High Frequency Acoustics in Colloid-Based Meso - and Nanostructures by Spontaneous Brillouin Light Scattering (Springer Theses) - Tim Still
High Frequency Acoustics in Colloid-Based Meso - and Nanostructures by Spontaneous Brillouin Light Scattering (Springer Theses) - Tim Still
High Frequency Acoustics in Colloid-Based Meso - and Nanostructures by Spontaneous Brillouin Light Scattering (Springer Theses) - Tim Still
The series ‘‘Springer Theses’’ brings together a selection of the very best
Ph.D. theses from around the world and across the physical sciences. Nominated
and endorsed by two recognized specialists, each published volume has been
selected for its scientific excellence and the high impact of its contents for the
pertinent field of research. For greater accessibility to non-specialists, the
published versions include an extended introduction, as well as a foreword by
the student’s supervisor explaining the special relevance of the work for the field.
As a whole, the series will provide a valuable resource both for newcomers to the
research fields described, and for other scientists seeking detailed background
information on special questions. Finally, it provides an accredited documentation
of the valuable contributions made by today’s younger generation of scientists.
123
Author Supervisor
Tim Still Prof. Hans-Jürgen Butt
Max Planck Institute for Polymer Research Max Planck Institute for Polymer Research
Ackermannweg 10 Ackermannweg 10
55128, Mainz, Germany 55128, Mainz, Germany
[email protected] [email protected]
DOI 10.1007/978-3-642-13483-8
This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcast-
ing, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this
publication or parts thereof is permitted only under the provisions of the German Copyright Law of
September 9, 1965, in its current version, and permission for use must always be obtained from
Springer. Violations are liable to prosecution under the German Copyright Law.
The use of general descriptive names, registered names, trademarks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
T. Still, M. D’Acunzi, D. Vollmer, G. Fytas, J. Coll. Interf. Sci. 2009, 340, 42.
Copyright 2009, Elsevier, Inc. All rights reserved.
Materials that can mold the flow of elastic waves of certain energy in certain
directions are called phononic materials. The present thesis deals essentially with
such phononic systems, which are structured in the mesoscale (\1 lm), and with
their individual components. Such systems show interesting phononic properties in
the hypersonic region, i.e., at frequencies in the GHz range. It is shown that
colloidal systems are excellent model systems for the realization of such phononic
materials. Therefore, different structures and particle architectures are investigated
by Brillouin light scattering, the inelastic scattering of light by phonons.
Both the mechanical properties of the individual colloidal particles, which
manifest in their resonance vibrations (eigenmodes), as well as the acoustic
propagation in colloidal structures have been investigated. The measurement of the
eigenmodes allows for new insights into physical properties at the mesoscale, e.g.,
confinement effects, copolymer behavior, or the non-destructive determination of
nanomechanical properties of core–shell particles, supporting the working groups
aim to achieve a deeper understanding of ‘soft mechanics’ at small length scales.
Another novel contribution assigned to this thesis is the first experimental reali-
zation of a phononic band gap arising from the interaction of these particle ei-
genmodes with the effective medium band (hybridization gap). This finding
already gave new impulses to the whole field of phononics.
The thesis was performed between 03/2007 and 08/2009 at Max Planck Institute
for Polymer Research, leading to several publications and presentations in inter-
national conferences, and was honored summa cum laude by the University
of Mainz.
ix
Acknowledgments
The work presented in this thesis would be nothing without the help of so many
people, who advised me, who provided samples, who eased my daily life, or
helped me in many other ways to make not only a happy but also a successful time
out of the last two and a half years.
First of all, my thanks go to Prof. Hans-Jürgen Butt, who gave me the oppor-
tunity to do my Ph.D. in his research group at Max Planck Institute for Polymer
Research. I enjoyed very much the freedom of science, which is not given
everywhere to a Ph.D. student.
Prof. Carsten Sönnichsen (J. Gutenberg-Univ. Mainz) and Prof. Alfons van
Blaaderen (University of Utrecht) kindly agreed to be the second and third referee
of my thesis.
The deepest gratitude I owe to Prof. George Fytas, who was a great supervisor,
giving me all the freedoms I needed to develop my own ideas, but on the other
hand pushing me in the best possible meaning of the word to go on with them in a
very consequent and productive way.
I would like to thank Prof. Werner Steffen for many scientific and non-scientific
discussions, technical support, proof-reading a part of this thesis, and for hiring
me.
When I started my Ph.D. in Mainz, there were two fellows that supported me a
lot by introducing me into BLS and some of the topics that accompanied me for
the next years, Dr. Cheng Wei and Dr. Eugenia Nuñez. Especially Cheng Wei was
always available to help me to solve my smaller and bigger problems with the
setup and to discuss my results. Beyond that, I enjoyed his special interpretation of
far eastern philosophy (‘You can even ride on a donkey!’).
When reading this thesis, it becomes clear that nearly all presented results are
based on diverse cooperations, since the success of my studies was up to the
continuous supply with all kinds of high quality samples. The most fruitful and
longest lasting of these cooperations was with Dr. Markus Retsch and Dr. Uli
Jonas. I really enjoyed our work together, since both of them came up with a great
deal of improvements and completely new ideas. Markus was the best conceivable
‘colloid cook’, who always delivered requested samples or realized just discussed
xi
xii Acknowledgments
ideas incredibly shortly and with impressive quality. Beyond science, he became a
good friend.
I also acknowledge the fruitful cooperation with Maria D’Acunzi, Gabriele
Schäfer and Dr. Doris Vollmer (all MPIP), leading to the investigation of PS–silica
core–shell particles.
I thank Dr. Peter Spahn and Diana Kiefer from Dr. Götz Hellmann’s group at
DKI Darmstadt for the ongoing cooperation on Silica–PMMA core–shell particles,
silica suspensions, and the phoXonic films.
Kenneth Kearns and Prof. Mark Ediger from University of Wisconsin must be
named thankfully for the cooperation on the stable organic glasses project, which
will be continued with Dr. Zahra Fakhraai.
For their theoretical support I would like to thank Dr. Rebecca Sainidou (Univ.
du Havre), Prof. Nikolaos Stefanou, and Georgios Gantzounis (both University of
Athens). Especially Rebecca was involved in many projects, and I want to thank
her not only for her calculations, but also for her patience in explaining me her
results.
I thank my office mates Akihiro Sato and Nikos Gomopoulos for many helpful
discussions, as well as Nikos for building the stretching machine.
I benefited from the great infrastructure at the Max Planck Institute for Polymer
Research. Therefore, I thank the people in the different service groups, the
mechanical shop, the electronics shop, and besides that the technical support by
Andreas Best, Gunnar Glasser and Maren Müller (SEM), Melanie Dröge and Petra
Räder (DSC), Dr. Konstantinos Mpoukouvalas (dielectric), and Uwe Rietzler
(AFM).
For the nice working atmosphere and the activities beyond the daily work I want
to thank all involved members of AK Butt, but also, in the same manner, the
people from the material research group (formerly known as AK Knoll).
This thesis was proof-read by Johannes Schmidt (MPI Golm), Daniel Szubrin
(University of Marburg), Martin Münzel (LMU Munich), and Justine Witosch
(MPI Martinsried). Thank you all.
Coming to the end of these acknowledgments, I want to thank the supervisor
of my diploma work, Prof. Wolf-Christian Pilgrim from Philipps-Universität
Marburg for introducing me into the field of inelastic scattering. Furthermore,
I thank many colleagues from all over the world that I met on several conferences
and who gave new impulses to me during many interesting discussions.
I want to thank my parents for all the support they gave me and all chances they
facilitated in my life.
And finally, I want to express my deepest thankfulness to the woman I love, who
gave me so much joy and happiness in the last years. Danke!
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 General Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Aims and Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1 Vertical Lifting Deposition and Colloidal Crystals. . . . . . . . . . . . 35
3.2 Melt Compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3 Polymer and Colloid Characterization Techniques . . . . . . . . . . . . 38
3.3.1 Photon Correlation Spectroscopy . . . . . . . . . . . . . . . . . . 38
3.3.2 Electron Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3.3 Differential Scanning Calorimetry . . . . . . . . . . . . . . . . . . 40
3.3.4 Density Gradient Column. . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.5 Wide Angle X-ray Scattering . . . . . . . . . . . . . . . . . . . . . 41
3.3.6 UV/VIS Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3.7 Gel Permeation Chromatography . . . . . . . . . . . . . . . . . . 42
3.4 Theoretical Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4.1 Single-sphere Scattering Cross-section Calculations . . . . . 44
xiii
xiv Contents
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Abbreviations
2D Two-dimensional
3D Three-dimensional
a Angle incident laser/sample
ag Mark–Houwink parameter
b(T,S) (Isothermal, adiabatic) compressibility
Dielectric constant
Extinction coefficient
/ Glancing angle
u, w, v Scalar functions
C Central point of fcc BZ
g Viscosity
k Wave length
k Lamé coefficient
K Phonon periodicity
l Lamé coefficient
m Velocity of light in vacuum
h Scattering angle
. Mass density
r Poisson’s ratio
r Scattering cross section
r
^ Normalized scattering cross section
rik Stress tensor
s Correlation time
f Tandem FP angle
x Angular frequency
A Free energy
A Area under the curve
A Absorbance
Alml’m’ Structure constants
BLS Brillouin light scattering
nBA n-Butyl acrylate
xvii
xviii Abbreviations
BG Bragg gap
cF Coefficient of finesse
cl/t/eff (Longitudinal/transverse/effective) sound velocity
Cp/T Specific heat (at constant p/T)
Cik Components of the stiffness matrix
CCD Charge-coupled device
d Diameter
d1/2 FP mirror distances
D Diffusion coefficient
DOS Density of states
DSC Differential scanning calorimetry
DTA Differential thermo analysis
DVB Divenyl benzene
E Young’s modulus
E Electric field
E0 Field amplitude
EMT Effective medium theory
f Frequency
F Finesse
F Force
flm(r, h, /) Solution of the scalar Helmholtz equation
fcc Face centered cubic
FP Fabry–Pérot interferometer
FSR Free spectral range
G Green’s function
G Reciprocal space vector
G(q, s) Time–correlation function
g2(q, s) Second order autocorrelation
GPC Gel permeation chromatography
h, h Planck quantum (/2p)
hcp Hexagonal close packing
HG Hybridization gap
I (Scattering) intensity
I Unit tensor
IMC Indomethacin
kB Boltzmann’s constant
KPS Potassium persulfate
K Gordon–Taylor parameter
K Bulk modulus
Kg Mark–Houwink parameter
k(i/sc) (Incident/scattered) Wave vector
l Longitudinal
L Distinct point in reciprocal fcc lattice
L Shear modulus
l, m, n Three independent vectors
Abbreviations xix
T Transmittance
TEM Transmission electron microscope
TNB aab-1,3,5-tris-Naphthylbenzene
u Displacement vector
uik Strain tensor
UV/VIS Ultraviolet/visible light (spectroscopy)
V (Scattering) volume
v Velocity
w Work
w Relative width of a band gap
wc Degree of crystallization
X Distinct point in reciprocal fcc lattice
x(n, l) Eigenmode constant
Ylm Spherical harmonics
Chapter 1
Introduction
One of the most excitatory and technically promising new field in physics in the
last decades was for sure the emerging possibility to control and manipulate the
flow of light by engineering of macroscopic media with periodic dielectric func-
tion, i.e., with periodic variation of the refractive index in one, two, or three
dimensions (Fig. 1.1). Since the first realization of such photonic crystals, an
abundance of publications on this topic have been released, including several new
high impact journals. Beside some quantum mechanical effects, the striking
capability of a photonic crystal is its aptitude to prevent light from propagating in
certain directions with specified frequencies, while they can operate as omni-
directional reflectors of light, independent from angle and polarization [1]. If the
propagation of electromagnetic waves is forbidden at certain frequencies, one
speaks about a photonic band gap.
The appearance of band gaps for distinct wavelengths in materials with periodic
changed dielectric constant can be rationalized as an interference effect. The
light is partially reflected at each layer interface. The multiple reflection interfere,
and if the periodicity spacing is commensurate to the wavelength of the incident
light, the destructive interference eliminates the forward propagation of the elec-
tromagnetic wave. In the simplest case of a one-dimensional (1D) photonic crystal,
i.e., with periodic modulation of only in z-direction, with normal incident light,
the gap occurs when the wavelength of the light kl is twice the crystals spatial
period a, kl = 2a. In the concept of the reciprocal space that is commonly used in
solid state physics, this belongs to a wave vector k = 2p/kl (in this case k is
identical to its component in z-direction) at the edge of the first Brillouin zone, i.e.,
k = p/a. The problem was first treated by Lord Rayleigh (1887), and such gap is
usually called Bragg gap. It can be shown that the width of such a photonic gap
depends on the dielectric contrast between the two components.
The concepts presented for the molding of electromagnetic waves can, in
principle, be transferred to mechanical waves, i.e., sound waves. A material with
Fig. 1.1 Photonic or phononic crystals with periodicity in one, two, or three directions. In
photonic crystals, white and black segments have different refractive index, while in phononic
crystals the components must have different elastic moduli or densities
periodic modulation of the elastic properties, i.e., of the elastic moduli or the
density, is called a phononic crystal after the quantized mode of vibration in a rigid
crystal lattice—the phonon (in analogy to the photon in the photonic crystals)
[2, 3]. However, the theoretical treatment of the propagation of mechanical waves
is more complicated. While the light propagation can be fully described by a wave
with distinct polarization and velocity, which depends on the refractive index
of the passed medium, a mechanical wave can, in principle, propagate in two
different ways.
Both mechanisms are schematically shown in Fig. 1.2. The longitudinal sound
wave on the left side shows displacement along the propagation direction, while in
the transverse sound wave the displacement is perpendicular to the propagation
direction symbolized by the phonon wave vector k. Both waves propagate with
different sound velocities cl (longitudinal) or ct (transverse), respectively. The
relevant propagation mechanisms depend on the type of the material. In solid
materials, both kind of waves can propagate, while in liquid or gaseous media only
longitudinal waves are supported. Hence, the different mechanical waves are often
distinguished between elastic waves in solids and acoustic waves in fluids.
However, in this thesis the terms elastic wave, acoustic wave, or sound wave will
be used synonymously for all kinds of mechanical waves.
(a) (b)
Fig. 1.2 Propagation mechanisms of mechanical waves: a longitudinal wave with displacement
along the propagation direction (given by the direction of the wave vector k); b transverse sound
wave with displacement perpendicular to the propagation direction
1.1 General Introduction 3
On an atomic scale, the concept of the phonon is used to describe for example
the heat capacity of solids. The frequencies of the allowed mechanical waves are
quantized as a function of the distance ‘ between neighboring atoms. The atoms
(the black points in Fig. 1.2) are displaced, and the displacement of each atom is
described by the displacement vector u(r, t) as a function of position and time.
However, in this thesis the interest lays on much lower frequencies with phonon
wavelength K ‘, thus the allowed mechanical frequencies can be regarded as
continuous. Figure 1.2 holds also in this case, but now the points represent volume
elements containing many atoms. It should be noted that the spatial displacement
of volume elements as a function of time leads to areas of instantaneously
increased or decreased pressure (or local density), thus mechanical waves can be
regarded as pressure waves. This picture also rationalizes why there is no prop-
agation of transverse sound waves in fluids since fluids do not support shear (aside
from very viscous liquids).
Phononic crystals can be designed by a one-, two-, or three-dimensional peri-
odicity of their elastic properties in analogy to their electromagnetic counterparts
(Fig. 1.1). The geometry of the periodic structure has strong influence on its
mechanical properties. One of the first examples of a two-dimensional phononic
crystal is a statue in Madrid, created by the artist E. Sempere [4]. The piece of art
consist of a large number of hollow stainless-steel cylinders with diameter of
2.9 cm in a simple cubic arrangement in a a = 10 cm unit cell, the whole statue
has a diameter of 4 m. A simple transmission experiment with a sound generator
and detector in the range of audible frequencies (1–5 kHz) has shown the existence
of phononic band gaps with strong sound attenuation depending on the orientation
of source and receiver relative to the statue, i.e., as a function of the crystallo-
graphic direction.
Like in the case of photonic crystals, the Bragg gap appears at the edge of the
first Brillouin zone, i.e., the phonon wavelength K is twice the lattice parameter a,
K = 2a, and the corresponding wave vector is again k = p/a. The frequency is
given by 2pf = cl/K. Since typical sound velocities are between 102 ms-1 (air)
and 105 ms-1 (condensed matter), attenuation in the audible range requires peri-
odicities in the centimeter to meter range, like in the example of the statue.
Structures that would attenuate seismic waves would have to be constructed even
in kilometer scales, the Bragg frequency scales with 1/a. According to this,
smaller, technically handier structures with spacing in the millimeter, micrometer,
or even nanometer-scale correspond to frequencies in the MHz to THz scale.
Most of the realized phononic band gap systems are restricted to sonic and
ultrasonic crystals with macroscopic periodicity, e.g., some millimeter-sized
crystals assembled manually, which could be probed by simple acoustic trans-
mission experiments in the commensurate frequency ranges [5–7]. Only recently,
Cheng et al. [8] could show the realization of a hypersonic phononic crystal based
on the self-assembly of sub-micron colloidal polymer crystals [9].
In fact, such mesoscopic structures have some advantages. They can be easily
prepared by self-assembly methods or by interference lithography [3] on a (relative
to the lattice spacing) large scale, they can show photo-thermal or photo-acoustic
4 1 Introduction
effects, and their size is also commensurate to the wavelength of visible light.
Therefore, it is possible to create a mesoscopic artificial crystal with periodic
variation of refractive index and mechanical properties, which would act as a
photonic and phononic crystal simultaneously. Sometimes, such ‘blind and deaf’
systems are called phoXonic crystals [10, 11]. Due to their size, colloidal crystals
from mesoscopic spheres show nice colorful reflections of light depending on the
scattering angle relative to their crystallographic direction as it can be found in
natural opals. Therefore, such systems are often (also in the present thesis) called
artificial opals. Figure 1.3 shows three photographs of a polished natural opal, an
artificial opal, which consists of dense packed silica spheres (d = 170 nm) in a
organic liquid matrix, and a layer of polystyrene spheres (d = 550 nm) on a glass
substrate.
These artificial opals are examples for colloid particle systems. In general, a
colloid is a particle or droplet of typical size between 1 nm and 10 lm that is
dispersed in another gaseous, liquid, or solid medium, e.g., milk (fat droplets in
water), fume (solid particles in gas), blood (solid particles in liquid), etc. In this
thesis the word colloid is used in a less general way; here it means in particular the
well defined colloidal solid particles, which are fabricated by the methods of
material chemistry. Systems of these particles in a matrix material are referred to
as colloidal systems.
Such colloidal systems have many advantages. Due to the onward development
of colloidal and material science, nowadays it became possible to control a wide
range of properties during the synthesis of colloids. Monodisperse spherical par-
ticles can be produced in a size range from few tens of nanometer to several
microns, utilizing a wide range of materials like metals, oxides (e.g., silica), or
Fig. 1.3 Natural and artificial opals: a Natural opal (Australia, Museum Idar-Oberstein,
Germany); b dense packed silica spheres (d = 219 nm, DKI Darmstadt) in organic liquid; c PS
spheres (d = 550 nm, M. Retsch, MPIP) on glass substrate
1.1 General Introduction 5
Colloids are very promising systems in many terms. They are auspicious candidates
to design hypersonic phononic systems that may be the basis of elaborated devices
that deal with the concurrent interaction of phonons and light (functional phoXonic
materials). However, so far there is no large knowledge about the details of the
nano- and mesomechanical behavior of individual colloids and the manipulation of
sound propagation in colloidal systems beyond the simple Bragg gap [8].
In this thesis a systematic approach is presented that captures the influence of
different parameters on the mechanical waves localized in spherical (as the most
accessible and theoretically best to capture model system), partially nanostruc-
tured, mesoscopic colloids as well as the propagation of sound waves in systems
composed of such particles [12, 32]. The first realization of a so-called hybrid-
ization gap in polymer based colloids will be presented, which originates from the
interaction between individual and collective colloid mechanical properties,
leading to crystalline as well as amorphous omnidirectional hypersonic band gap
systems [33].
These developments, however, must not be regarded as the end of the basic
experimental investigation of such systems. It will be shown that for silica based
systems with increasing mechanical contrast new effects occur, which will demand
further strong efforts in the theoretical and experimental conquest of this young
and promising field. Especially in consideration of the numerous appearing topics
related to photo-acoustics [34], thermo-acoustics [35], or nano- and microme-
chanics [36, 37], it seems worth to investigate colloidal model systems en detail
and the intimate relation of the heat conductivity in dielectric materials on phonons
1.2 Aims and Motivation 7
can have impact on a directional heat flow in the future. This work may be a
humble contribution to this exciting development.
1.3 Outline
References
1. Joannopoulos JD, Johnson SG, Winn JN, Meade RD (2008) Photonic crystals: molding the
flow of light, 2nd edn. Princeton University Press, Princeton
2. Economou EN, Zdetsis A (1989) Phys Rev B 40:1334
3. Maldovan M, Thomas EL (2008) Periodic materials and interference lithography. Wiley-
VCH, New York
4. Martinez-Salazar R et al (1995) Nature 378:241
8 1 Introduction
5. Montero de Espinosa FR, Jiménez E, Torres M (1998) Phys Rev Lett 80:1208
6. Liu ZY et al (2000) Science 289:1734
7. Vasseur JO et al (2001) Phys Rev Lett 86:3012
8. Cheng W et al (2006) Nat Mater 5:830
9. Thomas EL, Gorishnyy T, Maldovan M (2006) Nat Mater 5:773
10. Gorishnyy T, Maldovan M, Ullal C, Thomas E (2005) Phys World 18:24
11. Bernal MP, Roussey M, Baida F, Benchabane S, Khelif A, Laude V (2008) In: Ferraro P,
Grilli S, De Natale P (eds) Ferroelectric crystals for photonic applications, vol 307. Springer,
Heidelberg, pp 307–336
12. Still T et al (2008) Nano Lett 8:3194
13. Davis KE, Russel WB, Glantschnig WJ (1989) Science 245:507
14. Trau M, Saville DA, Aksay IA (1996) Science 272:706
15. Jiang P, Bertone JF, Hwang KS, Colvin VL (1999) Chem Mater 11:2132
16. Gu ZZ, Fujishima A, Sato O (2002) Chem Mater 14:760
17. Ruhl T, Spahn P, Hellmann GP (2003) Polymer 44:7625
18. Lyklema J (1991) Fundamentals of interface and colloid science. Academic Press, London
19. Lyklema J (2005) Fundamentals of interface and colloid science IV. Elsevier, London
20. Lyklema J (2005) Fundamentals of interface and colloid science V. Elsevier, Amsterdam
21. Magdassi S, Bassa A, Vinetsky Y, Kamyshny A (2003) Chem Mater 15:2208
22. Müller RH, Mäder K, Gohla S (2000) Eur J Pharm Biopharm 50:161
23. Pichot F, Pitts JR, Gregg BA (2000) Langmuir 16:5626
24. Meschede D (2006) Gerthsen Physik, 23rd edn. Springer, Heidelberg
25. Brillouin L (1922) Ann Phys (Paris) 17:88
26. Raman VV, Krishnan KS (1928) Nature 121:501
27. Mazurenko DA et al (2007) Phys Rev B 75:161102
28. Akimov AV et al (2008) Phys Rev Lett 101:033902
29. Mock R, Hillebrands B, Sandercock JR (1987) J Phys E Sci Instrum 20:656
30. Dyre JC, Olsen NB, