You are on page 1of 24

Chapter 6

Synchrotron Small-Angle X-Ray Scattering


Studies of Colloidal Suspensions
T. Narayanan
Abstract This chapter presents a review of recent advances in synchrotron small-
angle scattering for the investigation of structure and interactions in colloidal sys-
tems. Examples covered are representative to illustrate the quantitative features of
SAXS. Techniques include anomalous, time-resolved and ultra small-angle scat-
tering applied to bulk studies. Topics discussed span structure and interactions in
hard and soft colloids, dynamics of self-assembly of amphiphilic molecules, growth
kinetics of aerosol particles in ames and counterion distribution around polyelec-
trolyte brushes. For the sake of clarity, some basic notions of scattering from partic-
ulate systems are also presented.
6.1 Introduction
Small-angle X-ray and neutron scattering (SAXS and SANS, respectively) tech-
niques are widely used to probe the microstructure and interactions in colloidal sys-
tems [1, 2]. The scattering contrast in the case of X-rays originates from the spatial
uctuations of the electron density while neutron contrast arises from the atomic
nuclei without any systematic dependence on atomic number. As a result, these
techniques provide complementary structural information. This chapter primarily
deals with certain unique applications of SAXS in the investigations of colloidal
systems made possible by the combination of high brightness offered by modern
synchrotron sources and the high-performance detectors. Examples include interac-
tions in hard colloids, multi-scale structure of self-assembled soft colloids, charge
distribution around polyelectrolyte brushes and the amphiphilic self-assembly in the
millisecond time scale. Some of these studies also require combination of SAXS
with advanced sample environments. Such combined experiments have broadened
the scope of synchrotron scattering techniques in soft matter research.
T. Narayanan
European Synchrotron Radiation Facility F-38043, Grenoble, France, narayan@esrf.fr
Narayanan, T.: Synchrotron Small-Angle X-Ray Scattering Studies of Colloidal Suspensions.
Lect. Notes Phys. 776, 133156 (2009)
DOI 10.1007/978-3-540-95968-7 6 c Springer-Verlag Berlin Heidelberg 2009
134 T. Narayanan
It should be noted that when studying the morphology of self-assembled col-
loidal systems, SANS has certain advantage in terms of selectively investigating
their different constituents by contrast matching method [2]. The extensive liter-
ature on the self-assembled block copolymer and surfactant systems will not be
discussed here. Instead, certain specic examples are chosen to highlight the quan-
titative information that can be derived by synchrotron SAXS thanks to the high
angular and time resolution and the low detection threshold. The high brightness
of modern synchrotron sources is also exploited in dynamic scattering experiments
using the coherent part of the beam (X-ray photon correlation spectroscopy, XPCS)
which probe the equilibrium dynamics in the system. This topic is not discussed
here, so are the studies of variety of ordered colloidal states and their behaviour
in ow.
6.2 General Principles
The basic formalism of small-angle scattering is similar for light, neutrons and
X-rays. As noted before, the important difference is in the interaction of the radia-
tion with the scattering medium. Detailed derivations of the expressions are readily
available in many textbooks and therefore omitted in this section except in few cases
[1, 35]. In the following, some basic denitions relevant to examples discussed in
this chapter are provided.
6.2.1 Momentum Transfer and Differential Scattering
Cross Section
Figure 6.1 depicts the scattering geometry of a typical SAXS experiment. A highly
collimated and monochromatic X-ray beamof wavelength () impinges on a sample
and the scattered intensity in the forward direction is recorded by a two-dimensional
detector. The transmitted primary beam also could be recorded by a detector em-
bedded in the beamstop, and the entire ight path before and after the sample is in
vacuum to avoid absorption and scattering by air. In the experiment, the number of

detector sample beamstop


vacuum
>
>
k
s
k
i
k
s
k
i
q

q
X-ray beam ()
Fig. 6.1 Schematic layout of a SAXS set-up depicting the incident, scattered and transmitted X-ray
beams, the two-dimensional detector and the denition of the scattering vector (q)
6 Synchrotron SAXS of Colloidal Suspensions 135
photons scattered as a function of the scattering angle () is measured but for a given
sample, the quantity that can be compared in different experiments is the number of
photons scattered into unit solid angle, normalized by the incident photon ux (pho-
tons per second per unit area). For X-rays, the scattering originates from electrons
(Thomson scattering) [6], and at small angles the scattering is fully elastic. The
magnitudes of the incident and scattered wave vectors are equal, |k
i
| =|k
s
| = 2/
and the refractive index is close to unity. The momentum transfer or scattering vec-
tor q = k
s
k
i
, and its magnitude,
q =|q| =
4

sin(/2) (6.1)
This quantity indicates the typical length scales probed by the scattering experi-
ment. In synchrotron SAXS, the q-range covered can be three orders of magnitude,
typically 0.006nm
1
< q < 6nm
1
corresponding to real space dimension of 1m
down to 1 nm [7].
In the small angle (neglecting the polarization factor), the coherent scattered in-
tensity from an atom containing Z electrons is given by,
I
coh
= r
e
2
Z

1
f
e
2
= r
e
2
f
2
(6.2)
The quantity f is the atomic scattering factor and r
e
is the Thomson scattering length
equal to the classical electron radius (= 2.818 10
15
m) [6]. f Z except near
the atomic absorption edge and the product f r
e
is the scattering length b.
The interaction between the incident photons and the scattering medium within
the sample is contained in the quantity namely the differential scattering cross sec-
tion (d/d). In an experiment as that depicted in Fig. 6.1, the incident photon
intensity per unit area per unit time (I
0
) is scattered by a sample and the scattered
photons are acquired by each detector element subtending a solid angle with an
efciency . The measured scattered intensity is given by
I
S
= I
0
T
r
A
s
l
s
d
d
(6.3)
where T
r
is the sample transmission, A
s
is the cross section of the beam, l
s
is the
sample thickness and d/d is the differential scattering cross section per unit
volume. The transmitted intensity per unit area per second is I
T
= I
0
exp(
l
l
s
),
where
l
is the linear absorption coefcient and T
r
=I
T
/I
0
. The quantity that can be
directly compared to a model is d/d which contains information about the struc-
ture and the interactions in the system over the range of q spanned by the scattering
experiment, and it is expressed in units of reciprocal of length times solid angle
(m
1
sterad
1
). Therefore, an essential step to reach a quantitative understanding of
the measured intensities is the normalization of the experimental data to d/d
which henceforth will be denoted by I(q) and given simply in units of reciprocal
length.
136 T. Narayanan
To convert measured intensities to d/d, the absolute value of the detector ef-
ciency () is also needed. This can be measured using a sample of known d/d.
Of course, the cumulative parasitic background of the instrument and the container
should be much lower than the scattering from the calibration sample. Over the
q-range 13nm
1
scattering from a clean liquid, composed of light atoms with
known isothermal compressibility (K
T
) and density, is an appropriate intensity stan-
dard. The intensity corresponding to the low q-plateau of the liquid structure factor
(q < 3nm
1
) is given by d/d = N
2
n
2
e
r
2
e
k
B
TK
T
, where N is the number den-
sity of molecules, n
e
is the number of electrons per molecule, k
B
is the Boltzmann
constant and T is the absolute temperature. For example, water is a good intensity
standard for SAXS and after proper background subtraction, I(q) is nearly at over
the range 1nm
1
q 4nm
1
, with d/d 1.610
3
mm
1
at 25

C [7].
An essential step before attempting to understand the normalized intensity in
terms of a quantitative model is the accurate subtraction of the combined back-
ground scattering fromthe instrument, the sample cell and the mediumor the solvent
in the case of particulate systems. This background intensity should be normalized
in an identical manner as the sample scattered intensity.
6.2.2 Form Factor
For a dilute suspension containing N uniform particles per unit volume, the inter-
particle interactions can be neglected and I(q) mainly depend on the shape and size
of particles.
I(q) = N|F(q)|
2
(6.4)
where F(q) is the coherent sumof the scattering amplitudes of the individual scatter-
ing centres within the particle given by the Fourier transform of the electron density
distribution [1]. In small-angle scattering, the electron density can be approximated
as a continuous function. Therefore, the scattering amplitude of the particle, is given
by
F(q) =

V
(r)e
iq.r
dV (6.5)
with V the volume of the particle and the scattering length density which is the
product of electron density and r
e
. In the case of uniform electron density,
=
n
e
d
M
N
A
M
W
r
e
(6.6)
where n
e
is the number of electrons in a molecule, d
M
is the mass density, N
A
is
the Avogadro number and M
W
is the molar mass, e.g. for water (d
M
= 10
3
Kg/m
3
),
9.410
4
nm
2
or colloidal silica particles (d
M
210
3
Kg/m
3
), 1.7
10
3
nm
2
. When the scattering units are embedded in a medium (e.g. solvent), the
relative scattering length density or the contrast length density ( =
M
) is the
relevant parameter that determines the scattering power. Inserting the spatial average
6 Synchrotron SAXS of Colloidal Suspensions 137
of the phase factor into Eq. (6.5) and for isotropic particles, the Fourier transform
reduces to the one-dimensional form [1, 4],
F(q) = 4


0
(r)
sin(qr)
qr
r
2
dr (6.7)
For a uniform spherical particle of radius R
S
and volume V
S
, Eq. (6.7) leads to
|F(q)|
2
=V
2
S

3[sin(qR
S
) qR
S
cos(qR
S
)]
(qR
S
)
3

2
=V
2
S

2
P(q, R
S
) (6.8)
P(q, R
S
) is the scattering form factor that describes the shape of the particle and for a
sphere it is the Bessel function within the brackets. The product NV
S
is the volume
fraction of the particles,
S
. Table 6.1 lists P(q, R) function for a few shapes which
will be used latter in this chapter. A comprehensive list of P(q, R) functions for
different particle shapes frequently occurring in scattering from colloidal systems
can be found in Ref. [8, 9].
Table 6.1 Form factors of a few commonly observed shapes in scattering from colloidal systems
Uniform sphere of
radius, R
S
P(q, R
S
) = F
2
0
(qR
S
) =

3[sin(qR
S
)qR
S
cos(qR
S
)]
(qR
S
)
3

2
Randomly oriented
cylinder or circular
disk of radius R
C
and height H
P(q, R
C
, H) =

/2
0

2J
1
(qR
C
sin)
qR
C
sin

sin((qH/2)cos)
(qH/2)cos

2
sind
J
1
rst-order Bessel function and
orientation angle
Spherical shell of
inner and outer radii
R
1
and R
2
V
2
P(q, R
1
, R
2
) =
16
2
9

R
3
2
F
0
(qR
2
) R
3
1
F
0
(qR
1
)

2
V Volume of the shell
Spherical coreshell
of core and shell
radii R
1
and R
2
F
2
(q, R
1
, R
2
) = [V
2

2
F
0
(qR
2
) V
1

1
F
0
(qR
1
)]
2
V
1
and V
2
are volumes of inner and outer spheres,
1
and
2
are
contrast between shell and core, and shell and medium, respectively
6.2.3 Effect of Polydispersity
A distinguishing feature of objects in the colloidal scale from the atomic systems is
the nite polydispersity in size and other properties. For example, Eq. (6.8) has zeros
at qR
S
=4.5, 7.73 which for real particles appear as minima in the scattered intensity
138 T. Narayanan
10
1
10
2
10
3
10
4
10
2
10
1
10
1
10
0
Fig. 6.2 Scattered intensity from a dilute suspension of polymethylmethacrylate (PMMA) la-
tex particles (volume fraction,
S
0.003) [10]. The dotted line represents Eq. (6.8) with R
S

40.9nm. The continuous line is a t to polydisperse sphere function given by Eq. (6.9) for Gaus-
sian size distribution with mean radius 40.9 nm and RMS deviation 2.8 nm. The lower inset shows
the size distribution function
as depicted in Fig. 6.2. For particles with uniform density, these minima can be used
for determining the approximate radius. The polydispersity can be described by a
size distribution function, D(R), with

D(R)dr = 1 and a lower cut-off at R = 0.
The resulting I(q) is given by [8, 9],
I(q) = N
2


0
D(R)V
2
(R)P(q, R)dR (6.9)
Analytical expression for Eq. (6.9) exists in the case of spherical particles for size
distributions such as Gaussian, Schulz [10, 11]. The polydispersity is given by the
ratio of the root mean square (RMS) deviation of R
S
(
s
) and the mean value of R
S
.
In general, Eq. (6.9) can be numerically integrated using an experimentally deter-
mined D(R) or a theoretical function to deduce the mean radius and the polydisper-
sity. Moreover, in real systems the distribution may be skewed to larger sizes that
leads to better visibility of the oscillations at high q (see Fig. 6.2) for the same mean
radius and polydispersity. As a result, the polydispersity will be overestimated if the
analysis is based on the rst minimum alone.
6.2.4 Limiting Form of I(q)
For non-interacting particles, irrespective of the shape, the exponential factor in
Eq. (6.5) can be expanded in terms of the radius of gyration (R
G
) at very small
q-values. This leads to the well-known Guinier law [3],
6 Synchrotron SAXS of Colloidal Suspensions 139
I(q) = NV
2

2
exp

q
2
R
2
G
3

(6.10)
This approximation is valid only for qR
G
< 1 (or the leading term in q
2
) and it is
widely used in small-angle scattering for determining R
G
from the lnI(q) vs q
2
plot.
In the asymptotic limit, qR
G
1, the scattering probes the interface of the parti-
cles. In the case of homogenous particles with average surface area, S, this leads to
the Porod behaviour,
I(q) = 2N
2
S q
4
(6.11)
signifying a sharp interface between the particle and the medium [1, 4]. Power-law
variation of I(q) is very commonly observed in SAXS from particulate systems
composed of both compact and fractal morphologies as summarized below [1],
I(q) q
p

p = 4 sharp interface
3 p < 4 surface fractal
p < 3 mass fractal
p 2 gaussian polymer chain
(6.12)
The power-law exponent, p > 4, can be observed in the case of a non-fractal dif-
fuse interface. In practice, extreme care must be exercised when determining the
asymptotic power-law behaviour from I(q). Improper background subtraction, non-
linearity of the detector, etc. can undermine an accurate determination of p. In the
case of monodisperse particles, this power-law region is modulated by the oscilla-
tions in the Bessel function as shown in Fig. 6.2.
Many polydisperse systems consist of multiple structural levels such as primary
particles, aggregates and their agglomerates (e.g. pyrolytically grown aerosol parti-
cles). Such systems display structurally limited power-law regions with intervening
Guinier regions [12, 13]. In this case, the local scattering laws and their crossover
can be described by the so-called unied scattering function [12],
I(q) = Gexp

q
2
R
2
G
3

+B

er f

qR
G
/

3
q

p
(6.13)
where G=N
2
V
2
, B =2N
2
S and er f is the error function. This global func-
tion does not introduce additional parameters other than those involved in the local
scattering laws. For monodisperse spheres, BR
4
G
/G = 1.62 and analogous ratios can
be arrived for a variety of terminal size distributions [13]. The ratio BR
4
G
/(1.62 G) is
called the polydispersity index which increases from 1 in the case of monodisperse
spheres to about 10 for the DebyeBueche function. For a polydisperse system, R
G
is related to mean particle radius, R
S
, through the moments of the size distribu-
tion [13]. Equation (8.13) can be extended to n structural levels by adding the corre-
sponding number of terms for each of these levels. In this case, the power law of the
ith level should be cut-off at the high q-region by the form factor of the immediate
140 T. Narayanan
10
2
10
1
10
0
10
0
10
1
10
2
10
3
10
4
10
1
10
2
R
G1
= 84 nm
R
G2
= 3.5 nm
p
1
= 3.7
p
2
= 3.2
Buffer scattering level
Casein micelles
Unified function
Large globular structure
Small non-globular structure
Calcium phosphate particles
I
(
q
)

(
m
m

1
)
q (nm
1
)
Fig. 6.3 Scattered intensity from a dilute suspension of casein micelles (volume fraction,
S

0.01). The continuous line depicts a t to the unied function involving two structural levels.
Dotted and dashed lines represent different contributions in Eq. (6.13) with parameters indicated
in the legend. The continuous line for the lower level also show a polydisperse disk function given
in Table 6.1 with parameters R
C
and H, 2.4 and 1.4 nm, respectively
lower level [12]. That is, by multiplying the power-law term with exp(q
2
R
2
G
/3)
term of the lower level.
Figure 6.3 shows the application of the unied function to describe the scattering
from a dilute suspension of casein micelles. These micelles are very polydisperse
and they constitute the main protein component of milk [14]. The internal structure
of these micelles, formed by the self-assembly of different caseins (, and ) and
calcium phosphate, is an outstanding issue. The analysis based on two-level struc-
ture shows that the larger unit (R
G
=84nm) is nearly globular but displays a weaker
power law(p 3.7) than q
4
, suggestive of the brushy interface of these micelles. In
addition, the weak oscillations in the data around q 0.08 and 0.2nm
1
feature the
form factor of these polydisperse brushes. The smaller level (R
G
= 3.5nm) exhibits
even a weaker power law (p 3.2) which is constituted presumably by the cal-
cium phosphate particles reticulated in the inner protein matrix [15]. Alternatively,
this high q part can also be described by polydisperse disk-like objects (colloidal
calcium phosphate particles) embedded in a polymer-like mesh.
6.2.5 Structure Factor
When the particulate system is more concentrated, d/d involves an additional
term corresponding to the interparticle interactions [16, 17]. This interference term
or the structure factor, S(q), is a function of N and the interaction potential, U(r).
In a dilute non-interacting system, S(q) 1. For particles with spherical symmetry
and narrow size distribution, I(q) can be factorized as given below [11, 16, 17],
6 Synchrotron SAXS of Colloidal Suspensions 141
I(q) = NV
2

2
P(q)S(q) (6.14)
S(q) relates scattered intensity to the microstructure through the pair correlation
function, g(r), which is related to the probability of nding a particle at a distance r
from another particle [16],
S(q) = 1+4N


0
(g(r) 1)
sin(qr)
qr
r
2
dr (6.15)
Furthermore, in the limit q 0, S(0) = Nk
B
TK
T
, where K
T
is the compressibility
[16, 17]. Calculation of g(r) involves many-body correlations. The total correlation,
h(r) =g(r)1, is given in terms of the a priori unknown direct correlation function,
c(r), and indirect correlations by the OrnsteinZernike (OZ) integral equation [16].
Therefore, the calculation of c(r) involves further approximation which is called
a closure relation. The simplest case is the low-density limit (N 0) of the OZ
equation that leads to the mean spherical approximation (MSA),
c(r) =U(r)/k
B
T (6.16)
More accurate closure relations are obtained by including further terms in the den-
sity expansion [16, 17]. For example, the very commonly used PercusYevick (PY)
approximation is given by
c(r) = g(r)

1exp

U(r)
k
B
T

(6.17)
The advantage of PYclosure is that OZ equation can be solved analytically for short-
ranged interparticle potentials such as hard sphere, square well [1618]. For long-
ranged U(r), the hypernetted chain approximation (HNC) provides better results
but it needs to be solved numerically. Further improvements can be obtained by
thermodynamically self-consistent mixing schemes of the above closure relations
which have only numerical solutions. Once the c(r) is known, S(q) is given by
S(q) =
1
1Nc(q)
(6.18)
where c(q) is the Fourier transform of c(r). The complete analytical expression for
c(q) in the monodisperse case for hard-sphere potential [16], and with square-well
attraction [18, 19], or screened Coulomb repulsion [20, 21] can be found elsewhere.
The calculation of S(q) is further complicated in the case of polydisperse sys-
tems [11, 16, 17]. The complete separation of P(q) and S(q) is not possible and
instead the effective structure factor, S
M
(q), depends on the individual scattering
amplitudes, F(q).
S
M
(q) =
1
F
2
(q)

i, j
F
i
(q)F
j
(q)S
i j
(q) (6.19)
S
i j
(q) are the partial structure factors and
142 T. Narayanan
F
2
(q) =

i
x
i
F
2
i
(q) (6.20)
where x is the fraction of individual sizes. For polydisperse hard spheres, an analyt-
ical solution is available in the PY approximation [22]. The direct relationships of
S
M
(0) to K
T
and the maximum of S
M
(q) to N are not strictly valid for polydisperse
systems. For relatively low polydispersities, the calculation can be simplied by the
assumption that the position and the size of the particles are not correlated which
leads to the decoupling approximation [11, 17],
I(q) = NV
2

2
P(q)

1+
F(q)
2
F
2
(q)
(S(q) 1)

(6.21)
The quantity within the square brackets is an apparent structure factor and the S(q)
is calculated for the mean radius of the particles. The angular brackets denote the
average quantities over the size distribution. Analytical expression for Eq. (6.21) is
available for Schulz distribution [11]. The decoupling approximation works better
in repulsive systems but fails at high polydispersities [17]. Moreover, at high volume
fraction, it tends to overpredict S(q) at low q-values which makes the approximation
less suitable for modelling attractive systems.
A similar coupling between form and structure factors occurs even in the
monodisperse case for anisotropic particles [11, 16, 17]. The scattering amplitudes
and partial structure factors depend on the orientation of the particles. For small
anisotropies and low volume fractions, the interactions can be assumed to be in-
dependent of the orientation which leads to an analogous decoupling approxima-
tion [11, 17], as in the case of size polydispersity for spherical particles. Then the
averaging in Eq. (6.21) is performed over the orientation distribution of anisotropic
particles. The more general case of both size and shape polydispersities is a non-
trivial problem and difcult to model quantitatively. For spherical particles, the
effect is less pronounced in S(q) as compared to that in P(q) at low polydispersities.
Within the experimental errors, deviations from the monodisperse S(q) at least for
the rst maximum is less signicant below 5% polydispersity (e.g. hard spheres for

