421.
[46] Skarzynski T. Collecting data in the home laboratory: evolution of X-ray sources, detectors and
working practices. Acta Crystallogr D. 2013;69:1283–1288.
[47] Materlik G, Tschentscher T editors. TESLA technical design report. Part V: The X-ray free electron
laser. Hamburg (Germany): HASYLAB; 2001.
[48] Silva NJO, Peunte-Orench I, Martins M, Trindade T, Millán A, Campo J, Palacio F. Neutron diffrac-
tion and magnetism of CoO antiferromagnetic nanoparticles. J Phys: Conf Ser. 2011;325:012020.
[49] Blanco-Gutierrez V, Climent-Pascual E, Torralvo-Fernandex MJ, Saez-Puche R, Fernandez-Diaz
MT. Neutron diffraction study and superparamagnetic behaviour of ZnFe2 O4 nanoparticles obtained
with different conditions. J Solid State Chem. 2011;184:1608–1613.
[50] Reinhard D, Hall BD, Ugarte D, Monot R. Size-independent fcc-to-icosahedral structural transi-
tion in unsupported silver clusters: an electron diffraction study of clusters produced by inert-gas
aggregation. Phys Rev B. 1997;55:7868–7881.
[51] Hall BD, Flüeli M, Monot R, Borel JP. Multiply twinned structures in unsupported ultrafine silver
particles observed by electron diffraction. Phys Rev B. 1991;43:3906–3917.
[52] Strunz P, Mukherji D, Pigozzi G, Gilles R, Geue T, Pranzas K. Characterization of core–shell
nanoparticles by small angle neutron scattering. Appl Phys A. 2007;88:277–284.
[53] Von White G, Mohammed FS, Kitchens CL. Small-angle neutron scattering investigation of
gold nanoparticle clustering and ligand structure under antisolvent conditions. J Phys Chem C.
2011;115:18397–18405.
[54] Moglianetti M, Ong QK, Requera J, Harkness KM, Mameli M, Radulsecu A, Kohlbrecher J, Jud C,
Svergun DI, Stellacci F. Scanning tunnelling microscopy and small angle neutron scattering study of
mixed monolayer protected gold nanoparticles in organic solvents. Chem Sci. 2014;5:1232–1240.
[55] Calvin S, Luo SX, Caragianis-Broadbridge C, McGuinness JK, Anderson E, Lehman A, Wee KH,
Morrison SA, Kurihara LK. Comparison of extended X-ray absorption fine structure and Scherrer
analysis of X-ray diffraction as methods for determining mean sizes of polydisperse nanoparticles.
Appl Phys Lett. 2005;87:233102.
[56] Belyakova OA, Zubavichus YV, Neretin IS, Golub AS, Novikov YN, Mednikov EG, Vargaftik MN,
Moiseev II, Slovokhotov YL. Atomic structure of nanomaterials: combined X-ray diffraction and
EXAFS studies. J Alloys Compounds. 2004;382:46–53.
[57] Rockenberger J, Tröger L, Kornowski A, Vossmeyer T, Eychmüller A, Feldhaus J, Weller H. EXAFS
studies on the size dependence of structural and dynamic properties of CdS nanoparticles. J Phys
Chem B. 1997;101:2691–2701.
[58] Frenkel AI. Solving the structure of nanoparticles by multiple-scattering EXAFS analysis. J
Synchrotron Radiat. 1999;6:293–295.
[59] Frenkel A. Solving the 3D structure of metal nanoparticles. Z Kristallogr. 2007;222:605–611.
[60] Knecht MR, Weir MG, Frenkel AI, Crooks RM. Structural rearrangement of bimetallic alloy
PdAu nanoparticles within dendrimer templates to yield core/shell configurations. Chem Mater.
2008;20:1019–1028.
[61] Warren BE. X-ray diffraction. Reading (MA): Dover; 1969.
290 B. Ingham

[62] Cullity BD, Stock SR. Elements of X-ray diffraction. New York: Prentice Hall; 2001.
[63] The International Centre for Diffraction Data [Internet]. Available from: http://www.icdd.com/
[64] The Cambridge Crystallographic Data Centre [Internet]. Available from: http://www.ccdc.cam.ac.uk/
[65] FIZ Karlsruhe [Internet]. Available from: http://www.fiz-karlsruhe.de
[66] NIST Standard Reference Database 84 [Internet]. Available from: http://www.nist.gov/srd/nist84.cfm
[67] Taylor JC, Hinczak I. Rietveld made easy: a practical guide to the understanding of the method and
successful phase quantifications. Canberra: Sietronics Pty. Ltd.; 2001.
[68] Bragg WL, Williams EJ. The effect of thermal agitation on atomic arrangement in alloys. Proc R Soc
Lond A. 1934;145:699–730.
[69] Scardi P, Leoni M, Beyerlein KR. On the modelling of the powder pattern from a nanocrystalline
material. Z Kristallogr. 2011;226:924–933.
[70] Niederdraenk F, Seufert K, Luczak P, Kulkarni SK, Chory C, Neder RB, Kumpf C. Structure of small
II–VI semiconductor nanoparticles: a new approach based on powder diffraction. Phys Status Solidi
C. 2007;4:3234–3243.
[71] Hall BD, Zanchet D, Ugarte D. Estimating nanoparticle size from diffraction measurements. J Appl
Crystallogr. 2000;33:1335–1341.
[72] Solliard C, Flueli M. Surface stress and size effect on the lattice parameter in small particles of gold
and palladium. Surf Sci. 1985;156:487–494.
[73] Lamber R, Wetjen S, Jaeger NI. Size dependence of the lattice parameter of small palladium particles.
Phys Rev B. 1995;51:10968–10971.
[74] Qi W, Wang M, Su Y. Size effect on the lattice parameters of nanoparticles. J Mater Sci Lett.
2002;21:877–878.
[75] Qi W, Wang M. Size and shape dependent lattice parameters of metallic nanoparticles. J Nanoparticle
Res. 2005;7:51–57.
[76] Langford JI, Louër D, Scardi P. Effect of a crystallite size distribution on X-ray diffraction line
profiles and whole-powder-pattern fitting. J Appl Crystallogr. 2000;33:964–974.
[77] Williamson GK, Hall WH. X-ray line broadening from filed aluminium and wolfram. Acta
Metallurgica. 1953;1:22–31.
[78] Kandemir T, Kasatkin I, Girgsdies F, Zander S, Kühl S, Tovar M, Schlögl R, Behrens M. Microstruc-
tural and defect analysis of metal nanoparticles in functional catalysts by diffraction and electron
microscopy: the Cu/ZnO catalyst for methanol synthesis. Top Catal. 2014;57:188–206.
[79] Zhou XD, Huebner W. Size-induced lattice relaxation in CeO2 nanoparticles. Appl Phys Lett.
2001;79:3512–3514.
[80] Singla G, Singh K, Pandey OP. Williamson–Hall study on synthesised nanocrystalline tungsten
carbide (WC). Appl Phys A. 2013;113:237–242.
[81] Venkateswarlu K, Chandra Bose A, Rameshbabu N. X-ray peak broadening studies of nanocrys-
talline hydroxyapatite by Williamson–Hall analysis. Physica B. 2010;405:4256–4261.
[82] Warren BE, Averbach BL. The effect of cold-work distortion on X-ray patterns. J Appl Phys.
1950;21:595–599.
[83] Warren BE, Averbach BL. The separation of cold-work distortion and particle size broadening in
X-ray patterns. J Appl Phys. 1952;23:497.
[84] Ungár T, Gubicza J, Ribárik G, Borbély A. Crystallite size distribution and dislocation structure
determined by diffraction profile analysis: principles and practical application to cubic and hexagonal
crystals. J Appl Crystallogr. 2001;34:298–310.
[85] Balzar D, Audebrand N, Daymond MR, Fitch A, Hewat A, Langford JI, Le Bail A, Louër D, Masson
O, McCowan CN, Popa NC, Stephens PW, Toby BH. Size-strain line-broadening analysis of the
ceria round-robin sample. J Appl Crystallogr. 2004;37:911–924.
[86] Kaszkur Z, Mierzwa B, Pielaszek J. Ab initio test of the Warren–Averbach analysis on model
palladium nanocrystals. J Appl Crystallogr. 2005;38:266–273.
[87] Ruzicka JY, Abu Bakar F, Thomsen L, Cowie BC, McNicoll C, Kemmitt T, Brand HEA, Ingham
B, Andersson GG, Golovko VB. XPS and NEXAFS study of fluorine modified TiO2 nano-ovoids
reveals dependence of Ti3+ surface population on the modifying agent. RSC Adv. 2014;4:20649–
20658.
[88] Popa NC, Balzar D. Size-broadening anisotropy in whole powder pattern fitting. Application to zinc
oxide and interpretation of the apparent crystallites in terms of physical models. J Appl Crystallogr.
2008;41:615–627.
[89] Ingham B, Lim TH, Dotzler CJ, Henning A, Toney MF, Tilley RD. How nanoparticles coalesce: an
in situ study of Au nanoparticle aggregation and grain growth. Chem Mater. 2011;23:3312–3317.
Crystallography Reviews 291

[90] Behrens M, Studt F, Kasatkin I, Kühl S, Hävecker M, Abild-Pedersen F, Zander S, Girgsdies F, Kurr
P, Kniep BL, Tovar M, Fischer RW, Nørskov JK, Schlögl R. The active site of methanol synthesis
over Cu/ZnO/Al2 O3 industrial catalysts. Science. 2012;336:893–897.
[91] LaGrow AP, Ingham B, Cheong S, Williams GVM, Dotzler C, Toney MF, Jefferson DA, Corbos EC,
Bishop PT, Cookson J, Tilley RD. Synthesis, alignment, and magnetic properties of monodisperse
nickel nanocubes. J Am Chem Soc. 2012;134:855–858.
[92] Zeng H, Rice PM, Wang SX, Sun S. Shape-controlled synthesis and shape-induced texture of
MnFe2 O4 nanoparticles. J Am Chem Soc. 2004;126:11458–11459.
[93] Liu X, Zhao L, Shen H, Xu H, Lu L. Ordered gold nanoparticle arrays as surface-enhanced Raman
spectroscopy substrates for label-free detection of nitroexplosives. Talanta. 2011;83:1023–1029.
[94] Swami A, Kumar A, Selvakannan P, Mandal S, Pasricha R, Sastry M. Highly oriented gold nanorib-
bons by the reduction of aqueous chloroaurate ions by hexadecylaniline Langmuir monolayers. Chem
Mater. 2003;15:17–19.
[95] Kurszynska M, Borchert H, Parisi J, Kolny-Olesiak J. Synthesis and shape control of CuInS2
nanoparticles. J Am Chem Soc. 2010;132:15976–15986.
[96] Ingham B, Illy BN, Toney MF, Howdyshell ML, Ryan MP. In situ synchrotron X-ray diffraction
experiments on electrochemically deposited ZnO nanostructures. J Phys Chem C. 2008;112:14863–
14866.
[97] Park WI, Yi GU, Kim M, Pennycook SJ. Zno nanoneedles grown vertically on Si substrates by
non-catalytic vapour-phase epitaxy. Adv Mater. 2002;14:1841–1843.
[98] Wu W, Hu G, Cui S, Zhou Y, Wu H. Epitaxy of vertical ZnO nanorod arrays on highly (001)-oriented
ZnO seed monolayer by a hydrothermal route. Cryst Growth Design. 2008;8:4014–4020.
[99] Huang Y, Okumura H, Hadjipanayis GC, Weller D. Perpendicularly oriented FePt nanoparticles
sputtered on heated substrates. J Magn Magn Mater. 2002;242–245:317–320.
[100] Guo M, Diao P, Cai S. Hydrothermal growth of well-aligned ZnO nanorod arrays: dependence of
morphology and alignment ordering upon preparing conditions. J Solid State Chem. 2005;178:1864–
1873.
[101] Ahniyaz A, Sakamoto Y, Bergström L. Magnetic field-induced assembly of oriented superlattices
from maghemite nanocubes. Proc Natl Acad Sci USA. 2007;104:17570–17574.
[102] Wan Q, Wei M, Zhi D, MacManus-Driscoll JL, Blamire MG. Epitaxial growth of vertically aligned
and branched single-crystalline tin-doped indium oxide nanowire arrays. Adv Mater. 2006;18:234–
238.
[103] Kim HM, Kim DS, Kim DY, Kang TW, Cho YH, Chung KS. Growth and characterisation of single-
crystal GaN nanorods by hydride vapour phase epitaxy. Appl Phys Lett. 2002;81:2193–2195.
[104] Yu ZQ, Wang CM, Engelhard MH, Nachimuthu P, McCready DE, Lyubinetsky IV, Thevuthasan S.
Epitaxial growth and microstructure of Cu2 O nanoparticle/thin films on SrTiO3 (100). Nanotechnol-
ogy. 2007;18:115601.
[105] Tsybulya SV, Yatsenko DA. X-ray diffraction analysis of ultradisperse systems: The Debye formula.
J Struct Chem. 2012;53:S150–S165.
[106] Beyerlein KR. A review of Debye function analysis. Powder Diffraction. 2013;28:S2–S10.
[107] Bazin DC, Sayers DA, Rehr JJ. Comparison between X-ray absorption spectroscopy, anomalous
wide angle X-ray scattering, anomalous small angle X-ray scattering, and diffraction anomalous fine
structure techniques applied to nanometre-scale metallic clusters. J Phys Chem B. 1997;101:11040–
11050.
[108] Debye P. Scattering of X-rays. Ann Phys. 1915;351:809–823. German.
[109] Baranova EA, Le Page Y, Ilin D, Bock C, MacDougall B, Mercier PHJ. Size and composition for 1–5
nm Ø PtRu alloy nanoparticles from Cu Kα X-ray patterns. J Alloys Compounds. 2009;471:387–394.
[110] Zanchet D, Hall BD, Ugarte D. Structure population in thiol-passivated gold nanoparticles. J Phys
Chem B. 2000;104:11013–11018.
[111] Bawendi MG, Kortan AR, Steigerwald ML, Brus LE. X-ray structural characterisation of larger CdSe
semiconductor clusters. J Chem Phys. 1989;91:7282–7290.
[112] Masadeh AS, Božin ES, Farrow CL, Paglia G, Juhás P, Billinge SJL, Karkamkar A, Kanatzidis MG.
Quantitative size-dependent structure and strain determination of CdSe nanoparticles using atomic
pair distribution function analysis. Phys Rev B. 2007;76:115413.
[113] Hall BD, Ugarte D, Reinhard D, Monot R. Calculations of the dynamic Debye–Scherrer diffraction
patterns for small metal particles. J Chem Phys. 1995;103:2384–2394.
[114] Beyerlein KR, Snyder RL, Li M, Scardi P. Application of the Debye function to systems of
crystallites. Philos Magazine. 2010;90:3891–3905.
292 B. Ingham

[115] Masson O, Boulle A, Guinebretière R, Lecomte A, Dauger A. On the use of one-dimensional position
sensitive detector for X-ray diffraction reciprocal space mapping: data quality and limitations. Rev
Sci Instrum. 2005;76:063912.
[116] Schmitt B, Brönnimann C, Eikenberry EF, Gozzo F, Hörmann C, Horisberger R, Patterson B. Mythen
detector system. Nucl Instrum Meth Phys Res A. 2003;501:267–272.
[117] Bergamaschi A, Cervellino A, Dinapoli R, Gozzo F, Henrich B, Johnson I, Kraft P, Mozzanica A,
Schmitt B, Shi X. The MYTHEN detector for X-ray powder diffraction experiments at the Swiss
light source. J Synchrotron Radiat. 2010;17:653–668.
[118] Hülsen G, Brönnimann C, Eikenberry EF, Wagner A. Protein crystallography with a novel large-area
pixel detector. J Appl. Crystallogr. 2006;39:550–557.
[119] Brönnimann C, Eikenberry EF, Henrich B, Horisberger R, Hülsen G, Pohl E, Schmitt B, Schulze-
Briese C, Suzuki M, Tomizaki T, Toyokawa H, Wagner A. The PILATUS 1M detector. J Synchrotron
Radiat. 2006;13:120–130.
[120] Henrich B, Bergamaschi A, Brönnimann C, Dinapoli R, Eikenberry EF, Johnson I, Kobas M, Kraft P,
Mozzanica A, Schmitt B. PILATUS: A single photon counting pixel detector for X-ray applications.
Nucl Instrum Methods Phys Res A. 2009;607:247–249.
[121] Kraft P, Bergamaschi A, Brönnimann C, Dinapoli R, Eikenberry EF, Henrich B, Johnson I, Mozzan-
ica A, Schlepütz CM, Willmott PR, Schmitt B. Performance of single-photon-counting PILATUS
detector modules. J Synchrotron Radiat. 2009;16:368–375.
[122] Kraft P, Bergamaschi A, Brönnimann C, Dinapoli R, Eikenberry EF, Graafsma H, Henrich B,
Johnson I, Kobas M, Mozzanica A, Schlepütz CM, Schmitt B. Characterization and calibration of
PILATUS detectors. IEEE Trans Nucl Sci. 2009;56:758–764.
[123] Trueb P, Sobott BA, Schnyder R, Loeliger T, Schneebeli M, Kobas M, Rassool RP, Peake DJ,
Brönnimann C. Improved count rate corrections for highest data quality with PILATUS detectors. J
Synchrotron Radiat. 2012;19:347–351.
[124] Ayyub P, Chandra R, Taneja P, Sharma AK, Pinto R. Synthesis of nanocrystalline material by
sputtering and laser ablation at low temperatures. Appl Phys A. 2001;73:67–73.
[125] Reichel R, Partridge JG, Dunbar ADF, Brown SA, Caughley O, Ayesh A. Construction and
application of a UHV compatible cluster deposition system. J Nanoparticle Res. 2006;8:405–416.
[126] Petkov V, Peng Y, Williams G, Huang B, Tomalia D, Ren Y. Structure of gold nanoparti-
cles suspended in water studied by X-ray diffraction and computer simulations. Phys Rev B.
2005;72:195402.
[127] Baker JL, Jimison LH, Mannsfeld S, Volkman S, Yin S, Subramanian V, Salleo A, Alivisatos AP,
Toney MF. Quantification of thin film crystallographic orientation using X-ray diffraction with an
area detector. Langmuir. 2010;26:9146–9151.
[128] Wang C, Daimon H, Lee Y, Kim J, Sun S. Synthesis of monodisperse Pt nanocubes and their
enhanced catalysis for oxygen reduction. J Am Chem Soc. 2007;129:6974–6975.
[129] Shevchenko EV, Talapin DV, Murray CB, O’Brien S. Structural characterisation of self-assembled
multifunctional binary nanoparticle superlattices. J Am Chem Soc. 2006;128:3620–3637.
[130] Guo H, Chen Y, Ping H, Jin J, Peng DL. Facile synthesis of Cu and Cu@Cu–Ni nanocubes
and nanowires in hydrophobic solution in the presence of nickel and chloride ions. Nanoscale.
2013;5:2394–2402.
[131] Yang HJ, He SY, Chen HL, Tuan HY. Monodisperse copper nanocubes: synthesis, self-assembly,
and large-area dense-packed films. Chem Mater. 2014;26:1785–1793.
[132] Im SH, Lee YT, Wiley B, Xia Y. Large-scale synthesis of silver nanocubes: The role of HCl in
promoting cube perfection and monodispersity. Angew Chem Int Ed. 2005;44:2154–2157.
[133] Siekkinen AR, McLellan JM, Chen J, Xia Y. Rapid synthesis of small silver nanocubes by mediating
polyol reduction with a trace amount of sodium sulphide or sodium hydrosulfide. Chem Phys Lett.
2006;432:491–496.
[134] Yu D, Yam VWW. Controlled synthesis of monodisperse silver nanocubes in water. J Am Chem Soc.
2004;126:13200–13201.
[135] Xu D, Liu Z, Yang H, Liu Q, Zhang J, Fang J, Zou S, Sun K. Solution-based evolution and enhanced
methanol oxidation activity of monodisperse platinum–copper nanocubes. Angew Chem Int Ed.
2009;48:4217–4221.
[136] Choi SI, Choi R, Han SW, Park JT. Synthesis and characterisation of Pt9 Co nanocubes with high
activity for oxygen reduction. Chem Commun. 2010;46:4950–4952.
[137] Zhang J, Fang J. A general strategy for preparation of Pt 3d-transition metal (Co, Fe, Ni) nanocubes.
J Am Chem Soc. 2009;131:18543–18547.
[138] Toney MF, Huang TC, Brennan S, Rek Z. X-ray depth profiling of iron oxide thin films. J Mater Res.
1988;3:351–356.
Crystallography Reviews 293

