ArticlePDF Available

X-Ray Photon Correlation Spectroscopy Study of Brownian Motion of Gold Colloids in Glycerol

Authors:

Abstract and Figures

We report x-ray photon correlation spectroscopy studies of the static structure factor and dynamic correlation function of a gold colloid dispersed in the viscous liquid glycerol. We find a diffusion coefficient for Brownian motion of the gold colloid which agrees well with that extrapolated from measurements made with visible light, but which was determined on an optically opaque sample and in a wave-vector range inaccessible to visible light.
Content may be subject to copyright.
VOLUME 75, NUMBER 3 PHYSICAL REVIEW LETTERS 17JULY 1995
X-Ray Photon Correlation Spectroscopy Study of Brownian Motion of Gold Colloids in Glycerol
S.B. Dierker,1R. Pindak,2R.M. Fleming,2I.K. Robinson,3and L. Berman4
1Department of Physics, University of Michigan, Ann Arbor, Michigan 48109-1120
2AT&T Bell Laboratories, 600 Mountain Avenue, Murray Hill, New Jersey 07974-0636
3Department of Physics, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801-3080
4National Synchrotron Light Source, Brookhaven National Laboratory, Upton, New York 11973-5000
(Received 30 November 1994; revised manuscript received 5 May 1995)
We report x-ray photon correlation spectroscopy studies of the static structure factor and dynamic
correlation function of a gold colloid dispersed in the viscous liquid glycerol. We find a diffusion
coefficient for Brownian motion of the gold colloid which agrees well with that extrapolated from
measurements made with visible light, but which was determined on an optically opaque sample and in
a wave-vector range inaccessible to visible light.
PACS numbers: 61.10.Lx, 05.40.+j, 42.25.Kb, 82.70.Dd
Photon correlation spectroscopy (PCS) probes the dy-
namics of a material by analyzing the temporal correla-
tions among photons scattered by the material. Visible
light PCS has proven to be an indispensible technique
for studying the long wavelength hydrodynamics of flu-
ids, including simple liquids, liquid mixtures, liquid crys-
tals, polymers, and colloids [1]. However, visible PCS
cannot probe the short wavelength dynamics of materi-
als or opaque materials at all. The new field of x-ray PCS
(XPCS) offers an unprecedented opportunity to extend the
range of length scales over which a material’s low fre-
quency (10–3 to 106Hz) dynamics can be probed down to
interatomic spacings. It is clear that demonstration of the
ability to make dynamic XPCS measurements would be a
major step forward in this nascent field.
Many of the important problems in the low frequency
dynamics of condensed matter systems for which XPCS
should be uniquely suited arise in disordered materials, in
particular, liquids. These include studies of the dynamic
structure factor of liquids down to interatomic length
scales, density fluctuations in liquids undergoing a glass
transition, internal conformational dynamics and reptation
in polymers, and equilibrium concentration fluctuations
in polymer blends near a phase separation critical point.
Here we report results of equilibrium dynamic measure-
ments on a disordered system using XPCS. Specifically,
we measured the diffusion coefficient for Brownian mo-
tion of gold colloid particles dispersed in glycerol. Our
results unequivocally demonstrate the feasibility of the
XPCS technique, as well as illustrate the crucial impor-
tance of both matching the longitudinal coherence length
of the x rays to the experimental requirements and the
tremendous benefit of utilizing area detectors in XPCS
measurements.
While the principles of XPCS have been known for
decades, the very low flux of coherent x rays available
with previous sources has, until now, precluded its appli-
cation as a practical technique. The critical development
which has now made XPCS a feasible technique is the
use of insertion devices at second and third generation
synchrotron sources. The work reported here was done
on the wiggler beam line X25 at the Brookhaven National
Synchrotron Light Source (NSLS).
Previous work has consisted of measurements of the
static speckle patterns resulting from the structure of an-
tiphase domains in Cu3Au [2] or from grazing incidence
scattering from gold-coated polymer films of inhomoge-
neous thickness [3]. Identification of equilibrium critical
fluctuations in Fe3Al with XPCS has also been reported
very recently [4]. These studies benefited either from
the tremendous gain in scattering cross section of Bragg
scattering from essentially well-ordered materials [2,4] or
from the large reflectivity at grazing incidence [3]. In
contrast, the scattering typical of highly disordered mate-
rials, such as liquids or glasses, is dramatically weaker.
To demonstrate the difficulty of a coherent x-ray beam
by liquids, we estimate its average scattering efficiency.
For an incident x-ray beam of cross-sectional area Aiand
power P0(photonsysec) incident on a sample containing
Ntot uncorrelated atoms of charge Z, the fractional detected
count rate sPyP0dscattered into the solid angle DV is [5]
PyP0r2
0NtotZ2DVyAi. Here r0e2ymc22.8 3
10213 cm is the classical Thomson electron radius. In
terms of the density r, molecular weight M, Avogadro’s
number NA, and the scattering volume V, we have Ntot
rVNAyM. We take VAiymfor a sample of thickness
t1ym, where mrmmis the adsorption length and
mmis the mass absorption coefficient, and make use of the
approximate relationship [5] mmøCZl3NAyM, where C
is a constant. In a speckle experiment, the detector solid
angle of a single speckle is DV øl2yAi. Altogether this
gives PyP0ø8310226ZyClAi. This demonstrates the
benefit of having a high source brightness, i.e., P0yAi.
Aiis set by the transverse coherence length of the x-
ray beam. At X25, this requires a pinhole diameter of
5mm, for which Ai,20 mm2. For the case of liquid Au,
Z79, and taking l1.55 Å at 8 keV, C261 cm–1
for Au at 8 keV, and P0ø43107photonsysec at X25
with DEyE1.5%, we estimate the impracticably small
count rate of P,331024cps.
0031-9007y95y75(3)y449(4)$06.00 © 1995 The American Physical Society 449
VOLUME 75, NUMBER 3 PHYSICAL REVIEW LETTERS 17JULY 1995
However, if one had a liquid of Npgold particles, each
with naatoms (whether crystalline or amorphous, regard-
less of orientation), such that NpnaNtot, then the aver-
age scattering in the forward direction would be the sum
of the coherent scattering from Npparticles, i.e., PyP0~
Npn2
aNtotna, instead of Ntot. The qdependence of the
scattering will be that of the particle form factor [5] FsqRd
for a sphere of radius R(in the dilute gas limit at low q,
the structure factor will be unity),
FsqRd3hsinsqRd2qR cossqRdj
sqRd32
.(1)
For example, for a Au particle with R200 Å, na,
23106, giving a sufficiently large P,600 photonsysec.
A constraint of this approach is that one must work at
very low q, although not so low as to be dominated by the
tail of the Fraunhofer diffraction of the main beam. We
estimate that measurements should be possible at least over
the range 131023,q,331022Å–1. This extends
far beyond the upper qrange of visible light scattering, for
which qmax ,431023Å–1.
Monodisperse gold sols were prepared by reducing
HAuCl4with Na3-citrate [6]. The hydrodynamic radius
RHof the resultant gold colloid particles was determined
to be 335 Å for a small diluted aliquot of the resulting
sol using visible PCS [1]. A cumulant analysis of the dy-
namic correlation functions indicated a polydispersity of
,10%. The hydrophobic gold sol was stabilized against
aggregation by absorbing cold water fish skin gelatin onto
the surface of the particles [7]. RHand the hydrodynamic
polydispersity were remeasured for the stabilized sol and
determined to be 425 Å and 20%, respectively. The in-
creased magnitude of these quantities presumably reflects
the increased drag and statistical size fluctuations of the
“fuzzy” molecular coat covering the spheres.
The as-prepared sol had a gold volume fraction of
2.7 31024%, whereas the XPCS measurements required
a sample with a gold volume fraction of ,1%.At
this concentration, the x-ray absorption of the gold and
the glycerol are approximately equal. Thus, 450 ml
of sol was concentrated by centrifuging, aspirating out
the supernatant, and resuspending in 1.5 ml H2Oby
ultrasonicating. The concentrated sol was remeasured
with visible PCS, and RHand the polydispersity were
determined to be 590 Å and 26%, respectively. The sol
was then centrifuged and the supernatant was aspirated
and resuspended by ultrasonicating in 0.10 ml glycerol
to which 0.4MNaCl had been added in order to screen
out Coulomb interactions between the spheres. This was
repeated twice more. The end result was a concentration
enhancement of 4000:1, giving an estimated gold volume
fraction of 1.3% in a 0.1 ml sol. This concentration was
confirmed by x-ray absorption measurements. We note
that at this high concentration the sample is completely
opaque to visible light, and thus could not be studied with
visible PCS. The 1% Au sample was placed in a 1.0 mm
diameter hole drilled in a 0.6 mm thick aluminum plate,
i.e., about one absorption length thick at 8 keV.
A laterally coherent beam was prepared by passing the
x rays through a 5 mm diameter pinhole. The longitudi-
nal coherence length lcoh of the x-ray beam is given by
lslyDldand was set by using a two crystal WyB4C multi-
layer monochromator, giving lcoh ,100 Å. This resulted
in an incident coherent x-ray flux of 43107photonsysec.
In a transmission geometry, the maximum path length dif-
ference (PLD) is given by 2hsinutanu, where his the
sample thickness. Thus, the requirement [1] that PLD
,lcoh is satisfied up to scattering angles 2uof 8 mrad, cor-
responding to q3.2 31022Å–1. We also attempted
measurements using a Si(220) monochromator, for which
lcoh 2mm and the incident coherent flux was only
,23105photonsysec. Data collected under those con-
ditions were of significantly poorer quality than those re-
ported here. However, small-angle scattering data col-
lected under incoherent illumination conditions using the
Si(220) monochromator were of good quality, and all such
data presented here (Figs. 1–3) were collected using this
monochromator. This underscores the crucial importance
of not having lcoh any greater than required by PLD in
order to have the maximum intensity available for the
experiment.
X-ray detectors used included a scintillation detector
and a two-dimensional charged couple device (CCD)
area detector operated in the direct detection mode. The
CCD detector was custom-built and used a Kodak KAF-
1400 CCD, which is a front illuminated device having a
512 3768 array of 9 mm square pixels.
Figure 1 shows the measured scattering intensity, both
with and without the colloid. The strong signal centered at
q0is the direct beam transmitted through the sample.
The background scattering without sample is due to small
angle scattering from the slits and pinholes in the beam
line. The overall shape and intensity of the colloid scat-
tering qualitatively matches that expected from Eq. (1).
Measurements of the static structure factor were also
made with the CCD area detector. Figure 2 shows a plot
of a CCD image taken with an 80 mm front pinhole.
The channel cut through the center of the plot is due
to the shadow of a 200 mm diameter gold wire used to
block the main beam, which would have saturated the
FIG. 1. Ssqdas measured with incoherent x rays by scanning
an 80 mm pinhole followed by a scintillation detector. The
incident beam diameter was 200 mm. One contribution to the
background, a Kapton window, is indicated.
450
VOLUME 75, NUMBER 3 PHYSICAL REVIEW LETTERS 17JULY 1995
FIG. 2. Log plot of scattering intensity from the colloid
collected with the CCD camera.
detector. The CCD image was obtained by summing ten
1 sec exposures in order to extend the CCD’s dynamic
range. In comparison, each line scan with the pinhole-
scintillation detector combination over a similar qrange
took more than an hour.
A comparison of the structure factor measured in the
scintillation detector scans with that deduced from radial
averages of the CCD image is shown in Fig. 3. The
data have been scaled to coincide at q,0.015 Å–1.
The scintillation detector and CCD scans show good
general agreement apart from differences in the amount
of forward scatter around the main beam between the two
measurements.
In order to determine the size of the protected colloid
spheres and the degree of polydispersity, the scintillation
detector data shown in Fig. 3 were fit by a model
consisting of a Gaussian distribution of sphere scattering
radii, with each sphere according to Eq. (1), along with
FIG. 3. Comparison of Ssqdderived from scintillator or
pinhole scans with that from the CCD image in Fig. 2. Also
shown are the results of a fit of a model for Ssqdbased on
Eq. (1) and a Gaussian distribution of particle radii.
a constant background. The results indicate that the
mean radius was 270 Å and the standard deviation of
the distribution was 57 Å. These numbers agree well
with the mean hydrodynamic radius and polydispersity
measured on dilute aqueous solutions of the colloid with
visible PCS before protecting them with gelatin. This
suggests that the increased RHmeasured for the protected
colloid results from the added hydrodynamic drag of
the macromolecular coating. The further increase in RH
upon centrifuging presumably results from changes in the
molecular conformation of the gelatin molecules and not
from aggregation of the spheres.
We next turn to consideration of the Brownian motion
dynamics of this system. The Au colloid particles form
a dilute gas having diffusively relaxing concentration
fluctuation with a corresponding diffusion coefficient
given by [1] DkbTy6phRHand a relaxation time at
wave vector qof t1yDq2. Glycerol is a prototypical
glass former with a glass transition at Tg286 ±
C, at
which the shear viscosity happroaches ,1013 P [8].
The dynamic correlation function gstdis defined as
gstdknstdns0dlykns0dl2,(2)
where kl represents a time average and nstdis the
detected photon count at time t.gstdshould have a
maximum value of 2 at t0, decaying to 1 at infinite
time. Spatial averaging of the speckles by the detector
can reduce the t0intercept of gstd.
A dramatic increase in collection efficiency was ob-
tained by using the CCD detector to measure the scatter-
ing for 6750 pixels simultaneously. Since the dynamics
depend only on the magnitude of q, we can average the
autocorrelation functions measured in all of the pixels in a
band of qvalues. This is equivalent to performing an en-
semble average over pixels, as well as a time average for
each pixel, and should reduce the time needed to measure
the correlation functions with good statistics by the num-
ber of pixels averaged over [9]. With 6750 pixelsyimage,
the reduction is substantial.