S
0.10.2).
Figure 6.4 presents the normalized intensities illustrating P(q) and S(q) for a
hard-sphere-like system consisting of sterically stabilized silica particles suspended
in dodecane. P(q) is a scaled tted function with Schulz size distribution obtained
with a very dilute sample (
S
10
4
). The solid line for
S
0.5 corresponds
to Eq. (6.14) with the normalized P(q), known value of (9.9 10
4
nm
2
)
and a polydisperse hard sphere S(q) (hard-sphere diameter,
HS
134 nm). For a
comparison the t with a monodisperse S(q) is also shown. In addition, N and
S
are constrained by the relation
S
=N
3
HS
/6 and the good agreement between I(q)
and S(q) ts demonstrate that the particle number density can be reliably deduced
from the absolute intensity (d/d). At high volume fractions PY solutions usually
overpredict the structure. In Fig. 6.4, the monodisperse and polydisperse S(q)s are
obtained using a reduced volume fraction,

=
S

2
S
/16 given by the Verlet
6 Synchrotron SAXS of Colloidal Suspensions 143
(b)
S
(
q
)
S
~ 0.5
PY-monodisperse
PY-effective
0.00 0.05 0.10 0.15
1
2
3
0
0
1
2
(c)
S
= 0.5
Polydispersity:
S
M
(
q
)
q (nm
1
)
0.0
0.05
0.1
0.2
10
2
10
1
10
2
10
3
10
4
10
5
(a)
Polydisperse
Monodisperse
Form factor
I
(
q
)

(
m
m

1
)
q (nm
1
)
~ 0.50
S
Fig. 6.4 (a) Normalized SAXS intensity depicting form and structure factors for sterically stabi-
lized silica particles in dodecane at 60

C. The scaled P(q) is a tted function to Eq. (6.9) with a


Schulz size distribution of mean core radius 65 nm and polydispersity 0.085. The solid and dot-
ted lines for
S
0.5 correspond to Eq. (6.14) with the same P(q), and using polydisperse and
monodisperse hard sphere S(q)s (
HS
134 nm), respectively. (b) S(q) obtained by the division
of I(q) by P(q) in (a) and the corresponding ts to monodisperse and polydisperse models. (c)
Comparison of effective S(q) for different polydispersities with
S
= 0.5 and
HS
= 130 nm
correction [16]. With increase of polydispersity, height of the main peak decreases
and the oscillations in S(q) gradually smooth out [see Fig. 6.4(c)]. In addition, at
low q the polydisperse function lies above that of the monodisperse case.
An alternative description of a polydisperse system is via the local monodisperse
approximation which assumes that individual particles are surrounded by particles
with similar sizes. In this way, Eq. (6.14) can be integrated over the size distribu-
tion as in Eq. (6.9) [8, 9]. While analytical methods are relatively easy to handle
for experimentalists, the structure and interactions in many colloidal systems are
more complex than the simple potentials involved in these models. In addition, the
intrinsic polydispersity of real systems further complicates the data analysis. An
alternate approach towards obtaining model-free real space information is the gen-
eralized indirect fourier transformation (GIFT) [23]. In this method, P(q) is model
independent but S(q) requires a model. For known shape of the particles, P(q) can
be evaluated for an arbitrary size distribution when analysing SAXS data from an
unknown system.
More recently, the polymer reference interaction site model (PRISM) originally
developed for polymer solutions and melts [24] has found to be very successful for
modelling interactions in binary and multicomponent colloidal systems (including
colloidpolymer mixtures). Within the PRISM integral equation approach, the poly-
dispersity is dealt by a multicomponent site model and the theory is applicable to
both charged and uncharged colloidal suspensions [25].
144 T. Narayanan
6.3 Applications of SAXS Methods
In this section a few complementary aspects of SAXS is presented. The techniques
described are primarily used for bulk studies and extend the range of q, resolution
in q and time, and allows limited contrast variation. In principle, XPCS allows to
access the corresponding equilibrium dynamics over this q-range [26]. The third-
generation synchrotron sources have played a key role in the development of these
complementary techniques in soft matter studies.
6.3.1 Ultra Small-Angle X-Ray Scattering
Typically the ultra small-angle X-ray scattering (USAXS) range covers sizes from
100 nm to several microns and above. USAXS is an alternative to light scattering
and microscopy for optically opaque samples. The high collimation of an undula-
tor beam readily permits accessing the lower end of the USAXS region (100 nm
1 m) by a pinhole SAXS instrument (as schematically depicted in Fig. 6.1) with
a long sample to detector distance. However, the Bonse and Hart scheme involving
channel-cut crystal analyser has several advantages [27]. The Bragg reection from
a thick crystal has a nite width (Darwin width,
D
) that depends on the order of
reection [6], and the parasitic background can be effectively curtailed by means of
a perfect channel-cut crystal. Typical beam divergence of an undulator beam is com-
parable to the width of the rocking curve (e.g. Si-111 22rad), and therefore the
loss of intensity with multiple reections is affordable. The measured rocking curve
is a superposition of the sample scattering and the reection curve of the analyser
crystal. Furthermore, the measured intensity can be directly transformed to d/d
without requiring a calibration standard [28]. In the case of colloidal suspensions,
the absolute scattered intensity can be used to deduce the particle number density
(Eq. (6.14)), provided is known from independent measurements as shown in
Fig. 6.4 [29].
In pinhole geometry, the intensity dynamic range is most often limited by the
detector point spread function which can also introduce certain peculiar artefacts
near the beamstop [30]. The BonseHart instrument provides unsmeared intensity
proles spanning over seven to eight orders of magnitude for a strongly scattering
sample [31]. For example, Fig. 6.5 displays the scattered intensity from a colloidal
system which can be transformed from repulsive to short-ranged attractive parti-
cles by a temperature-dependent absorption process [32]. In the attractive region,
particles aggregate to form droplet-like clusters. The peak around q 0.05nm
1
signies dense liquid-like packing within the clusters. When the same scattering
curve is measured by a pinhole instrument, the power-law scattering from the clus-
ters indicated by the shadowed window in the USAXS region will be nearly blocked
by the beamstop and the low q minimum will be inuenced by the detector point
spread function. This obscured information is crucial for the complete modelling of
I(q). The tted line corresponds to Eq. (6.14) with a hard-sphere structure factor
6 Synchrotron SAXS of Colloidal Suspensions 145
Fig. 6.5 USAXS froma silica colloidal suspension (R
S
60nm) in a binary mixture of 2,6-lutidine
and heavy water in the attractive and the repulsive states [32]. Attractive particles form droplet-
like colloidal clusters as schematically depicted in the inset. The high dynamic range covered by
the measurement allows distinguishing the different structural levels in the system. The continuous
line corresponds to Eq. (6.22) with polydisperse P(q), a monodisperse PercusYevick S(q), and
p = 4 and 1.2 m. The upper curve is displaced by a factor 3 for the sake of clarity
to which an additional DebyeBueche term (p = 4) [33] is added to describe the
scattering by the droplet-like clusters,
I(q) = NV
2

2
P(q)

S
PY
(q) +
I
M
(1+q
2

2
)
p/2

(6.22)
where I
M
and are proportional to the mean cluster mass and size, respectively [1].
The kinetically arrested states in short-ranged interacting colloids is a topic of
contemporary interest [34]. In this case, USAXS is a powerful method to probe the
subtle features in the microstructure of these systems. Figure 6.6 illustrates an ex-
ample of a thermally induced transition from hard-sphere repulsive to short-ranged
attractive interactions in a sterically stabilized colloidal system consisting of stearyl
silica particles in n-dodecane [29]. USAXS allows distinguishing the evolution in
S(q) corresponding to the transition from repulsive to attractive interactions and the
onset of clustering as the interparticle attraction (u) is progressively increased. Mod-
elling of the USAXS intensities provide parameters of u (depth and range) as well as
the structure of aggregates. The morphology of clusters in this case is characterized
by p 2.3 in Eq. (6.22). This corresponds to a fractal dimensionality (d
f
) 2.3,
suggesting the formation of a colloidal gel.
Another interesting application of USAXS is in the in situ studies of nucleation
and growth of aerosol particles [35]. In this case again USAXS provides access to
the size and morphology of aggregates and agglomerates. Fig. 6.7 shows the typical
146 T. Narayanan
10
8
10
7
10
6
10
5
10
4
10
3
10
2
10
3
10
2
10
1
Clustering (u/k
B
T2, 340 nm)
Attractive (u/k
B
T1)
Repulsive
I
(
q
)

(
m
m

1
)
q (nm
1
)
21.0
o
C (x10)
23.0
o
C (x3)
25.0
o
C
T
A
V
r
V
r
u
silica
silica

S
~ 0.2
Fig. 6.6 The evolution of the USAXS intensity in a short-range interacting colloidal system con-
sisting of stearyl silica particles in n-dodecane illustrating a thermally induced transition from
repulsive hard spheres to attractive spheres and the subsequent formation of colloidal clusters.
For better visibility, the absolute I(q) is multiplied by the factor indicated in the legend. The in-
set schematically depicts the transformation of the particles and the corresponding interparticle
potential above and below the transition temperature, T
A
scattering features of ame soot in an acetylene ame at two different heights above
the burner (HAB). Fits to unied scattering function (Eq. (6.13)) reveal compact
primary particles and their fractal aggregates [36]. The volume fraction of primary
particles is of the order of 10
6
. From the gas ow rate, HAB can be converted
10
3
10
3
10
2
10
1
q
4
q
2
10
2
10
1
10
0
10
1
10
2
10
3
10
4

Fig. 6.7 Combined USAXS and SAXS from an aggregating ame soot at two heights above the
burner, 10 and 20 mm corresponding to residence times of 27 and 54 ms, respectively. The power-
law regions elucidate the underlying morphology of spherical primary particles, their fractal ag-
gregates and agglomerates. The dashed and dotted lines represent the contribution from different
structural levels (Eq. (6.13))
6 Synchrotron SAXS of Colloidal Suspensions 147
to the residence time which is equivalent to a kinetic time. At 20 mm above the
burner, primary particles have reached their terminal size with R
g1
27nm and ag-
gregates R
g2
100nm with d
f
2. USAXS allows mapping of the growth kinetics
of primary particles and their aggregates as a function of residence time which pro-
vides a detailed understanding of the underlying mechanism [36]. While nucleation
and growth of primary particles can be probed by SAXS, distinguishing the onset
of aggregation is important in determining the terminal particle size distribution.
Moreover, in the case of ame-grown particles, the application of light scattering is
restricted by the optical emission from the ame.
With advanced instruments, resolution effects are not signicant when study-
ing particles of size 100 nm and polydisperse systems. However, limitations arise
in the case of larger particles with a narrow size distribution, especially when in-
vestigating their ordered states [37]. Figure 6.8 shows the scattering from a silica
colloidal glass (particle size 600 nm and polydispersity 2%), formed upon shear
melting a colloidal crystal, recorded with a BonseHart instrument. Both P(q) and
S(q) at low q-range have been signicantly smeared by the nite instrument resolu-
tion (q 0.001nm
1
). The low q minima and maxima in P(q) are well described
after the convolution by a gaussian resolution function, R(q),
I(q) =

R(q)
d
d
dq (6.23)
Here d/d is given by Eq. (6.14) with parameters indicated in the legend. The
repulsive interactions are modelled by effective hard-sphere parameters which are
larger than the R
S
and
S
deduced from P(q) and absolute level of I(q), respectively.
The convoluted S(q) is broader than the experimental data presumably due to the
particular history of the glassy sample.
I
(
q
)

(
m
m

1
)
10
2
10
1
Model convoluted ( q = 0.001 nm
1
)
Fig. 6.8 Resolution effects in the analysis of large particles with relatively narrow size distribution.
The low q part of P(q) and S(q) has been signicantly smeared by the nite instrumental resolution
(q 0.001nm
1
). S(q) is modelled by the effective hard-sphere parameters indicated in the
legend
148 T. Narayanan
6.3.2 Anomalous Small-Angle X-Ray Scattering
Until now the atomic scattering factor was assumed as a constant ( f Z) which is
valid when the incident X-ray energy (E) is removed from the atomic absorption
edge of all the constituent elements. Near the absorption edge f becomes a complex
function of E and not all electrons involve in the scattering process [38]. Above the
edge, some absorb the photon and the corresponding inner shell electron will be pro-
moted. This absorbed photon will be later emitted at a lower energy as uorescence.
The complex atomic scattering factor is given by
f (E) = f
0
+ f

(E) +i f

(E) (6.24)
f
0
Z except for a small relativistic correction at high photon energies [38]. The
imaginary part is responsible for absorption, and f

and f

are related through the


KramersKronig dispersion relation [38, 39].
f

=
1
r
e
hc


0
E
2

a
(E

)
E
2
E
2
dE

(6.25)
f

=

a
2r
e

(6.26)
where
a
is the atomic photoabsorption cross section which is related to mass ab-
sorption coefcient, , as
a
= A
W
/N
A
, A
W
is the atomic weight, N
A
is the Avo-
gadro number, h is the Planck constant and c is the velocity of light. The energy
dependence of f can be exploited to vary the contrast in SAXS for a large num-
ber of elements having an absorption edge in the range of 5 25 keV. This contrast
variation SAXS is known as anomalous SAXS (ASAXS). Figure 6.9(a) shows the
typical variation of f

and f

for rubidium near its K-edge.