[139] Anders S, Toney MF, Thomson T, Farrow RFC, Thiele JU, Terris BD, Sun S, Murray CB. X-
ray absorption and diffraction studies of thin polymer/FePt nanoparticle assemblies. J Appl Phys.
2003;93:6299–6304.
[140] Sun S, Anders S, Thomson T, Baglin JEE, Toney MF, Hamann HF, Murray CB, Terris BD. Controlled
synthesis and assembly of FePt nanoparticles. J Phys Chem B. 2003;107:5419–5425.
[141] Hsieh YC, Zhang Y, Su D, Volkov V, Si R, Wu L, Zhu Y, An W, Liu P, He P, Ye S, Adzic RR, Wang
JX. Ordered bilayer ruthenium-platinum core–shell nanoparticles as carbon monoxide-tolerant fuel
cell catalysts. Nature Commun. 2013;4:2466.
[142] LaGrow AP, Cheong S, Watt J, Ingham B, Toney MF, Jefferson DA, Tilley RD. Can polymorphism
be used to form branched metal nanostructures? Adv Mater. 2013;25:1552–1556.
[143] Huang X, Li H, Li S, Wu S, Boey F, Ma J, Zhang H. Synthesis of gold square-like plates from
ultrathin gold square sheets: The evolution of structure phase and shape. Angew Chem Int Ed.
2011;50:12245–12248.
[144] Huang X, Li S, Wu Q, Huang Y, Boey F, Gan CL, Zhang H. Graphene oxide-templated synthe-
sis of ultrathin or tadpole-shaped Au nanowires with alternating hcp and fcc domains. Adv Mater.
2012;24:979–983.
[145] Koga K, Ikeshoji T, Sugawara K. Size- and temperature-dependent structural transitions in gold
nanoparticles. Phys Rev Lett. 2004;92:115507.
[146] Hendy SC, Hall BD. Molecular-dynamics simulations of lead clusters. Phys Rev B. 2001;64:085425.
[147] Cervellino A, Giannini C, Guagliardi A, Zanchet D. Quantitative analysis of gold nanoparticles from
synchrotron data by means of least-squares techniques. Eur Phys J B. 2004;41:485–493.
[148] Cervellino A, Giannini C, Guagliardi A. Determination of nanoparticle structure type, size and strain
distribution from X-ray data for monatomic f.c.c.-derived non-crystallographic nanoclusters. J Appl
Crystallogr. 2003;36:1148–1158.
[149] Nayral C, Viala E, Fau P, Senocq F, Jumas JC, Maisonnat A, Chaudret B. Synthesis of tin and tin
oxide nanoparticles of low size dispersity for application in gas sensing. Chem Eur J. 2000;6:4082–
4090.
[150] Stevens KJ, Ingham B, Toney MF, Brown SA, Partridge J, Ayesh A, Natali F. Structure of oxidized
bismuth nanoclusters. Acta Crystallogr B. 2007;63:569–576.
[151] Lei P, Boies AM, Calder S, Girshick SL. Thermal plasma synthesis of superparamagnetic iron oxide
nanoparticles. Plasma Chem Plasma Process. 2012;32:519–531.
[152] Teng X, Yang H. Effects of surfactants and synthetic conditions on the sizes and self-assembly of
monodisperse iron oxide nanoparticles. J Mater Chem. 2004;14:774–779.
[153] Hou Y, Xu Z, Sun S. Controlled synthesis and chemical conversions of FeO nanoparticles. Angew
Chem Int Ed. 2007;46:6329–6332.
[154] Teleki A, Pratsinis SE. Blue nano titania made in diffusion flames. Phys Chem Chem Phys.
2009;11:3742–3747.
[155] Mahadevan S, Behera SP, Gnanaprakash G, Jayakumar T, Philip J, Rao BPC. Size distribution of
magnetic iron oxide nanoparticles using Warren–Averbach XRD analysis. J Phys Chem Solids.
2012;73:867–872.
[156] Tsunekawa S, Ishikawa K, Li ZQ, Kawazoe Y, Kasuya A. Origin of anomalous lattice expansion in
oxide nanoparticles. Phys Rev Lett. 2000;85:3440–3443.
[157] Zhang F, Chan SW, Spanier JE, Apak E, Jin Q, Robinson RD, Herman IP. Cerium oxide
nanoparticles: size-selective formation and structure analysis. Appl Phys Lett. 2002;80:127–129.
[158] Özgür Ü, Alivov YA, Liu C, Teke A, Reshchikov MA, Doğan S, Avrutin V, Cho SJ, Morkoç H. A
comprehensive review of ZnO materials and devices. J Appl Phys. 2005;98:041301.
[159] Wang ZL. Zinc oxide nanostructures: growth, properties and applications. J Phys Condens Matter.
2004;16:R829–R858.
[160] Weintraub B, Zhou Z, Li Y, Deng Y. Solution synthesis of one-dimensional ZnO nanomaterials and
their applications. Nanoscale. 2010;2:1573–1587.
[161] Willander M, Nur O, Zhao QX, Yang LL, Lorenz M, Cao BQ, Zúñiga Pérez J, Czekalla C, Zim-
mermann G, Grundmann M, Bakin A, Behrends A, Al-Suleiman M, El-Shaer A, Che Mofor A,
Postels B, Waag A, Boukos N, Travlos A, Kwack HS, Guinard J, Le Si Dang D. Zinc oxide nanorod
based photonic devices: recent progress in growth, light emitting diodes and lasers. Nanotechnology.
2009;20:332001.
[162] Janotti A, Van de Walle CG. Fundamentals of zinc oxide as a semiconductor. Rep Prog Phys.
2009;72:126501.
[163] Djurisić AB, Leung YH. Optical properties of ZnO nanostructures. Small. 2006;2:944–961.
[164] Chory C, Neder RB, Korsunskiy VI, Niederdraenk F, Kumpf C, Umbach E, Schumm M, Lentze M,
Geurts J, Astakhov G, Ossau W, Müller G. Influence of liquid-phase synthesis parameters on particle
294 B. Ingham

sizes and structural properties of nanocrystalline ZnO powders. Phys Stat Solidi C. 2007;4:3260–
3269.
[165] Zhang YL, Yang Y, Zhao JH, Tan RQ, Cui P, Song WJ. Preparation of ZnO nanoparticles by a
surfactant-assisted complex sol–gel method using zinc nitrate. J Sol–Gel Sci Technol. 2009;51:198–
203.
[166] Hynes AP, DOremus RH, Siegel RW. Sintering and characterisation of nanophase zinc oxide. J Am
Ceram Soc. 2002;85:1979–1987.
[167] Ingham B, Linklater R, Kemmitt T. Grain growth kinetics of ZnO:Al nanocrystalline powders during
calcination from sol–gels. J Phys Chem C. 2011;115:21034–21040.
[168] Shukla S, Seal S, Vij R, Bandyopadhyay S. Reduced activation energy for grain growth in
nanocrystalline yttria-stabilized zirconia. Nano Lett. 2003;3:397–401.
[169] Shek CH, Lai JKL, Lin GM. Grain growth in nanocrystalline SnO2 prepared by sol–gel route.
Nanostruct Mater. 1999;11:887–893.
[170] Lai JKL, Shek CH, Lin GM. Grain growth kinetics of nanocrystalline SnO2 for long-term isothermal
annealing. Scripta Mater. 2003;49:441–446.
[171] Kuo CW, Shen YH, Hung IM, Wen SB, Lee HE, Wang MC. Effect of Y2 O3 addition on the crys-
tal growth and sintering behaviour of YSZ nanopowders prepared by a sol–gel process. J Alloys
Compounds. 2009;472:186–193.
[172] Niederdraenk F, Seufert K, Stahl A, Bhalerao-Panajkar RS, Marathe S, Kulkarni SK, Neder
RB, Kumpf C. Ensemble modelling of very small ZnO nanoparticles. Phys Chem Chem Phys.
2011;13:498–505.
[173] Zhu Y, Zhang X, Li R, Li Q. Planar-defect-rich zinc oxide nanoparticles assembled on carbon
nanotube films as ultraviolet emitters and photocatalysts. Sci Rep. 2014;4:4728.
[174] Nakagawa Y, Kageyama H, Oaki Y, Imai H. Direction control of oriented self-assembly for
1D, 2D, and 3D microarrays of anisotropic rectangular nanoblocks. J Am Chem Soc. 2014;136:
3716–3719.
[175] Banerjee R, Jayakrishnan R, Ayyub P. Effect of the size-induced structural transformation on the
band gap in CdS nanoparticles. J Phys Condens Matter. 2000;12:10647–10654.
[176] Patra S, Satpati B, Pradhan SK. Microstructure characterisation of mechanically synthesised ZnS
quantum dots. J Appl Phys. 2009;106:034313.
[177] Kim YH, Jun YW, Jun BH, Lee SM, Cheon J. Sterically induced shape and crystalline phase control
of GaP nanocrystals. J Am Chem Soc. 2002;124:13656–13657.
[178] Manna L, Scher EC, Alivisatos AP. Synthesis of soluble and processable rod-, arrow-, teardrop- and
tetrapod-shaped CdSe nanocrystals. J Am Chem Soc. 2000;122:12700–12706.
[179] Peng X, Manna L, Yang W, Wickham J, Scher E, Kadavanich A, Alivisatos AP. Shape control of
CdSe nanocrystals. Nature. 2000;404:59–61.
[180] Baranov AV, Rakovich YP, Donegan JF, Perova TS, Moore RA, Talapin DV, Rogach AL, Masumoto
Y, Nabiev I. Effect of ZnS shell thickness on the phonon spectra in CdSe quantum dots. Phys Rev B.
2003;68:165306.
[181] Dabbousi BO, Rodriguez-Viejo J, Mikulec FV, Heine JR, Mattoussi H, Ober R, Jensen KF, Bawendi
MG. (Cdse)ZnS core–shell quantum dots: synthesis and characterisation of a size series of highly
luminescent nanocrystallites. J Phys Chem B. 1997;101:9463–9475.
[182] Peng X, Schlamp MC, Kadavanich AV, Alivisatos AP. Epitaxial growth of highly luminescent
CdSe/CdS core/shell nanocrystals with photostability and electronic accessibility. J Am Chem Soc.
1997;119:7019–7029.
[183] Lee YJ, Kim TG, Sung YM. Lattice distortion and luminescence of CdSe/ZnSe nanocrystals.
Nanotechnology. 2006;17:3539–3542.
[184] Smith AM, Mohs AM, Nie S. Tuning the optical and electronic properties of colloidal nanocrystals
by lattice strain. Nature Nanotechnol. 2008;4:56–63.
[185] Kumpf C, Neder RB, Niederdraenk F, Luczak P, Stahl A, Cheuermann M, Joshi S, Kulkarni SK,
Barglik-Chory C, Heske C, Umbach E. Structure determination of CdS and ZnS nanoparticles: direct
modelling of synchrotron-radiation diffraction data. J Chem Phys. 2005;123:224707.
[186] Murray CB, Norris DJ, Bawendi MG. Synthesis and characterization of nearly monodisperse CdE
(E = S, Se, Te) semiconductor nanocrystallites. J Am Chem Soc. 1993;115:8706–8715.
[187] Moreau LM, Ha DH, Zhang H, Hovden R, Muller DA, Robinson RD. Defining crys-
talline/amorphous phases of nanoparticles through X-ray absorption spectroscopy and X-ray
diffraction: The case of nickel phosphide. Chem Mater. 2013;25:2394–2403.
[188] Ingham B. In situ, real-time synchrotron X-ray scattering. In: Bauwens CM, editor. X-ray scattering.
New York: Nova; 2012; p. 55–86.
Crystallography Reviews 295

[189] Cheong S, Watt J, Ingham B, Toney MF, Tilley RD. In situ and ex situ studies of platinum
nanocrystals: growth and evolution in solution. J Am Chem Soc. 2009;131:14590–14595.
[190] Watt J, Cheong S, Toney MF, Ingham B, Cookson J, Bishop PT, Tilley RD. Ultrafast growth of highly
branched palladium nanostructures for catalysis. ACS Nano. 2010;4:396–402.
[191] Wu Q, Duchstein LDL, Chiarello GL, Christensen JM, Damsgaard CD, Elkjær CF, Wagner JB,
Temel B, Grunwaldt JD, Jensen AD. In situ observation of Cu-Ni alloy nanoparticle formation by
X-ray diffraction, X-ray absorption spectroscopy, and transmission electron microscopy: influence
of Cu/Ni ratio. ChemCatChem. 2014;6:301–310.
[192] Oezaslan M, Hasché F, Strasser P. In situ observation of bimetallic alloy nanoparticle formation and
growth using high-temperature XRD. Chem Mater. 2011;23:2159–2165.
[193] Mishra YK, Avasthi DK, Kulriya PK, Singh F, Kabiraj D, Tripathi A, Pivin JC, Bayer IS, Biswas
A. Controlled growth of gold nanoparticles induced by ion irradiation: an in situ X-ray diffraction
study. Appl Phys Lett. 2007;90:073110.
[194] Peng S, Okasinski JS, Almer JD, Ren Y, Wang L, Yang W, Sun Y. Real-time probing of the synthesis
of colloidal silver nanocubes with time-resolved high-energy synchrotron X-ray diffraction. J Phys
Chem C. 2012;116:11842–11847.
[195] Sun J, Ma D, Zhang H, Liu X, Han X, Bao X, Weinberg G, Pfänder N, Su D. Toward
monodispersed silver nanoparticles with unusual thermal stability. J Am Chem Soc. 2006;128:
15756–15764.
[196] Vogel W, Bradley J, Vollmer O, Abraham I. Transition from five-fold symmetric to twinned FCC
gold particles by thermally induced growth. J Phys Chem B. 1998;102:10853–10859.
[197] Nguyen HL, Howard LEM, Giblin SR, Tanner BK, Terry I, Hughes AK, Ross IM, Serres A, Bür-
ckstümmer H, Evans JSO. Synthesis of monodispersed fcc and fct FePt/FePd nanoparticles by
microwave irradiation. J Mater Chem. 2005;15:5136–5143.
[198] Champion Y, Bernard F, Guigue-Millot N, Perriat P. Sintering of copper nanopowders under
hydrogen:an in situ X-ray diffraction analysis. Mater Sci Eng A. 2003;360:258–263.
[199] Ingham B, Hendy SC, Fong DD, Fuoss PH, Eastman JA, Lassesson A, Tee KC, Convers PY,
Brown SA, Ryan MP, Toney MF. Synchrotron X-ray diffraction measurements of strain in metallic
nanoparticles with oxide shells. J Phys D: Appl Phys. 2010;43:075301.
[200] Champion Y, Bernard F, Millot N, Perriat P. Surface adsorption effects on the lattice expansion of
copper nanocrystals. Appl Phys Lett. 2005;86:231914.
[201] Hu J, Cai W, Li C, Gan Y, Chen L. In situ X-ray diffraction study of the thermal expansion of silver
nanoparticles in ambient air and vacuum. Appl Phys Lett. 2005;86:151915.
[202] Hendy SC. A thermodynamic model for the melting of supported metal nanoparticles. Nanotechnol-
ogy. 2007;18:175703.
[203] Buffat P, Borel JP. Size effect on the melting temperature of gold particles. Phys Rev A.
1976;13:2287–2298.
[204] Depero LE, Bontempi E, Sangaletti L, Pagliara S. Melting of nanostructured Sn probed by in-situ
X-ray diffraction. J Chem Phys. 2003;118:1400–1403.
[205] Peters KF, Chung YW, Cohen JB. Surface melting on small particles. Appl Phys Lett. 1997;71:2391–
2393.
[206] Peters KF, Cohen JB, Chung YW. Melting of Pb nanocrystals. Phys Rev B. 1998;57:13430–13438.
[207] Sheng HW, Lu K, Ma E. Melting of embedded Pb nanoparticles monitored using high-temperature
in situ XRD. Nanostructured Mater. 1998;10:865–873.
[208] Gu QF, Krauss G, Steurer W, Gramm F, Cervellino A. Unexpected high stiffness of Ag and Au
nanoparticles. Phys Rev Lett. 2008;100:045502.
[209] Martin CD, Antao SM, Chupas PJ, Lee PL, Shastri SD, Parise JB. Quantitative high-pressure pair
distribution function analysis of nanocrystalline gold. Appl Phys Lett. 2005;86:061910.
[210] Syassen K, Holzapfel WB. Isothermal compression of Al and Ag to 120 kbar. J Appl Phys.
1978;49:4427–4430.
[211] Heinz DL, Jeanloz R. The equation of state of the gold calibration standard. J Appl. Phys.
1984;55:885–893.
[212] Gilbert B, Huang F, Zhang H, Waychunas GA, Banfield JF. Nanoparticles: strained and stiff. Science.
2004;305:651–654.
[213] Araujo LL, Kluth P, Azevedo GDM, Ridgway MC. Vibrational properties of Ge nanocrystals
determined by EXAFS. Phys Rev B. 2006;74:184102.
[214] Ouyang G, Zhu WG, Sun CQ, Zhu ZM, Liao SZ. Atomistic origin of lattice strain on stiffness of
nanoparticles. Phys Chem Chem Phys. 2010;12:1543–1549.
296 B. Ingham

[215] Cuenot S, Frétigny C, Demoustier-Champagne S, Nysten B. Surface tension effect on the mechanical
properties of nanomaterials measured by atomic force microscopy. Phys Rev B. 2004;69:165410.
[216] Suleiman M, Faupel J, Borchers C, Krebs HU, Kirchheim R, Pundt A. Hydrogen absorption
behaviour in nanometer sized palladium samples stabilised in soft and hard matrix. J Alloys
Compounds. 2005;404–406:523–528.
[217] Pundt A, Suleiman M, Bähtz C, Reetz MT, Kirchheim R, Jisrawi NM. Hydrogen and Pd-clusters.
Mater Sci Eng B. 2004;108:19–23.
[218] Suleiman M, Jisrawi NM, Dankert O, Reetz MT, Bähtz C, Kirchheim R, Pundt A. Phase transi-
tion and lattice expansion during hydrogen loading of nanometer sized palladium clusters. J Alloys
Compounds. 2003;356–357:644–648.
[219] Ingham B, Toney MF, Hendy SC, Cox T, Fong DD, Eastman JA, Fuoss PH, Stevens KJ, Lassesson
A, Brown SA, Ryan MP. Particle size effect of hydrogen-induced lattice expansion of palladium
nanoclusters. Phys Rev B. 2008;78:245408.
[220] Baranowski B, Majchrzak S, Flanagan TB. The volume increase of fcc metals and alloys due to
interstitial hydrogen over a wide range of hydrogen contents. J Phys F Metal Phys. 1971;1:258–261.
[221] van Lith J, Lassesson A, Brown SA, Schulze M, Partridge JG, Ayesh A. A hydrogen sensor based on
tunnelling between palladium clusters. Appl Phys Lett. 2007;91:181910.
[222] Lange U, Hirsch T, Mirsky VM, Wolfbeis OS. Hydrogen sensor based on a graphene–palladium
nanocomposite. Electrochim Acta. 2011;56:3707–3712.
[223] Eastman JA, Thompson LJ, Kestel BJ. Narrowing of the palladium–hydrogen miscibility gap in
nanocrystalline palladium. Phys Rev B. 1993;48:84–92.
[224] Suleiman M. Size-selective synthesis of nanometre-sized palladium clusters and their hydrogen
solvation behaviour [PhD dissertation]. Göttingen: University of Göttingen; 2003.
[225] Kaszkur Z, Rzeszotarski P, Juszczyk W. Powder diffraction in studies of nanocrystal surfaces:
chemisorption on Pt. J Appl Crystallogr. 2014;47:2069–2077.
[226] Ingham B, Illy BN, Ryan MP. Direct observation of distinct nucleation and growth processes
in electrochemically deposited ZnO nanostructures using in situ XANES. J Phys Chem C.
2008;112:2820–2824.
[227] Fall S, Konovalov O, Maaza M, Beye AC, Gibaud A. In situ monitoring by synchrotron radiation of
the formation of ZnO nanoparticles by aqueous chemical growth. J Appl Crystallogr. 2009;42:815–
819.
[228] Bøjesen ED, Jense KMØ, Tyrsted C, Lock N, Christensen M, Iversen BB. In situ powder diffraction
study of the hydrothermal synthesis of ZnO nanoparticles. Cryst Growth Design. 2014;14:2803–
2810.
[229] Chen W, Lu YH, Wang M, Kroner L, Paul H, Fecht HJ, Bednarcik J, Stahl K, Zhang ZL, Wiedwald
U, Kaiser U, Ziemann P, Kikegawa T, Wu CD, Jiang JZ. Synthesis, thermal stability and properties
of ZnO2 nanoparticles. J Phys Chem C. 2009;113:1320–1324.
[230] Yoon S, Baik S, Kim MG, Shin N. Formation mechanisms of tetragonal barium titanate nanoparticles
in alkoxide-hydroxide sol-precipitation synthesis. J Am Ceram Soc. 2006;89:1816–1821.
[231] Rehan M, Lai X, Kale GM. Hydrothermal synthesis of titanium dioxide nanoparticles studied
employing in situ energy dispersive X-ray diffraction. Cryst Eng Comm. 2011;13:3725–3732.
[232] Eltzholtz JR, Tyrsted C, Jensen KMØ, Bremholm M, Christensen M, Becker-Christensen J, Iversen
BB. Pulsed supercritical synthesis of anatase TiO2 nanoparticles in a water–isopropanol mixture
studied by in situ powder X-ray diffraction. Nanoscale. 2013;5:2372–2378.
[233] Sietsma JRA, Meeldijk JD, den Breejen JP, Versluijs-Helder M, van Dillen AJ, de Jongh PE, de
Jong KP. The preparation of supported NiO and Co3 O4 nanoparticles by the nitric oxide controlled
thermal decomposition of nitrates. Angew Chem Int Ed. 2007;46:4547–4549.
[234] Althues H, Kaskel S. Sulfated zirconia nanoparticles synthesised in reverse microemulsions:
preparation and catalytic properties. Langmuir. 2002;18:7428–7435.
[235] Philip J, Gnanaprakash G, Panneerselvam G, Antony MP, Jayakumar T, Raj B. Effect of ther-
mal annealing under vacuum on the crystal structure, size, and magnetic properties of ZnFe2 O4
nanoparticles. J Appl Phys. 2007;102:054305.
[236] Lai J, Shafi KVPM, Loos K, Ulman A, Lee Y, Vogt T, Estournès C. Doping γ-Fe2 O3 nanoparticles
with Mn(III) suppresses the transition to the α-Fe2 O3 structure. J Am Chem Soc. 2003;125:11470–
11471.
[237] Di Luccio T, Laera AM, Tapfer L, Kempter S, Kraus R, Nickel B. Controlled nucleation and growth
of CdS nanoparticles in a polymer matrix. J Phys Chem B. 2006;110:12603–12609.
[238] Zhang H, Gilbert B, Huang F, Banfield JF. Water-driven structure transformation in nanoparticles at
room temperature. Nature. 2003;424:1025–1029.
Crystallography Reviews 297

[239] Mainz R, Singh A, Levcenko S, Klaus M, Genzel C, Ryan KM, Unold T. Phase-transition-driven
growth of compound semiconductor crystals from ordered metastable nanorods. Nature Commun.
2014;5:3133.
[240] Waseda Y. Anomalous X-ray scattering for materials characterisation: atomic-scale structure
determination. Springer tracts in modern physics. Vol. 179. Berlin: Springer; 2002.
[241] Hodeau JL, Favre-Nicolin V, Bos S, Renevier H, Lorenzo E, Berar JF. Resonant diffraction. Chem
Rev. 2001;101:1843–1868.
[242] Samant MG, Bergeret G, Meitzner G, Boudart M. Anomalous wide angle X-ray scattering and X-ray
absorption spectroscopy of supported Pt-Mo bimetallic clusters. 1. Experimental technique. J Phys
Chem. 1988;92:3542–3546.
[243] Waasmaier D, Kirfel A. New analytical scattering factor functions for free atoms and ions. Acta
Cryst A. 1995;51:416–431.
[244] Henke BL, Gullikson EM, Davis JC. X-ray interactions: photoabsorption, scattering, transmission,
and reflection at E = 50–30,000 eV, Z = 1–92. Atomic Data Nuclear Data Tables. 1993;54:181–342.
[245] Adora S, Simon JP, Soldo-Olivier Y, Faure R, Chaînet E, Durand R. Electrochemical deposi-
tion of platinum nanoparticles on carbon: A study by standard and anomalous X-ray diffraction.
ChemPhysChem. 2004;5:1178–1184.
[246] Jeng US, Lai YH, Sheu HS, Lee JF, Sun YS, Chuang WT, Huang YS, Liu DG. Anomalous small-
and wide-angle X-ray scattering and X-ray absorption spectroscopy for Pt and Pt–Ru nanoparticles.
J Appl Crystallogr. 2007;40:S418–S422.
[247] Strasser P, Koh S, Anniyev T, Greeley J, More K, Yu C, Liu Z, Kaya S, Nordlund D, Ogasaware H,
Toney MF, Nilsson A. Lattice-strain control of the activity in dealloyed core–shell fuel cell catalysts.
Nature Chem. 2010;2:454–460.
[248] Samant MG, Bergeret G, Meitzner G, Gallezot P, Boudart M. Anomalous wide angle X-ray scattering
and X-ray absorption spectroscopy of supported Pt-Mo bimetallic clusters. 2. Atomic and electronic
structure. J Phys Chem. 1988;92:3547–3554.
[249] Connor ST, Weil BD, Misra S, Cui Y, Toney MF. Behaviours of Fe, Zn, and Ga substitution in
CuInS2 nanoparticles probed with anomalous X-ray diffraction. Chem Mater. 2013;25:320–325.
[250] Billinge SJL, Kanatzidis MG. Beyond crystallography: the study of disorder, nanocrystallinity
and crystallographically challenged materials with pair distribution functions. Chem Commun.
2004;7:749–760.
[251] Proffen T, Kim H. Advances in total scattering analysis. J Mater Chem. 2009;19:5078–5088.
[252] Egami T, Billinge SJL. Underneath the Bragg peaks: structural analysis of complex materials.
Pergamon Materials Series. 2nd ed. Vol. 16. Oxford: Elsevier Ltd.; 2012.
[253] Billinge SJL, Levin I. The problem with determining atomic structure at the nanoscale. Science.
2007;316:561–565.
[254] Peterson PF, Božin ES, Proffen T, Billinge SJL. Improved measures of quality for the atomic pair
distribution function. J Appl Crystallogr. 2003;36:53–64.
[255] Kaszkur Z. Convolutional approach to the normalization of intensity scattered by polycrystalline
substances. J Appl Crystallogr. 1990;23:180–185.
[256] Cumbrera FL, Sanchez-Bajo F, Muñoz A. An improved method for the normalization of the scattered
intensity and accurate determination of the macroscopic density of noncrystalline materials. J Appl
Crystallogr. 1995;28:408–415.
[257] Haverkamp R, Wallwork, K. X-ray pair distribution function analysis of nanostructured materials
using a Mythen detector. J. Synchrotron Radiat. 2009;16:849–856.
[258] Dykhne T, Taylor R, Florence A, Billinge SJL. Data requirements for the reliable use of atomic pair
distribution functions in amorphous pharmaceutical fingerprinting. Pharm Res. 2011;28:1041–1048.
[259] Chupas PJ, Qiu X, Hanson JC, Lee PL, Grey CP, Billinge SJL. Rapid acquisition pair distribution
function (RA-PDF) analysis. J Appl Crystallogr. 2003;36:1342–1347.
[260] Farrow CL, Billinge SJL. Relationship between the atomic pair distribution function and small-angle
scattering: implications for modeling of nanoparticles. Acta Crystallogr A. 2009;65:232–239.
[261] Petkov V, Shastri SD. Element-specific structure of materials with intrinsic disorder by high-
energy resonant X-ray diffraction and differential atomic pair-distribution functions: a study of PtPd
nanosized catalysts. Phys Rev B. 2010;81:165428.
[262] Qiu X, Proffen T, Mitchell JF, Billinge SJL. Orbital correlations in the pseudo-cubic O and
rhombohedral R phases of LaMnO3 . Phys Rev Lett. 2005;94:177203.
[263] Peterson PF, Proffen T, Jeong IK, Billinge SJL, Choi KS, Kanatzidis MG, Radaelli PG. Local
atomic strain in ZnSe1−x Tex from high real-space resolution neutron pair distribution function
measurements. Phys Rev B. 2001;63:165211.
298 B. Ingham

[264] Page K, Proffen T, Niederberger M, Seshadri R. Probing local dipoles and ligand structure in BaTiO3
nanoparticles. Chem Mater. 2010;22:4386–4391.
[265] Kodama K, Iikubo S, Taguchi T, Shamoto S. Finite size effects of nanoparticles on the atomic pair
distribution functions. Acta Crystallogr A. 2006;62:444–453.
[266] Bedford N, Dablemont C, Viau G, Chupas P, Petkov V. 3-D structure of nanosized catalysts by
high-energy X-ray diffraction and reverse Monte Carlo simulations: study of Ru. J Phys Chem C.
2007;111:18214–18219.
[267] Petkov V, Gateshki M, Choi J, Gillan EG, Ren Y. Structure of nanocrystalline GaN from X-ray
diffraction, Rietveld and atomic pair distribution function analyses. J Mater Chem. 2005;15:4654–
4659.
[268] Proffen T, Neder RB. DISCUS: a program for diffuse scattering and defect-structure simulation.
J Appl Crystallogr. 1997;30:171–175.
[269] Proffen T, Neder RB. DISCUS, a program for diffuse scattering and defect structure simulations –
update. J Appl Crystallogr. 1999;32:838–839.
[270] Page K, Hood TC, Proffen T, Neder RB. Building and refining complete nanoparticle structures with
total scattering data. J Appl Crystallogr. 2011;44:327–336.
[271] Page K, Proffen T, Terrones H, Terrones M, Lee L, Yang Y, Stemmer S, Seshadri R, Cheetham
AK. Direct observation of the structure of gold nanoparticles by total scattering powder neutron
diffraction. Chem Phys Lett. 2004;393:385–388.
[272] Juhás P, Cherba DM, Duxbury PM, Punch WF, Billinge SJL. Ab initio determination of solid-state
nanostructure. Nature. 2006;440:655–658.
[273] Juhás P, Granlund L, Duxbury PM, Punch WF, Billinge SJL. The Liga algorithm for ab initio
determination of nanostructure. Acta Crystallogr A. 2008;64:631–640.
[274] Juhás P, Granlund L, Gujarathi SR, Duxbury PM, Billinge SJL. Crystal structure solution from
experimentally determined atomic pair distribution functions. J Appl Crystallogr. 2010;43:623–629.
[275] Neder RB, Korsunskiy VI, Chory C, Müller G, Hofmann A, Dembski S, Graf C, Rühl E. Structural
characterization of II–VI semiconductor nanoparticles. Phys Stat Solidi C. 2007;4:3221–3233.
[276] Petkov V, Gateshki M, Niederberger M, Ren Y. Atomic-scale structure of nanocrystalline
Bax Sr1−x TiO3 (x = 1, 0.5, 0) by X-ray diffraction and the atomic pair distribution function technique.
Chem Mater. 2006;18:814–821.
[277] Welborn M, Tang W, Ryu J, Petkov V, Henkelman G. A combined density functional and X-ray
diffraction study of Pt nanoparticle structure. J Chem Phys. 2011;135:014503.
[278] Knecht MR, Weir MG, Myers S, Pyrz WD, Ye, H, Petkov V, Buttrey DJ, Frenkel AI, Crooks RM.
Synthesis and characterization of Pt dendrimer-encapsulated nanoparticles: effect of the template on
nanoparticle formation. Chem Mater. 2008;20:5218–5228.
[279] Simpson CA, Farrow CL, Tian P, Billinge SJL, Huffman BJ, Harkness KM, Cliffel DE. Tiopronin
gold nanoparticle precursor forms aurophilic ring tetramer. Inorg Chem. 2010;49:10858–10866.
[280] Alayoglu S, Zavalij P, Elchhorn B, Wang Q, Frenkel AI, Chupas P. Structural and architectural
evaluation of bimetallic nanoparticles: a case study of Pt–Ru core–shell and alloy nanoparticles.
ACS Nano. 2009;3:3127–3137.
[281] Jensen, KMØ, Christensen M, Juhas P, Tyrsted C, Bøjesen ED, Lock N, Billinge SJL, Iversen BB.
Revealing the mechanisms behind SnO2 nanoparticle formation and growth during hydrothermal
synthesis: an in situ total scattering study. J Amer Chem Soc. 2012;134:6785–6792.
[282] Fratzl P. Small-angle scattering in materials science – a short review of applications in alloys,
ceramics and composite materials. J Appl Crystallogr. 2003;36:397–404.
[283] Glatter O, Kratky O, editors. Small angle X-ray scattering. London: Academic Press; 1982.
[284] Pedersen JS. Analysis of small-angle scattering data from colloids and polymer solutions: modelling
and least-squares fitting. Adv Colloid Interface Sci. 1997;70:171–210.
[285] Caponetti E, Floriano MA, Di Dio E, Triolo R. On the shape of the radial distribution function of an
assembly of monodisperse ellipsoidal scatterers. J Appl Crystallogr. 1993;26:612–615.
[286] Kučerka N, Kiselev MA, Balgavý P. Determination of bilayer thickness and lipid surface area in
unilamellar dimyristoylphosphatidylcholine vesicles from small-angle neutron scattering curves: a
comparison of evaluation methods. Eur Biophys J. 2004;33:328–334.
[287] Pedersen JS. Determination of size distributions from small-angle scattering data for systems with
effective hard-sphere interactions. J Appl Crystallogr. 1994;27:595–608.
[288] Sharma RV, Sharma KC. The structure factor and the transport properties of dense fluids having
molecules with square well potential, a possible generalisation. Physica A. 1977;89:213–218.
[289] Menon SVG, Manohar C, Rao KS. A new interpretation of the sticky hard sphere model. J Chem
Phys. 1991;95:9186–9190.
Crystallography Reviews 299

[290] Beaucage G, Schaefer DW. Structural studies of complex systems using small-angle scattering: a
unified Guinier/power-law approach. J Non-Crystalline Solids. 1994;172–174:797–805.
[291] Beaucage G. Approximations leading to a unified exponential/power-law approach to small-angle
scattering. J Appl Crystallogr. 1995;28:717–728.
[292] Beaucage G. Small-angle scattering from polymeric mass fractals of arbitrary mass-fractal dimen-
sion. J Appl Crystallogr. 1996;29:134–146.
[293] Suryawanshi CN, Pakdel P, Schaefer DW. Effect of drying on the structure and dispersion of
precipitated silica. J Appl Crystallogr. 2003;36:573–577.
[294] Hammouda B. Analysis of the Beaucage model. J Appl Crystallogr. 2010;43:1474–1478.
[295] Hammouda B. A new Guinier-Porod model. J Appl Crystallogr. 2010;43:716–719.
[296] SasView for small angle scattering analysis [Internet]. Available from: http://www.sasview.org
[297] Ilavsky J, Jemian PR. Irena: tool suite for modelling and analysis of small-angle scattering. J Appl
Crystallogr. 2009;42:347–353.
[298] Irena package for analysis of small-angle scattering data [Internet]. Chicago (IL): Argonne National
Laboratory. Available from: http://usaxs.xray.aps.anl.gov/staff/ilavsky/irena.html
[299] Konarev PV, Petoukhov MV, Volkov VV, Svergun DI. ATSAS 2.1, a program package for small-angle
scattering data analysis. J Appl Crystallogr. 2006;39:277–286.
[300] Petoukhov MV, Franke D, Shkumatov AV, Tria G, Kikhney AG, Gajda M, Gorba C, Mertens
HDT, Konarev PV, Svergun DI. New developments in the ATSAS program package for small-angle
scattering data analysis. J Appl Crystallogr. 2012;45:342–350.
[301] Data analysis software ATSAS 2.5.2 [Internet]. Hamburg: European Molecular Biology Laboratory.
Available from: http://www.embl-hamburg.de/biosaxs/software.html
[302] Ingham B, Dickie S, Nanjo H, Toney MF. In situ USAXS measurements of titania colloidal paint
films. J Colloid Interface Sci. 2009;336:612–615.
[303] Stevens DA, Zhang S, Chen Z, Dahn JR. On the determination of platinum particle size in carbon-
supported platinum electrocatalysts for fuel cell applications. Carbon. 2003;41:2769–2777.
[304] Lin JM, Lin TL, Jeng US, Zhong YJ, Yeh CT, Chen TY. Fractal aggregates of the Pt nanoparticles
synthesised by the polyol process and poly(N-vinyl-2-pyrrolidone) reduction. J Appl Crystallogr.
2007;40:S540–S543.
[305] Pavlopoulou E, Portale G, Christodoulakis KE, Vamvakaki M, Bras W, Anastasiadis SH. Fol-
lowing the synthesis of metal nanoparticles within pH-responsive microgel particles by SAXS.
Macromolecules. 2010;43:9828–9836.
[306] Gröhn F, Kim G, Bauer BJ, Amis EJ. Nanoparticle formation within dendrimer-containing polymer
networks: route to new organic-inorganic hybrid materials. Macromolecules. 2001;34:2179–2185.
[307] Ceolín M, Gálvez N, Sánchez P, Fernández B, Domínguez-Vera JM. Structural aspects of the growth
mechanism of copper nanoparticles inside apoferritin. Eur J Inorg Chem. 2008;5:795–801.
[308] Biswas K, Das B, Rao CNR. Growth kinetics of ZnO nanorods: capping-dependent mechanism and
other interesting features. J Phys Chem C. 2008;112:2404–2411.
[309] Champagnon B, Andrianasolo B, Ramos A, Gandals M, Allais M, Benoit JP. Size of Cd(S,Se) quan-
tum dots in glasses: correlation between measurements by high-resolution transmission electron
microscopy, small-angle X-ray scattering, and low-frequency inelastic Raman scattering. J Appl.
Phys. 1993;73:2775–2780.
[310] Kagan CR, Murray CB, Nirmal M, Bawendi MG. Electronic energy transfer in CdSe quantum dot
solids. Phys Rev Lett. 1996;76:1517–1520.
[311] Metzger TH, Kegel I, Paniago R, Lorke A, Peisl J, Schulze J, Eisele I, Schittenhelm P, Abstreiter
G. Shape, size, strain and correlations in quantum dot systems studied by grazing incidence X-ray
scattering methods. Thin Solid Films. 1998;336:1–8.
[312] Polte J, Tuaev X, Wuithschick M, Fischer A, Thuenemann AF, Rademann K, Kraehnert R, Emmer-
ling F. Formation mechanism of colloidal silver nanoparticles: analogies and differences to the
growth of gold nanoparticles. ACS Nano. 2012;6:5791–5802.
[313] Polte J, Emmerling F, Radtke M, Reinholz U, Riesemeier H, Thünemann AF. Real-time monitoring
of copolymer stabilised growing gold nanoparticles. Langmuir. 2010;26:5889–5894.
[314] Beaucage G, Kammler HK, Mueller R, Strobel R, Agashe N, Pratsinis SE, Narayanan T. Probing the
dynamics of nanoparticle growth in a flame using synchrotron radiation. Nat Mater. 2004;3:370–373.
[315] Kammler HK, Beaucage G, Kohls DJ, Agashe N, Ilavsky J. Monitoring simultaneously the
growth of nanoparticles and aggregates by in situ ultra-small-angle X-ray scattering. J Appl Phys.
2005;97:054309.
[316] Henkel A, Schubert O, Plech A, Sönnichsen C. Growth kinetic of a rod-shaped metal nanocrystal. J
Phys Chem C. 2009;113:10390–10394.
300 B. Ingham

[317] Harada M, Katagiri E. Mechanism of silver particle formation during photoreduction using in situ
time-resolved SAXS analysis. Langmuir. 2010;26:17896–17905.
[318] Harada M, Tamura N, Takenaka M. Nucleation and growth of metal nanoparticles during photore-
duction using in situ time-resolved SAXS analysis. J Phys Chem C. 2011;115:14081–14092.
[319] LaGrow AP, Ingham B, Toney MF, Tilley RD. Effect of surfactant concentration and aggregation on
the growth kinetics of nickel nanoparticles. J Phys Chem C. 2013;117:16709–16718.
[320] Watzky MA, Finke RG. Transition metal nanocluster formation kinetic and mechanistic studies. A
new mechanism when hydrogen is the reductant:slow, continuous nucleation and fast autocatalytic
surface growth. J Am Chem Soc. 1997;119:10382–10400.
[321] Abécassis B, Testard F, Spalla O, Barboux P. Probing in situ the nucleation and growth of gold
nanoparticles by small-angle X-ray scattering. Nano Lett. 2007;7:1723–1727.
[322] Abécassis B, Testard F, Kong Q, Francois B, Spalla O. Influence of monomer feeding on a fast gold
nanoparticles synthesis: time-resolved XANES and SAXS experiments. Langmuir. 2010;26:13847–
13854.
[323] Takesue M, Tomura T, Yamada M, Hata K, Kuwamoto S, Yonezawa T. Size of elementary clusters
and process period in silver nanoparticle formation. J Am Chem Soc. 2011;133:14164–14167.
[324] Polte J, Erler R, Thünemann AF, Sokolov S, Ahner TT, Rademann K, Emmerling F, Kraehnert
R. Nucleation and growth of gold nanoparticles studied via in situ small angle X-ray scattering at
millisecond time resolution. ACS Nano. 2010;4:1076–1082.
[325] Carrado KA, Sandi G, Kizilel R, Seifert S, Castagnola N. Platinum nanoclusters immobilised on
polymer-clay nanocomposite films. Appl Clay Sci. 2005;30:94–102.
[326] Sakamoto N, Harada M, Hashimoto T. In situ and time-resolved SAXS studies of Pd nanopar-
ticle formation in a template of block copolymer microdomain structures. Macromolecules.
2006;39:1116–1124.
[327] Ustarroz J, Hammons JA, Van Ingelgem Y, Tzedaki M, Hubin A, Terryn H. Multipulse electrodeposi-
tion of Ag nanoparticles on HOPG monitored by in-situ by small-angle X-ray scattering. Electrochem
Commun. 2011;13:1320–1323.
[328] Winans RE, Vajda S, Lee B, Riley DJ, Seifert S, Tikhonov GY, Tomczyk NA. Thermal sta-
bility of supported platinum clusters studied by in situ GISAXS. J Phys Chem B. 2004;108:
18105–10107.
[329] Smith MC, Gilbert JA, Mawdsley JR, Seifert S, Myers DJ. In situ small-angle X-ray scattering
observation of Pt catalyst particle growth during potential cycling. J Am Chem Soc. 2008;130:8112–
8113.
[330] Yu C, Holby EF, Yang R, Toney MF, Morgan D, Strasser P. Growth trajectories and coarsening
mechanisms of metal nanoparticle electrocatalysts. ChemCatChem. 2012;4:766–770.
[331] Teixeira AV, Morfin I, Ehrburger-Dolle F, Rochas C, Geissler E, Licinio P, Panine P. Scattering from
dilute ferrofluid suspensions in soft polymer gels. Phys Rev E. 2003;67:021504.
[332] Tokumoto MS, Pulcinelli SH, Santilli CV, Craievich AF. SAXS study of the kinetics of formation of
ZnO colloidal suspensions. J Non-Crystalline Solids. 1999;247:176–182.
[333] Caetano BL, Santilli CV, Meneau F, Briois V, Pulcinelli SH. In situ and simultaneous UV–vis/SAXS
and UV–vis/XAFS time-resolved monitoring of ZnO quantum dots formation and growth. J Phys
Chem C. 2011;115:4404–4412.
[334] Lee EJH, Ribeiro C, Longo E, Leite ER. Growth kinetics of tin oxide nanoparticles in colloidal
suspensions under hydrothermal conditions. Chem Phys. 2006;328:229–235.
[335] Lee EJH, Ribeiro C, Longo E, Leite ER. Oriented attachment: an effective mechanism in the
formation of anisotropic nanocrystals. J Phys Chem B. 2005;109:20842–20846.
[336] Boissiere C, Grosso D, Amenitsch H, Gibaud A, Coupé A, Baccile N, Sanchez C. First in-situ
SAXS studies of the mesostrucuration of spherical titania particle during spray-drying process. Chem
Commun. 2003;22:2798–2799.
[337] Brumberger H, Hagrman D, Goodisman J, Finkelstein KD. In situ anomalous small-angle X-
ray scattering from metal particles in supported-metal catalysts. I. Theory. J Appl Crystallogr.
2005;38:147–151.
[338] Goerigk G, Haubold HG, Lyon O, Simon JP. Anomalous small-angle X-ray scattering in materials
science. J Appl Crystallogr. 2003;36:425–429.
[339] Haubold HG, Wang XH, Jungbluth H, Goerigk G, Schilling W. In situ anomalous small-angle X-ray
scattering and X-ray absorption near-edge structure investigation of catalyst structures and reactions.
J Mol Struct. 1996;383:283–289.
[340] Haubold HG, Wang XH, Goerigk G, Schilling W. In situ anomalous small-angle X-ray scattering
investigation of carbon-supported electrocatalysts. J Appl Crystallogr. 1997;30:653–658.
Crystallography Reviews 301

[341] Vainio U, Pirkkalainen K, Kisko K, Goerigk G, Kotelnikova NE, Serimaa R. Copper and copper
oxide nanoparticles in a cellulose support studied using anomalous small-angle X-ray scattering. Eur
Phys J D. 2007;42:93–101.
[342] Bóta A, Varga Z, Goerigk G. Biological systems as nanoreactors: anomalous small-angle scattering
study of the CdS nanoparticle formation in multilamellar vesicles. J Phys Chem B. 2007;111:1911–
1915.
[343] Benedetti A, Bertoldo L, Canton P, Goerigk G, Pinna F, Riello P, Polizzi S. ASAXS study of Au, Pd
and Pd–Au catalysts supported on active carbon. Catal Today. 1999;49:485–489.
[344] Haubold HG, Wang XH. ASAXS studies of carbon supported electrocatalysts. Nucl Instrum Methods
B. 1995;97:50–54.
[345] Yu C, Koh S, Leisch JE, Toney MF, Strasser P. Size and composition distribution dynamics of alloy
nanoparticle electrocatalysts probed by anomalous small angle X-ray scattering. Faraday Discuss.
2009;140:283–296.
[346] Haubold HG, Vad T, Waldöfner N, Bönnemann H. From Pt molecules to nanoparticles: in-situ
(anomalous) small-angle X-ray scattering studies. J Appl Crystallogr. 2003;36:617–620.
[347] Vad T, Haubold HG, Waldöfner N, Bönnemann H. Three-dimensional Pt-nanoparticle networks
studied by anomalous small-angle X-ray scattering and X-ray absorption spectroscopy. J Appl
Crystallogr. 2002;35:459–470.
[348] Lee B, Seifert S, Riley SJ, Tikhonov G, Tomczyk NA, Vajda S, Winans RE. Anomalous grazing inci-
dence small-angle X-ray scattering studies of platinum nanoparticles formed by cluster deposition. J
Chem Phys. 2005;123:074701.
[349] Murray CB, Kagan CR, Bawendi MG. Synthesis and characterisation of monodisperse nanocrystals
and close-packed nanocrystal assemblies. Annu Rev Mater Sci. 2000;30:545–610.
[350] Saunders AE, Korgel BA. Observation of an AB phase in bidisperse nanocrystal superlattices.
ChemPhysChem. 2005;6:61–65.
[351] Shevchenko EV, Talapin DV, Kotov NA, O’Brien S, Murray CB. Structural diversity in binary
nanoparticle superlattices. Nature. 2006;439:55–59.
[352] Talapin DV, Shevchenko EV, Bodnarchuk MI, Ye X, Chen J, Murray CB. Quasicrystalline order in
self-assembled binary nanoparticle superlattices. Nature. 2009;461:964–967.
[353] Murray CB, Sun S, Doyle H, Betley T. Monodisperse 3d transition-metal (Co, Ni, Fe) nanoparticles
and their assembly into nanoparticle superlattices. Mater Res Soc Bull. 2001;26:985–991.
[354] Kalsin AM, Fialkowski M, Paszewski M, Smoukov SK, Bishop KJM, Grzybowski BA. Electrostatic
self-assembly of binary nanoparticle crystals with a diamond-like lattice. Science. 2006;312:420–
424.
[355] Murray CB, Kagan CR, Bawendi MG. Self-organisation of CdSe nanocrystallites into three-
dimensional quantum dot superlattices. Science. 1995;270:1335–1338.
[356] Dong A, Chen J, Vora PM, Kikkawa JM, Murray CB. Binary nanocrystal superlattice membranes
self-assembled at the liquid–air interface. Nature. 2010;466:474–477.
[357] Korgel BA, Zaccheroni N, Fitzmaurice D. “Melting transition” of a quantum dot solid: collec-
tive interactions influence the thermally-induced order-disorder transition of a silver nanocrystal
superlattice. J Am Chem Soc. 1999;121:3533–3534.
[358] Heitsch AT, Patel RN, Goodfellow BW, Smilgies DM, Korgel BA. GISAXS characterisation of order
in hexagonal monolayers of FePt nanocrystals. J Phys Chem C. 2010;114:14427–14432.
[359] Park SY, Lytton-Jean AKR, Lee B, Weigand S, Schatz GC, Mirkin CA. DNA-programmable
nanoparticle crystallisation. Nature. 2008;451:553–556.
[360] Nykypanchuk D, Maye M, van der Lelie D, Gang O. DNA-guided crystallisation of colloidal
nanoparticles. Nature. 2008;451:549–552.
[361] Disch S, Wetterskog E, Hermann RP, Salazar-Alvarez G, Busch P, Brückel T, Bergström L, Kamali
S. Shape induced symmetry in self-assembled mesocrystals of iron oxide nanocubes. Nano Lett.
2011;11:1651–1656.
[362] Siffalovic P, Majkova E, Chitu L, Jergel M, Luby S, Satka A, Roth SV. Self-assembly of iron oxide
nanoparticles studied by time-resolved grazing-incidence small-angle X-ray scattering. Phys Rev B.
2007;76:195432.
[363] Smith DK, Goodfellow B, Smilgies DM, Korgel BA. Self-assembled simple hexagonal AB2 binary
nanocrystal superlattices: SEM, GISAXS and defects. J Am Chem Soc. 2009;131:3281–3290.
[364] Yu Y, Bosoy CA, Smilgies DM, Korgel BA. Self-assembly and thermal stability of binary
superlattices of gold and silicon nanocrystals. J Phys Chem Lett. 2013;4:3677–3682.
[365] McNeil BWJ, Thompson NR. X-ray free-electron lasers. Nat Photonics. 2010;4:814–821.
302 B. Ingham

[366] Patterson BD. Crystallography using an X-ray free-electron laser. Crystallogr Rev. 2014;20:
242–294.
[367] Ackermann W, Asova G, Ayvazyan V, Azima A, Baboi N, Bähr J, Balandin V, Beutner B, Brandt
A, Bolzmann A, Brinkmann R, Brovko OI, Castellano M, Castro P, Catani L, Chiadroni E, Choroba
S, Cianchi A, Costello JT, Cubaynes D, Dardis J, Decking W, Delsim-Hashemi H, Delserieys A, Di
Pirro G, Dohlus M, Düsterer S, Eckhardt A, Edwards HT, Faatz B, Feldhaus J, Flöttmann K, Frisch J,
Fröhlich L, Garvey T, Gensch U, Gerth C, Görler M, Golubeva N, Grabosch HJ, Grecki M, Grimm
O, Hacker K, Hahn U, Han JH, Honkavaara K, Hott T, Hüning M, Ivanisenko Y, Jaeschke E, Jal-
muzna W, Jezynski T, Kammering R, Katalev V, Kavanagh K, Kennedy ET, Khodyachykh S, Klose
K, Kocharyan V, Körfer M, Kollewe M, Koprek W, Korepanov S, Kostin D, Krassilnikov M, Kube
G, Kuhlmann M, Lewis CLS, Lilje L, Limberg T, Lipka D, Löhl F, Luna H, Luong M, Martins M,
Meyer M, Michelato P, Miltchev V, Möller WD, Monaco L, Müller WFO, Napieralski O, Napoly O,
Nicolosi P, Nölle D, Nuñez T, Oppelt A, Pagani C, Paparella R, Pchalek N, Pedregosa-Gutierrez J,
Petersen B, Petrosyan B, Petrosyan G, Petrosyan L, Pflüger J, Plönjes E, Poletto L, Pozniak K, Prat
E, Proch D, Pucyk P, Radcliffe P, Redlin H, Rehlich K, Richter M, Roehrs M, Roensch J, Romaniuk
R, Ross M, Rossbach J, Rybnikov V, Sachwitz M, Saldin EL, Sandner W, Schlarb H, Schmidt B,
Schmitz M, Schmüser P, Schneider JR, Schneidmiller EA, Schnepp S, Schreiber S, Seidel M, Ser-
tore D, Shabunov AV, Simon C, Simrock S, Sombrowski E, Sorokin AA, Spanknebel P, Spesyvtsev
R, Staykov L, Steffen B, Stephan F, Stulle F, Thom H, Tiedtke K, Tischer M, Toleikis S, Treusch R,
Trines D, Tsakov I, Vogel E, Weiland T, Weise H, Wellhöfer M, Wendt M, Will I, Winter A, Witten-
burg K, Wurth W, Yeates P, Yurkov MV, Zagorodnov I, Zapfe K. Operation of a free-electron laser
from the extreme ultraviolet to the water window. Nat Photonics. 2007;1:336–342.
[368] FLASH [Internet]. Available from: http://photon-science.desy.de/facilities/flash/index_eng.html
[369] Emma P, Akre R, Arthur J, Bionta R, Bostedt C, Bozek J, Brachmann A, Bucksbaum P, Coffee R,
Decker FJ, Ding Y, Dowell D, Edstrom S, Fisher A, Frisch J, Gilevich S, Hastings J, Hays G, Hering
P, Huang Z, Iverson R, Loos H, Messerschmidt M, Miahnahri A, Moeller S, Nuhn HD, Pile G, Ratner
D, Rzepiela J, Schultz D, Smith T, Stefan P, Tompkins H, Turner J, Welch J, White W, Wu J, Yocky G,
Galayda J. First lasing and operation of an angstrom-wavelength free-electron laser. Nat Photonics.
2010;4:641–647.
[370] Linac coherent light source [Internet]. Available from: http://lcls.slac.stanford.edu
[371] Ishikawa T, Aoyagi H, Asaka T, Asano Y, Azumi N, Bizen T, Ego H, Fukami K, Fukui T, Furukawa
Y, Goto S, Hanaki H, Hara T, Hasegawa T, Hatsui T, Higashiya A, Hirono T, Hosoda N, Ishii M,
Inagaki T, Inubushi Y, Itoga T, Joti Y, Kago M, Kameshima T, Kimura H, Kirihara Y, Kiyomichi A,
Kobayashi T, Kondo C, Kudo T, Maesaka H, Marchal XM, Masuda T, Matsubara S, Matsumoto T,
Matsushita T, Matsui S, Nagasono M, Nariyama N, Ohashi H, Ohata T, Ohshima T, Ono S, Otake
Y, Saji C, Sakurai T, Sato T, Sawada K, Seike T, Shirasawa K, Sugimoto T, Suzuki S, Takahashi
S, Takebe H, Takeshita K, Tamasaku K, Tanaka H, Tanaka R, Tanaka T, Togashi T, Togawa K,
Tokuhisa A, Tomizawa H, Tono K, Wu S, Yabashi M, Yamaga M, Yamashita A, Yanagida K, Zhang
C, Shintake T, Kitamura H, Kumagai N. A compact X-ray free electron laser emitting in the sub-
angstrom region. Nat Photonics. 2012;6:540–544.
[372] SACLA (XFEL) [Internet]. Available from: http://xfel.riken.jp/eng/index.html
[373] Allaria E, Castronovo D, Cinquegrana P, Craievich P, Dal Forno M, Danailov MB, D;Auria G, Demi-
dovich A, De Ninno G, Di Mitri S, Diviacco B, Fawley WM, Ferianis M, Ferrari E, Froehlich L, Gaio
G, Gauthier D, Giannessi L, Ivanov R, Mahieu B, Mahne N, Nikolov I, Parmigiani F, Penco G, Rai-
mondi L, Scafuri C, Serpico C, Sigalotti P, Spampinati S, Spezzani C, Svandrlik M, Svetina C, Trovo
M, Veronese M, Zangrando D, Zangrando M. Two-stage seeded soft-X-ray free-electron laser. Nat
Photonics. 2013;7:913–918.
[374] Elettra and FERMI lightsources [Internet]. Available from: http://www.elettra.eu/lightsources/fermi.
html
[375] Prat E, Aiba M, Bettoni S, Beutner B, Reiche S, Schietinger T. Emmittance measurements and
minimisation at the SwissFEL injector test facility. Phys Rev Spec Top Accelerators Beams.
2014;17:104401.
[376] SwissFEL, Paul Scherrer Institut [Internet]. Available from: http://www.psi.ch/swissfel/swissfel
[377] European XFEL [Internet]. Available from: http://www.xfel.eu/
[378] Choi J, Huang JY, Kang HS, Kim MG, Yim CM, Lee TY, Oh JS, Parc YW, Park JH, Park SJ, Ko IS,
Kim YJ. Design of the PAL XFEL. J Korean Phys Soc. 2007;50:1372–1376.
[379] Han JH, Hong J, Lee JH, Chae MS, Baek SY, Choi HJ, Ha T, Hu J, Hwang WH, Jung SH, Kim
CB, Kim CH, Kim IY, Kim JM, Kim SH, Lee HS, Lee SJ, Lee YY, Min CK, Mun GY, Na DH,
Crystallography Reviews 303

Park SJ, Park SS, Park YJ, Son YK, Yang HR, Kang HS, Ko IS. Beam operation of the PAL-
XFEL injector test facility. Proc FEL2014, Basel, Switzerland. 2014; WEB02. Available from:
http://www.fel2014.ch/prepress/FEL2014/papers/web02.pdf
[380] Takahashi Y, Suzuki A, Zettsu N, Oroguchi T, Takayama Y, Sekiguchi Y, Kobayashi A, Yamamoto
M, Nakasako M. Coherent diffraction imaging analysis of shape-controlled nanoparticles with
focused hard X-ray free-electron laser pulses. Nano Lett. 2013;13:6028–6032.
[381] Pedersoli E, Loh NH, Capotondi F, Hampton CY, Sierra RG, Starodub D, Bostedt C, Bozek J, Nelson
AJ, Aslam M, Li S, Dravid VP, Martin AV, Aquila A, Barty A, Fleckenstein H, Gumprecht L, Liang
M, Nass K, Schulz J, White TA, Coppola N, Bajt S, Barthelmess M, Graafsma H, Hirsemann H,
Wunderer C, Epp SW, Erk B, Rudek B, Rudenko A, Foucar L, Kassemeyer S, Lomb L, Rolles
D, Shoeman RL, Steinbrener J, Hartmann R, Hartmann A, Hauser G, Holl P, Kimmel N, Reich
C, Soltau H, Weidenspointer G, Benner WH, Farquar GR, Hau-Riege SP, Hunter MS, Ekeberg T,
Hantke M, Maiai FRNC, Tobias HJ, Marchesini S, Frank M, Strüder L, Schlichting I, Ullrich J,
Chapman HN, Bucksbaum PH, Kiskinova M, Bogan MJ. Mesoscale morphology of airborne core–
shell nanoparticle clusters: X-ray laser coherent diffraction imaging. J Phys B: At Mol Opt Phys.
2013;46:164033.
[382] Bogan MJ, Boutet S, Barty A, Benner WH, Frank M, Lomb L, Shoeman R, Starodub D, Seibert
MM, Hau-Riege SP, Woods B, Decorwin-Martin P, Bajt S, Schulz J, Rohner U, Iwan B, Timneanu
N, Machesini S, Schlichting I, Hajdu J, Chapman HN. Single-shot femtosecond X-ray diffrac-
tion from randomly oriented ellipsoidal nanoparticles. Phys Rev Spec Top Accelerators Beams.
2010;13:094701.
[383] Carnis J, Cha W, Wingert J, Kang J, Jiang Z, Song S, Sikorski M, Robert A, Gutt C, Chen SW, Dai
Y, Ma Y, Guo H, Lurio LB, Shpyrko O, Narayanan S, Cui M, Kosif I, Emrick T, Russell TP, Lee HC,
Yu CJ, Grübel G, Sinha SK, Kim H. Demonstration of feasibility of X-ray free electron laser studies
of dynamics of nanoparticles in entangled polymer melts. Sci Rep. 2014;4:6017.

Subject index

Anomalous SAXS, 51 Oxide nanoparticles, ex situ XRD studies of, 11


Anomalous XRD, 27 Oxide nanoparticles, in situ SAXS studies of, 50
Crystallite size, 4 Oxide nanoparticles, in situ XRD studies of, 23
Debye scattering, 6, 11, 15 Pair distribution function analysis, 29
Detectors, SAXS, 40 Preferred orientation, 6
Detectors, XRD, 6 Sample cell design, in situ SAXS, 44
Grazing-incidence SAXS, 40, 55 Sample cell design, in situ XRD, 16 SAXS analysis meth-
Metal nanoparticles, ex situ SAXS studies of, 41 ods, 39
Metal nanoparticles, ex situ XRD studies of, 9 Semiconductor nanoparticles, ex situ SAXS studies
Metal nanoparticle post-synthesis processing, in situ of, 43
SAXS studies of, 48 Semiconductor nanoparticles, ex situ XRD studies of, 14
Metal nanoparticle post-synthesis processing, in situ Semiconductor nanoparticles, in situ XRD studies of, 26
XRD studies of, 19 Shape anisotropy, 5
Metal nanoparticle synthesis, in situ SAXS studies of, 44 Stacking faults, 6Strain, 5
Metal nanoparticle synthesis, in situ XRD studies of, 17 Unified model for SAXS, 39
Ordered nanoparticle assemblies, 53 Williamson-Hall analysis, 5
Oxide nanoparticles, ex situ SAXS studies of, 42 XRD analysis methods, 4

You might also like