Specifically, we recorded 1920 images of the inten-
sity in two separate 90±arcs corresponding to q’s of
3.3 31023and 5.5 31023Å–1, with widths of 10% of
their average q’s and containing 1750 and 5000 pixels,
respectively (see Fig. 4). The larger qis well above the
maximum value of ,431023Å–1 attainable by visible
PCS. The images were taken once per second for a to-
tal duration of 32 min, with the exact time of exposure
recorded for each image. A total of ,3.4 3106pho-
tons were detected, corresponding to an average count rate
of 0.36 photonsypixel sec. This compares well with the
value of 0.68 derived from our earlier estimate when the
sample transmission (0.35), the ratio of the area of a pixel
to that of a speckle (0.13), the estimated quantum effi-
ciency of the CCD detector (0.05), and Eq. (1) (0.5) are
taken into account.
The time autocorrelation function of each pixel was then
calculated and the ensemble average of the resulting 1750
451
VOLUME 75, NUMBER 3 PHYSICAL REVIEW LETTERS 17JULY 1995
FIG. 4. Autocorrelation functions of colloid scattering inten-
sity as collected with the CCD camera for two different wave
vectors along with single exponential fits and relaxation times,
as indicated. The inset schematically depicts the scattering
“halo,” the beam stop, and the subarrays which were ensem-
ble averaged over. Curve (b) is offset by 0.15 for clarity.
or 5000 correlation functions calculated. The results are
shown in Fig. 4, which is plotted as gstd21. The large
signal-to-noise ratio of the data in Fig. 4 can leave no doubt
that they correspond to dynamic x-ray scattering from
colloid concentration fluctuations. The large amplitude of
the decays is also as expected considering the small size of
the pixels relative to the coherence area for the scattering.
Also shown in Fig. 4 are fits to the data of single
exponential relaxations. They give characteristic decay
times of 43.1 and 24.1 sec for q3.3 31023and 5.5 3
1023Å–1, respectively. The measurements were made at
a temperature of 229 ±
C. The viscosity of pure glycerol
[8] should be 5.5 3103P at this temperature. Thus,
the nominal expected relaxation times are 61 and 22 sec,
which are in good agreement with the measurements.
The deviation of the relaxation rates from a precise
q2dependence may be an indication of the onset of
nonhydrodynamic effects at large wave vector [10].
While these measurements clearly establish the feasi-
bility of XPCS studies, more experiments are called for
in order to fully understand the dynamics of this colloid.
For example, nonhydrodynamic effects at large wave vec-
tor [10] as well as nonlinear relaxation effects predicted
by mode coupling theories [11] near the glass transition
should be fruitful areas of future study with this new
technique.
In conclusion, we have made small angle x-ray scat-
tering measurements of the static structure factor of an
optically opaque gold colloid dispersed in glycerol. We
also reported a determination of dynamic correlation func-
tions and the diffusion coefficient for the Brownian mo-
tion of the colloid particles, at a wavelength inaccessible
to visible PCS, by using the new technique of XPCS. Our
results demonstrate the importance of using a multilayer
monochromator, or just the natural bandwidth of undula-
tor radiation (,1%), in situations where a large lcoh is not
required, such as surface diffraction or small-angle scat-
tering experiments, as well as the benefits of ensemble av-
eraging with CCD area detectors in XPCS measurements.
The present results clearly show that XPCS is a fea-
sible technique and can have wide application to cer-
tain classes of systems, especially complex fluids and
surfaces, even using existing 2nd generation synchrotron
x-ray sources such as the NSLS. This is particularly
true given several significant forthcoming enhancements:
(1) Optimization of the use of an area detector for making
ensemble measurements of gstd:3101–102. (2) Beam-
line optics which utilize the full coherent output of the
undulator in both the vertical and horizontal directions:
310. (3) Introduction of undulator sources such as the
Prototype Small Gap Undulator recently installed at the
NSLS: 310. These will enhance the data collection effi-
ciency by 103and 104, relative to the current experiment.
Those classes of problems demanding the utmost coherent
incident x-ray flux will require the 3rd generation undula-
tor sources at the European Synchrotron Radiation Facility
and the Advanced Photon Source: 310 2–103, for a total
gain of 104–106, relative to the current experiment. When
these developments come to fruition, XPCS will realize its
full potential as a unique and important new technique.
It is a pleasure to acknowledge the assistance and
advice of Christoph Schmidt on the preparation of the
gold colloids. Robert MacHarrie assisted in the x-
ray measurements. S.B. D. acknowledges support under
NSF Grant No. DMR 92-17956 and also the donors
of the Petroleum Research Fund, administered by the
American Chemical Society, for the partial support of this
research under Grant ACS-PRF No. 26389-AC9. I. K. R.
acknowledges support under NSF Grant No. DMR 93-
15691. The NSLS is supported by the U.S. Department
of Energy under Contract No. DE-AC02-76CH00016.
[1] See, for example, Benjamin Chu, Laser Light Scattering:
Basic Principles and Practices (Academic Press, San
Diego, 1991), 2nd ed.
[2] M. Sutton et al., Nature (London) 352, 608 (1991).
[3] Z.H. Cai et al., Phys. Rev. Lett. 73, 82 (1994).
[4] S. Brauer et al., Phys. Rev. Lett. 74, 2010 (1995).
[5] G.B. E. Warren, X-Ray Diffraction (Addison-Wesley,
New York, 1968).
[6] G. Frens, Nature (London) 241, 20 (1973).
[7] Colloidal Gold: Principles, Methods, and Applications,
edited by M.A. Hayat (Academic Press, San Diego,
1989).
[8] M.L. Williams et al., J. Phys. Chem. 77, 3701 (1955).
[9] Apollo P.Y. Wong and P. Wiltzius, Rev. Sci. Instrum. 64,
2547 (1993).
[10] W. Hess and R. Klein, Adv. Phys. 32, 214 (1983).
[11] J. Jackle, J. Phys. Condens. Matter 1, 267 (1989).
452
... fastest dynamics [9,24,25]. After more than two decades of development, XPCS has been extensively applied to condensed systems [7,9], including colloidal suspensions [23,26], polymers [27,28], metallic glasses [29,30], proteins [31][32][33], and strongly correlated materials [10,34,35]. ...
... An average size of R SEM = 524.4 nm and dispersion of p = 7.2% were obtained [26]. The volume fraction of silica in the stock solution was 1% . ...
... Here, represents the speckle contrast, which depends on the coherence of the incident beam, optical layout of the beamline and pixel size of the detector. In this study, the diluted colloidal suspension was expected to exhibit Brownian motion [26,51,52], such that Combining Eqs. 4 and 5, the correlation equation can be simplified to: Γ(q) = 1∕ c (q) represents the relaxation rate, which is the reciprocal of the characteristic relaxation time c (q): ...
Article
X-ray photon correlation spectroscopy (XPCS) has emerged as a powerful tool for probing the nanoscale dynamics of soft condensed matter and strongly correlated materials owing to its high spatial resolution and penetration capabilities. This technique requires high brilliance and beam coherence, which are not directly available at modern synchrotron beamlines in China. To facilitate future XPCS experiments, we modified the optical setup of the newly commissioned BL10U1 USAXS beamline at the Shanghai Synchrotron Radiation Facility (SSRF). Subsequently, we performed XPCS measurements on silica suspensions in glycerol, which were opaque owing to their high concentrations. Images were collected using a high frame rate area detector. A comprehensive analysis was performed, yielding correlation functions and several key dynamic parameters. All the results were consistent with the theory of Brownian motion and demonstrated the feasibility of XPCS at SSRF. Finally, by carefully optimizing the setup and analyzing algorithms, we achieved a time resolution of 2 ms, which enabled the characterization of millisecond dynamics in opaque systems.
... g 2 (, Q) = hI(t, Q) I(t + )i t / hI 2 (t, Q)i t . For the simplest XPCS measurements, the beam on the sample is fixed and the statistics of g 2 can be greatly enhanced by binning detector pixels over Q and averaging over t assuming that the dynamics are stationary during intensity collection (Dierker et al., 1995). For more details about XPCS theories and experimental setup, readers can refer to Grü bel et al. (2008) and Shpyrko (2014). ...
... The versatility of XPCS, combined with full compatibility with in situ sample environments and wide temporo-spatial coverage, has led to unique insight into the collective atomic motion in metallic glasses (Evenson et al., 2015;Giordano & Ruta, 2016) and supercooled liquid alloys (Ruta et al., 2020), arrested dynamics of nanoparticles in polymer matrices (Dierker et al., 1995;Senses et al., 2017;Yavitt et al., 2021), liquid-liquid phase separation in protein (Begam et al., 2021) and micelle suspensions (Sheyfer et al., 2020), and the mesoscopic structure dynamics in both abrasive manufacturing such as ion-beam patterning (Myint et al., 2021) and additive manufacturing such as 3-D printing (Lin et al., 2021), where the results common to all these XPCS applications are intensity correlation functions in the form of multidimensional spectra over time and reciprocal space. However, generating and interpreting XPCS results can be very challenging due to the multi-dimensional and statistical nature of the data that require efficient reduction and sophisticated visualization tool suites. ...
Article
Full-text available
pyXPCSviewer, a Python-based graphical user interface that is deployed at beamline 8-ID-I of the Advanced Photon Source for interactive visualization of XPCS results, is introduced. pyXPCSviewer parses rich X-ray photon correlation spectroscopy (XPCS) results into independent PyQt widgets that are both interactive and easy to maintain. pyXPCSviewer is open-source and is open to customization by the XPCS community for ingestion of diversified data structures and inclusion of novel XPCS techniques, both of which are growing demands particularly with the dawn of near-diffraction-limited synchrotron sources and their dedicated XPCS beamlines.
... To assess the structure and heterogeneity of SLs during annealing and the nature of the annealing process itself requires direct space visualization of a three-dimensional (3D) object in solution at the nanoscale and characterization of similarly small fluctuations in the structure as a function of time. Coherent hard x-ray scattering [14] techniques such as coherent diffractive imaging (CDI) [15][16][17][18][19][20][21][22][23], x-ray photon correlation spectroscopy (XPCS) [24][25][26][27][28][29][30][31][32], or a combination of the two [33] in principle allow such characterizations. In this particular case, however, they also require a high degree of translational and rotational immobilization of individual condensed structures in a focused x-ray beam in solution for observation over at least tens of minutes. ...
Article
Full-text available
Solution-phase bottom up self-assembly of nanocrystals into superstructures such as ordered superlattices is an attractive strategy to generate functional materials of increasing complexity, including very recent advances that incorporate strong interparticle electronic coupling. While the self-assembly kinetics in these systems have been elucidated and related to the product characteristics, the weak interparticle bonding interactions suggest the superstructures formed could continue to order within the solution long after the primary nucleation and growth have occurred, even though the mechanism of annealing remains to be elucidated. Here, we use a combination of Bragg coherent diffractive imaging and x-ray photon correlation spectroscopy to create real-space maps of supercrystalline order along with a real-time view of the strain fluctuations in aging strongly coupled nanocrystal superlattices while they remain suspended and immobilized in solution. By combining the results, we deduce that the self-assembled superstructures are polycrystalline, initially comprising multiple nucleation sites, and that shear avalanches at grain boundaries continue to increase crystallinity long after growth has substantially slowed. This multimodal approach should be generalizable to characterize a breadth of materials in their native chemical environments, thus extending the reach of high-resolution coherent x-ray characterization to the benefit of a much wider range of physical systems. Published by the American Physical Society 2024
... Although large-emittance first-and second-generation light sources produced low-coherence X-ray beams, experimental techniques exploiting coherence started to be developed at second-generation sources (Sutton et al., 1991;Dierker et al., 1995). The considerably higher brightness and coherence of the third-and fourth-generation light sources allowed for maturity and led to further proliferation of coherence exploiting techniques. ...
Article
Full-text available
Physical optics simulations for beamlines and experiments allow users to test experiment feasibility and optimize beamline settings ahead of beam time in order to optimize valuable beam time at synchrotron light sources like NSLS-II. Further, such simulations also help to develop and test experimental data processing methods and software in advance. The Synchrotron Radiation Workshop ( SRW ) software package supports such complex simulations. We demonstrate how recent developments in SRW significantly improve the efficiency of physical optics simulations, such as end-to-end simulations of time-dependent X-ray photon correlation spectroscopy experiments with partially coherent undulator radiation (UR). The molecular dynamics simulation code LAMMPS was chosen to model the sample: a solution of silica nanoparticles in water at room temperature. Real-space distributions of nanoparticles produced by LAMMPS were imported into SRW and used to simulate scattering patterns of partially coherent hard X-ray UR from such a sample at the detector. The partially coherent UR illuminating the sample can be represented by a set of orthogonal coherent modes obtained by simulation of emission and propagation of this radiation through the coherent hard X-ray (CHX) scattering beamline followed by a coherent-mode decomposition. GPU acceleration is added for several key functions of SRW used in propagation from sample to detector, further improving the speed of the calculations. The accuracy of this simulation is benchmarked by comparison with experimental data.
... The temporal resolution of the technique is defined by the number of images taken in time. In the past decades, considerably slow systems have been in the focus of XPCS studies [6][7][8][9][10][11][12][13][14]. Recently, new advanced detector technology has begun to offer 2D detectors with the microsecond temporal resolutions in XPCS measurements (e.g., VIPIC [15], UFXC32k [16], Tristan detector [17], and the XSPA-500k [18]). ...
Article
Full-text available
The ability of pulsed nature of synchrotron radiation opens up the possibility of studying microsecond dynamics in complex materials via speckle-based techniques. Here, we present the study of measuring the dynamics of a colloidal system by combining single and multiple X-ray pulses of a storage ring. In addition, we apply speckle correlation techniques at various pulse patterns to collect correlation functions from nanoseconds to milliseconds. The obtained sample dynamics from all correlation techniques at different pulse patterns are in very good agreement with the expected dynamics of Brownian motions of silica nanoparticles in water. Our study will pave the way for future pulsed X-ray investigations at various synchrotron X-ray sources using individual X-ray pulse patterns.
... X-ray photon correlation spectroscopy (XPCS) is an emerging technique to probe nanoscale and microscale dynamics [1][2][3][4]. It has been applied to various sample systems including amorphous materials [5][6][7][8], colloidal dispersions [9][10][11], and strongly correlated materials [12,13]. ...
Article
Full-text available
X-ray free-electron lasers (XFELs) with megahertz repetition rates enable X-ray photon correlation spectroscopy (XPCS) studies of fast dynamics on microsecond and sub-microsecond time scales. Beam-induced sample heating is one of the central concerns in these studies, as the interval time is often insufficient for heat dissipation. Despite the great efforts devoted to this issue, few have evaluated the thermal effects of X-ray beam profiles. This work compares the effective dynamics of three common beam profiles using numerical methods. Results show that under the same fluence, the effective temperatures increase with the nonuniformity of the beam, such that the Gaussian beam profile yields a higher effective temperature than the donut-like and uniform profiles. Moreover, decreasing the beam sizes is found to reduce beam-induced thermal effects, in particular the effects of beam profiles.
Article
Recent technological breakthroughs in synchrotron and x-ray free electron laser facilities have revolutionized nanoscale structural and dynamic analyses in condensed matter systems. This review provides a comprehensive overview of the advancements in coherent scattering and diffractive imaging techniques, which are now at the forefront of exploring materials science complexities. These techniques, notably Bragg coherent diffractive imaging and x-ray photon correlation spectroscopy, x-ray magnetic dichroism, and x-ray correlation analysis leverage beam coherence to achieve volumetric three-dimensional imaging at unprecedented sub-nanometer resolutions and explore dynamic phenomena within sub-millisecond timeframes. Such capabilities are critical in understanding and developing advanced materials and technologies. Simultaneously, the emergence of chiral crystals—characterized by their unique absence of standard inversion, mirror, or other roto-inversion symmetries—presents both challenges and opportunities. These materials exhibit distinctive interactions with light, leading to phenomena such as molecular optical activity, chiral photonic waveguides, and valley-specific light emissions, which are pivotal in the burgeoning fields of photonic and spintronic devices. This review elucidates how novel x-ray probes can be leveraged to unravel these properties and their implications for future technological applications. A significant focus of this review is the exploration of new avenues in research, particularly the shift from conventional methods to more innovative approaches in studying these chiral materials. Inspired by structured optical beams, the potential of coherent scattering techniques utilizing twisted x-ray beams is examined. This promising direction not only offers higher spatial resolution but also opens the door to previously unattainable insights in materials science. By contextualizing these advancements within the broader scientific landscape and highlighting their practical applications, this review aims to chart a course for future research in this rapidly evolving field.
Article
Full-text available
Unraveling the mechanism of water’s glass transition and the interconnection between amorphous ices and liquid water plays an important role in our overall understanding of water. X-ray photon correlation spectroscopy (XPCS) experiments were conducted to study the dynamics and the complex interplay between the hypothesized glass transition in high-density amorphous ice (HDA) and the subsequent transition to low-density amorphous ice (LDA). Our XPCS experiments demonstrate that a heterodyne signal appears in the correlation function. Such a signal is known to originate from the interplay of a static component and a dynamic component. Quantitative analysis was performed on this heterodyne signal to extract the intrinsic dynamics of amorphous ice during the HDA–LDA transition. An angular dependence indicates non-isotropic, heterogeneous dynamics in the sample. Using the Stokes–Einstein relation to extract diffusion coefficients, the data are consistent with the scenario of static LDA islands floating within a diffusive matrix of high-density liquid water.
Article
Full-text available
In conventional frame-based image sensors, every pixel records brightness information and sends this information to a receiver serially in a scanning fashion. This full-frame readout approach suffers from high bandwidth requirements and increased power consumption with the increasing size of the pixel array. Event-based image sensors are gaining popularity for reducing the bandwidth and power requirements by sending only meaningful data in an event-driven approach with the help of address-event representation (AER) communication protocol. However, the event-based readout suffers from increased latency and timing error when the number of pixels with an event increase. In this paper, we introduce a new field-programmable AER (FP-AER) encoding scheme which offers benefits of both frame-based and event-based approaches. The readout design can be configured “in the field” using configuration bits. We also compare the performance of the proposed design against existing AER-based approaches for imaging applications and show that FP-AER performs best in both scanning and event-based readout.
Chapter
Transport of nanoparticles has received increasing attention for both fundamental interest to test prevailing models as well as practical interest in separations and self‐healing applications. Methods for measuring particle dynamics are briefly reviewed to orient the reader to different methodologies. The transport of nanoparticles through polymer melts, solutions, and cross‐linked networks is described with emphasis on systems where the nanoparticle is on the length scale of the defining structure including the tube size, correlation length, and mesh size, respectively. Because nanoparticles can interact with the different mediums, the interactions between nanoparticles and polymer and the subsequent impacts on dynamics are described. The transport of particles at liquid–solid (polymer brush) interfaces is addressed due to growing interest in applications ranging from understanding dynamics in porous media to the translocation of DNA. Both hard particles and polymer grafted particles (soft particles) are described. The objective of this article is to provide the reader with fundamental descriptions of phenomena while reviewing the current state of understanding. Further directions will leverage advances in inorganic chemistry to prepare monodisperse, functional nanoparticles as well as polymer chemistry to create novel media, such as networks with monodisperse mesh size.
Article
REFLECTED light from a coherent light source such as a laser shows a graininess known as speckle. In general, a speckle pattern is produced whenever randomly distributed regions of a material introduce different phase shifts into the scattering of coherent incident light. If the arrangement of the regions evolves with time, the speckle pattern will also change. Observation of the intensity fluctuations at a single point in the speckle pattern provides a direct measure of the time correlation function of the inhomogeneity. This leads to a technique1,2, alternatively called light-beating spectroscopy, dynamic light scattering or intensity fluctuation spectroscopy, which has been widely used with visible light to study processes such as critical fluctuations near phase transitions in fluids and the diffusion of particles in liquids. But it is not possible to study processes involving length scales less than about 200 nm, or those in opaque materials, using visible light. If intensity fluctuation spectroscopy could be carried out using coherent X-rays, however, one could probe the dynamics of processes involving atomic length scales in a wide range of materials. Here we show that by using a high-brilliance X-ray source it should be possible to perform this type of measurement using radiation of wavelength ~0.15nm. Specifically, we have observed a speckle pattern in the diffraction of coherent X-rays from randomly arranged antiphase domains in a single crystal of the binary alloy Cu3Au.
Article
A generalized hydrodynamic theory is developed for systems of interacting Brownian particles on the basis of a Fokker-Planck equation. General results are derived for correlation functions, frequency- and wave-vector dependent transport coefficients. Explicit expressions for moments, cumulants and the hydrodynamic limits of the transport coefficients are given. For the special cases of overdamped systems with and without hydrodynamic interaction the general results are simplified. As an example for the application of this approach the system of charged spherical polystyrene spheres in aqueous solution is treated in detail. The generalized transport functions are evaluated in mode-mode coupling approximation and detailed numerical results are presented for various collective and single-particle properties. Finally, the relationship to a corresponding Smoluchowski approach is discussed.
Article
MANY properties of colloids and suspensions depend on the particle size. Series of monodisperse suspensions of the same chemical composition but of rather different particle sizes may be used to study particle size dependent phenomena, such as Brownian motion, light scattering, sedimentation and electrophoresis of small particles. We have used such series to demonstrate the increased tendency of metal suspensions to coagulate in the presence of electrolytes as the radius of the particles increases1.
Article
We have successfully implemented a method to measure intensity autocorrelation functions with a CCD camera and a fast computer. We report light scattering experiments on solutions of diffusing latex spheres in glycerol and compare our results to those obtained with a conventional hardwired electronic correlator. Averaging allows significant reduction of measurement times and makes this technique suitable for the study of systems with very long time scales. Moreover, the CCD setup allows measurement of two‐time correlation functions under nonequilibrium conditions where conventional correlators which rely on time‐averaging fail. The new method may offer significant advantage for low‐intensity applications like x‐ray correlation spectroscopy.
Article
We have observed x-ray speckle at grazing incidence from gold-coated films of symmetric diblock copolymers of polystyrene (PS) and polymethylmethacrylate (PMMA). The polymer films consisted of micron-sized islands on a uniform surface. Coherent 6-keV photons were selected by monochromatizing and collimating the x-ray beam from a bending magnet synchrotron source. The visibility of the speckle patterns was enhanced with increasing degrees of spatial coherence of the photons. Speckles due to coherent scattering from islands with length scales up to 100 [mu]m were clearly identified.
Article
We have carried out intensity fluctuation spectroscopy measurements using coherent x rays to study the dynamics of critical fluctuations in a binary alloy at equilibrium. An intense coherent hard x-ray beam, produced from an undulator source, was scattered from a single crystal of Fe3Al held at temperatures near the B2-DO3 order-disorder transition. A speckle pattern was observed at the \(12 12 12\) superlattice reflection. Below Tc it was essentially static, while above Tc it fluctuated in time. The behavior of the normalized time correlation function is consistent with predictions of theory.
  • M L Williams
M. L. Williams et al., J. Phys. Chem. 77, 3701 (1955).
Laser Light Scattering: Basic Principles and Practices
  • Benjamin See
  • Chu
See, for example, Benjamin Chu, Laser Light Scattering: Basic Principles and Practices (Academic Press, San Diego, 1991), 2nd ed.
  • P Y Apollo
  • P Wong
  • Wiltzius
Apollo P. Y. Wong and P. Wiltzius, Rev. Sci. Instrum. 64, 2547 (1993).