For experiment, a special cell is required with calibration standards for inten-
sity (e.g. Lupolen or water) and energy (a solution containing known concentra-
tion of the anomalous atoms) installed besides the actual sample, preferably in an
in-vacuum set-up as schematically shown in Fig. 6.9(b). In this way, the detector
10 15 20
10
5
0
5
Rb 15.2 keV
f
'
f
''
f
(a ( ) b)
f
E (keV)
f
"
f
'
motorized syringe
capillary cell
calibrants
Rb
vacuum chamber
Fig. 6.9 (a) Atomic scattering factors ( f

and f

) of rubidiumnear its K-edge (15200 eV) depicting


the anomalous effect. (b) Schematic view of in-vacuum ow-through cell for in situ calibration of
absolute intensity and incident energy
6 Synchrotron SAXS of Colloidal Suspensions 149
efciency, the distance from the absorption edge and the uorescence contribution
to the intensity can be measured in situ. The motorized syringe could be used to
renew the sample during an energy scan for radiation-sensitive samples.
The potential of ASAXS was recognized decades ago [39] but it was mainly re-
stricted to metallic alloys [1]. The primary advantage of a synchrotron source in
ASAXS is that the energy can be varied continuously. The real and imaginary parts
of f ( f

and f

) as a function of E are available for most elements [38]. The K-edge


of certain intermediate elements (iron to strontium) and L-edges of some heavy
elements (e.g. gold, platinum or lead) are in a suitable energy range to perform
quantitative SAXS. However, the absorption edges of most relevant elements in soft
matter and biological materials (carbon calcium) are at lower energies where quan-
titative SAXS is less feasible due to strong absorption effect and radiation damage.
Nevertheless, in many practical situations, sodium or potassium can be replaced by
rubidium, magnesium or calcium by strontium, sulphur by selenium, chlorine by
bromine, etc. without dramatically altering the chemistry of the system. This al-
lowed ASAXS experiments on such systems but quantitative information remained
elusive as a result of the relatively low concentration of the ions and resulting small
changes in I(q) as a function of E. However, recent developments in high brilliance
SAXS has rectied these limitations, thereby permitting to obtain good intensity
statistics from very dilute and low contrast samples.
As mentioned before, for dilute non-interacting systems, S(q) 1, the measured
intensity is given by Eqs. (8.5) and (8.7). Near the absorption edge, part of (r)
becomes a complex function. Therefore,
(r)
M
=(r) =
0
(r) +n
R
(r)r
e
( f

(E) +i f

(E)) (6.27)
where
0
(r) is the non-resonant part or the usual contrast well below the absorp-
tion edge and n
R
is the number density of resonant ions. Now, F(q) can be decom-
posed to [40, 41]
F(q) = F
0
(q) +F
R
(q) (6.28)
with F
2
0
(q) being the non-resonant SAXS intensity measured far away from the
absorption edge and using Eq. (6.7)
F
R
(q) = 4r
e


0
( f

(E) +i f

(E))n
R
(r)
sin(qr)
qr
r
2
dr = ( f

(E) +i f

(E))N
R
(q)
(6.29)
where N
R
(q) is the product of Fourier transform of the spatial distribution of res-
onant atoms. The total scattered intensity by the macroions and the counterions is
given by
I(q) = NF(q)F

(q) = N(F
2
0
(q) +2f

F
0
(q)N
R
(q) +( f
2
+ f
2
)N
2
R
(q)) (6.30)
The anomalous effect appears as two different terms, the dominant cross-term in-
volving the resonant and non-resonant scattering amplitudes and a smaller self-
term due to the square of the resonant amplitude [39, 40]. The fact that f

(E)
150 T. Narayanan
0.0 0.1 0.2
10
2
10
1
10
0
10
1
I
(
q
)

(
m
m

1
)
q (nm
1
)
12460 eV
14800 eV
15184 eV
Fig. 6.10 Normalized SAXS intensity from a spherical polyelectrolyte brush consisting of a
polystyrene core, chemically grafted polyacrylic acid chains and Rb counterions as schematically
depicted in the inset. The lowering of I(q) with energy corresponds to the decrease of f

(E) of
Rb
+
ions as the absorption edge is neared [41]
decreases near the absorption edge implies that the scattered intensity decreases.
This is shown in Fig. 6.10 for the case of a spherical polyelectrolyte brush com-
posed of polystyrene core with radius 60 nm and a shell of grafted polyacrylic acid
chains of thickness 30 nm with Rb
+
counterions [41].
The challenge is to decompose the self-term (N
2
R
) whose Fourier transform di-
rectly yields the spatial distribution of the resonant atoms. Figure 6.11 shows
the separation of resonant and non-resonant terms in the case of the polyelec-
trolyte brush. From the individual scattering amplitudes, the corresponding elec-
tron density proles can be extracted as displayed in the inset. These electron
density proles directly reveal that the counterions are strongly correlated to the
macroion [41]. Figure 6.11 also illustrates the inherent difculties involved in per-
forming an ASAXS experiment, especially when N
R
(q) is relatively small. To ex-
tract the anomalous terms reliably, the relative accuracies of the measured I(q) at
different energies have to be very high (typically 0.1% or better). Furthermore, ab-
solute intensity scale is important to determine the spatial distribution of resonant
atoms. This means that precise calibration of incident ux and energy, sample trans-
mission and detector efciency over the required energy range is essential for an
ASAXS experiment [see Fig. 6.9(b)]. In the close vicinity of the absorption edge, the
nite energy resolution of the crystal monochromator (E/E 0.015% for Si-111)
introduces a smearing of f

(E) and an effective f

(E) has to be used in Eq. (6.30).


Furthermore, part of the energy spectrum can lie above the absorption energy result-
ing in an additional at background contribution from the uorescence which needs
to be subtracted. The uorescence contribution can be estimated from the measure-
ments done at two different energies above and below the absorption edge having
the same value of f

(E) [41].
6 Synchrotron SAXS of Colloidal Suspensions 151

Fig. 6.11 Decomposition of SAXS intensity into non-resonant (F


2
0
) and resonant (N
2
R
) terms in
the case of the polyelectrolyte brush depicted in Fig. 6.10. The inset presents the individual radial
electron density proles deduced from the corresponding scattering amplitudes [41]
The above example demonstrates that high-resolution ASAXS can reveal the ne
details of the counterion distribution in polyelectrolytes. In general, this method has
a great potential to investigate charged soft matter and biological systems [40].
6.3.3 Stopped-Flow Time-Resolved Small-Angle X-Ray Scattering
A direct implication of high brilliance for SAXS applications is the ability to
perform time-resolved experiments which are very valuable for probing the non-
equilibrium dynamics in soft matter [42]. In these studies, the system is driven out
of equilibrium by an externally imposed condition and the relaxation of the system
to the new quasi-equilibrium state is monitored as a function of time. This is differ-
ent from the equilibrium dynamics probed by dynamic scattering experiments but
both are related by uctuation dissipation. Time-resolved experiments can be per-
formed in two ways: (1) real-time experiments wherein the scattering is followed
continuously as a function of time and (2) stroboscopic experiments in which a
certain time window is chosen and the experiment is repeated to accumulate the re-
quired statistics. In the latter case, the time course is followed by shifting the time
window relying on the precise synchronization and temporal reproducibility of the
process. The synchronization of the physical process is achieved by ancillary tech-
niques collectively called the sample environment. The effective time resolution is
determined by the speed of the data acquisition as well as the ability to synchro-
nize the phenomenon (time scale over which the entire sample behaves like a single
entity) with the imposed condition.
The stopped-ow mixing technique is widely used for triggering kinetics in
the millisecond range by rapid change in concentration or pH [42]. Figure 6.12
152 T. Narayanan
mixer

1

2
hard stop
syringes
sink
capillary cell
Fig. 6.12 Schematic diagram of a stopped-ow device consisting of two syringes, one mixer and
a hard stop. After mixing, the sample is transferred to the observation capillary and the incident
beam probes the mixture at the indicated position. The deadtimes
1
and
2
correspond to mixing
and transfer to the beam position, respectively
schematically depicts a stopped-ow device consisting of two motor or pneumat-
ically driven syringes, a turbulent mixer and a hard stop. The data acquisition is
usually hardware synchronized with the movement of the syringes and the activa-
tion of the hard stop. Additionally, syringes and mixing chambers can be appended
to the ow circuit when needed to mix more than two liquids. The observation cell
is a thin-walled quartz capillary (10m) of diameter 1.52 mm. The key param-
eter is the dead time of the apparatus which is determined by the time of mixing
(
1
) and the transfer time between the mixer and the observation point (
2
) [42]. In
isotropic turbulence, for two liquids having water-like viscosities, with typical mixer
size l 100m and ow velocities, 10 m/s,
1
is in the range of 100200s [42].
The turbulent mixing ceases at the viscous dissipation range below which the mix-
ing occurs via diffusion. For this reason, the observation point is usually located a
few millimetres down from the mixing point. The corresponding transfer (aging)
time is of the order of a millisecond (e.g. for a distance of 5 mm and ow velocity of
5 m/s,
2
= 1ms). The total dead time will be larger for lower ow rates and higher
viscosities. Therefore, the practical limit for complete homogenous mixing is in the
range of several milliseconds and during the intervening period, the scattering is
dominated by concentration uctuations.
The spontaneous self-assembly of unilamellar vesicles is a topic of longstand-
ing interest [43]. The self-assembly process can be initiated by the rapid mixing
of two micellar solutions with different charges. Figure 6.13 displays an example
for real-time kinetic study of this self-assembly. In this case, equimolar amounts of
zwitterionic and anionic micelles (tetradecyldimethylamine oxide, M
1
, and lithium
peruorooctanoate, M
2
, respectively) were rapidly mixed and the scattering was
followed as a function of time [44, 45]. The SAXS patterns were recorded by a
two-dimensional detector with millisecond time resolution [46]. Figure 6.13 shows
the static scattering fromthe initial micelles and the rst observable state of disk-like
mixed micelles formed in the mixing process. The analysis of mixed micelle form
factor by a disk scattering function (Table 6.1) revealed their equimolar composition.
However, these disk-like micelles are unstable and they grow by an expo-
nential kinetics in a subsequent step spanning a few hundred milliseconds [44,
45]. Figure 6.14 depicts the growth of disk-like micelles and their closure to
6 Synchrotron SAXS of Colloidal Suspensions 153
10
5
10
1
10
0
10
4
10
3
10
2
10
1
I
(
q
)

(
m
m

1
)
q (nm
1
)
Fig. 6.13 Formation of disk-like mixed micelles following the rapid mixing of 50 mM solutions
of zwitterionic tetradecyldimethylamine oxide (M
1
) and anionic lithium peruorooctanoate (M
2
)
micelles. Inset is a cartoon of the transformation
form unilamellar vesicles. Here, the mixed micelles and unilamellar vesicles are
described by Eq. (6.9) with the disk and the shell scattering functions given in
Table 6.1. The driving force for the growth of disk-like micelles is their unfavourable
edge energy [47]. Above a critical size, the bending energy of the bilayers favours
closure of the disks to form unilamellar vesicles. The high time resolution enabled
to capture the transient intermediate states and to probe their dynamics in this multi-
step self-assembly process [44, 45].
10
1
10
0
10
4
10
3
10
2
10
1
10
0
10
1
10
2
10
3
I
(
q
)

(
m
m

1
)
q (nm
1
)
Fig. 6.14 Intermediate stages involved in the transformation of an equimolar mixture of an-
ionic/zwitterionic micelles to vesicles. The initial four curves correspond to disk-like mixed mi-
celles and their growth, and the two upper curves represent unilamellar vesicles. For clarity the
successive curves have been multiplied by a factor

10. Inset is a cartoon of the underlying trans-


formation
154 T. Narayanan
An alternative to stopped-ow mixing is a continuous ow device in which the
mixing and transfer are continued at a desired rate and the kinetics at time (
1
+
2
)
is probed as a function of distance from the mixing point and ow rate using a beam
with small cross section (see Fig. 6.12). Each kinetic time point is a static measure-
ment and thereby obtaining good intensity statistics. The dimensions of the mixing
stage and the associated time scales can be reduced in a microuidic environment
(especially the distance between the mixing and observation points) [48]. However,
within such devices the ow is laminar corresponding to a larger
1
and poor mix-
ing efciency [42]. The inuence of incomplete mixing on the kinetic pathway of
the reaction is a signicant problem at short time scales (below a millisecond) in
any mixing device for samples with water-like viscosities. The mixing deadtime in-
creases correspondingly for more viscous systems and it is important to check the
precise temporal reproducibility of the kinetics under different mixing conditions
before reaching any quantitative conclusion.
In general, for strong scatterers, the time resolution in the millisecond time range
is primarily limited by the detector. For weakly scattering systems, both the detec-
tor and the sample scattering power limit the time resolution below 10 ms. In this
case, stroboscopic experiments become more appropriate (e.g. the continuous-ow
mixing scheme described above).
6.4 Summary and Outlook
The application of SAXS and related techniques to colloidal systems has steadily
advanced over the past decades. This chapter presented a few representative and
relatively simple examples to illustrate the different features of SAXS without an
extensive review of the literature on the application of SAXS in colloid science. The
signicant improvement in sensitivity and resolution of SAXS techniques realized
at modern synchrotron sources offer new possibilities in the studies of structure and
dynamics of colloidal systems. Examples include time-resolved SAXS, ASAXS,
USAXS, XPCS, etc. As a result, real-time studies on very dilute systems, charge dis-
tribution in ionic systems, large-scale structures and dynamics in optically opaque
samples can be investigated in great detail. In addition, the scattering experiments
are often combined with a variety of thermophysical and rheological techniques.
Therefore, systems driven out of equilibrium or subjected to controlled shear stress
can be studied in situ [42, 49]. Despite the advances in imaging techniques, scatter-
ing experiments remain essential for investigating colloidal samples. Most powerful
approach is to combine scattering and imaging methods in a complementary fashion.
Kinetic studies in the sub-millisecond regime still remain largely unexplored.
With recent advances in detector technology, the microsecond range is becoming
accessible for real-time studies. To exploit these developments, appropriate trigger-
ing methods are also required especially that the soft matter systems are dominated
by entropy which makes the synchronization in the fast time scale non-trivial. The
increasing brightness of modern X-ray sources continues to improve the quality of
6 Synchrotron SAXS of Colloidal Suspensions 155
the scattering data as long as the sample can withstand the high photon ux without
radiation damage. A comprehensive understanding of the radiation damage is essen-
tial to advance soft matter research using new generation X-ray sources which are
expected to be many orders of magnitude brighter and provide coherent X-ray beam
with sub-micrometer beam size. Further advances in the application of scattering
techniques to colloidal systems critically depend on the ability to model polydis-
perse systems with complex interactions.
Acknowledgments The experimental results presented in this chapter involved collaborations
with many colleagues, especially M. Ballauff, G. Beaucage, N. Dingenouts, M. Gradzielski,
P. Panine, F. Pignon, M. Sztucki, and T. Weiss. ESRF is acknowledged for the provision of syn-
chrotron beam time.
References
1. Brumberger H. (Ed.) Modern Aspects of Small-Angle Scattering. Kluwer Academic,
Dordrecht (1995). 133, 134, 136, 137, 139, 145, 149
2. Lindner P., Zemb T. (Eds.) Neutrons, X-Rays and Light: Scattering Methods Applied to Soft
Condensed Matter. Elsevier, Amsterdam (2002). 133, 134
3. Guinier A., Fournet G. Small-Angle Scattering of X-rays. Wiley, New York (1955). 134, 138
4. Glatter O., Kratky O. (Eds.) Small-Angle X-ray Scattering. Academic Press, London (1982). 134, 137, 139
5. Feigin L.A., Svergun, D.I. Structure Analysis by Small-Angle X-ray and Neutron Scattering.
Plenum Press, New York (1987). 134
6. Warren B.E. X-Ray Diffraction. Dover, New York (1990). 135, 144
7. Narayanan T. Synchrotron small-angle X-ray scattering, in Soft-Matter Characterization,
Chap. 17, p 899. R. Borsali and R. Pecora (Eds.), Springer, Heidelberg (2008). 135, 136
8. Pedersen J.S. Adv. Colloid Interface Sci., 70, 171 (1997). 137, 138, 143
9. Lindner P., Zemb T. (Eds.) Neutrons, X-Rays and Light: Scattering Methods Applied to Soft
Condensed Matter, p. 391, Elsevier, Amsterdam (2002). 137, 138, 143
10. Dingenouts N., Bolze J. Poetschke D., Ballauff M., Adv. Poly. Sci. 144, 1(1999). 138
11. Kotlarchyk M., Chen S.-H. J. Chem. Phys. 79, 2461 (1983). 138, 140, 141, 142
12. Beaucage G. J. Appl. Cryst. 28, 717 (1995). 139, 140
13. Beaucage G., Kammler H.K., Pratsinis S.E. J. Appl. Cryst. 37, 523 (2004). 139
14. Holt C., de Kruif C.G., Tuinier R., Timmins P.A. Colloids Surf. A, 213, 275 (2003). 140
15. Pignon F., Belina G., Narayanan T., Paubel X., Magnin A., Gesan-Guiziou G. J. Chem. Phys.
121, 8138 (2004). 140
16. Klein R., DAguanno B. In Light Scattering: Principles and Development, p. 30, W. Brown
(Ed.), Clarendon Press, Oxford (1996). 140, 141, 142, 143
17. Kaler E.W. In Modern Aspects of Small-Angle Scattering, p. 329, H. Brumberger (Ed.),
Kluwer Academic, Dordrecht (1995). 140, 141, 142
18. Menon S.V.G., Manohar C., Srinivasa Rao K. J. Chem. Phys. 95, 9186 (1991). 141
19. Chen W.-R., Chen S.-H., Mallamace F., Phys. Rev. E 66, 021403 (2002). 141
20. Hayter J.B., Penfold J., Mol. Phys. 42, 109 (1981). 141
21. Hansen J.-P., Hayter J.B. Mol. Phys. 46, 651 (1982). 141
22. Vrij A. J. Chem. Phys. 71, 3267 (1979). 142
23. Fritz G., Bergmann A., Glatter O. J. Chem. Phys. 113, 9733 (2000). 143
24. Schweizer K.S., Curro J.G. Adv. Chem. Phys. 98, 1 (1997). 143
25. Harnau L., Hansen J.-P. J. Chem. Phys. 116, 9051 (2002). 143
156 T. Narayanan
26. Gr ubel G., Stephenson G.B., Gutt C., Sinn H., and Tschentscher T. Nucl. Instr. Meth. Phys.
Res. B 262, 357 (2007). 144
27. Bonse U., Hart M., Appl. Phys. Lett. 7, 238 (1965). 144
28. Long G.G., Jemian P.R., Weertman J.R., Black D.R., Burdette H.E., Spal R. J. Appl. Cryst.
24, 30 (1991). 144
29. Sztucki M., Narayanan T., Belina G., Moussaid A., Pignon F., Hoekstra H. Phys. Rev. E 74,
051504 (2006). 144, 145
30. Pontoni D., Narayanan T., Rennie A.R. J. Appl. Cryst. 35, 207 (2002). 144
31. Sztucki M., Narayanan T. J. Appl. Cryst. 40, S459 (2007). 144
32. Pontoni D., Narayanan T., Petit J.-M., Gr ubel G., Beysens, D. Phys. Rev. Lett. 90, 188301
(2003). 144, 145
33. Debye P., Anderson H.R., Brumberger H. J. Appl. Phys. 28, 679 (1957). 145
34. Zaccarelli E. J. Phys. Condens. Matter 19, 323101 (2007). 145
35. Beaucage G., Kammler H.K., Strobel R., Mueller R., Pratsinis S.E., Narayanan T. Nat. Mater.
3, 370 (2004). 145
36. Sztucki M., Narayanan T., Beaucage G. J. Appl. Phys. 101, 114304 (2007). 146, 147
37. Petukhov A.V., Thijssen J.H.J., t Hart D.C., Imhof A., van Blaaderen A., Dolbnya I.P.,
Snigirev A., Moussaid A., Snigireva I. J. Appl. Cryst. 39, 137 (2006). 147
38. Thompson A.C., Vaughan D. (Eds.) X-Ray Data Booklet. LBNL, University of California,
Berkeley ( 2001). 148, 149
39. Stuhrmann H.B. Adv. Polym. Sci. 67, 123 (1985). 148, 149
40. Ballauff M., Jusi A. Colloid Polym. Sci. 284, 1303 (2006). 149, 151
41. Dingenouts N., Patel M., Rosenfeldt S., Pontoni D., Narayanan T., Ballauff M. Macro-
molecules 37, 8152 (2004). 149, 150, 151
42. Panine P., Finet S., Weiss T., Narayanan T. Adv. Colloid Interface Sci. 127, 9 (2006). 151, 152, 154
43. Gradzielski M. J. Phys. Condens. Matter. 15, R655 (2003). 152
44. Weiss T., Narayanan T., Wolf C., Gradzielski M., Panine P., Finet S., Helsby W. Phys. Rev.
Lett. 94, 038303 (2005). 152, 153
45. Weiss T., Narayanan T., Gradzielski M. Langmuir 24, 3759 (2008). 152, 153
46. Lewis R.A., Helsby W.I., Jones A.O., Hall C.J., Parker B., Sheldon J., Clifford P., Hillen M.,
Sumner I., Fore N.S., Jones R.W.M., Roberts K.M. Nucl. Instrum. Methods Phys. Res. A 392,
32 (1997). 152
47. Shioi A. Hatton T.A. Langmuir 18, 7341 (2002). 153
48. Atencia J., Beebe D.J. Nature 437, 648 (2005). 154
49. Panine P., Gradzielski M., Narayanan T. Rev. Sci. Instrum. 74, 2451 (2003). 154

You might also like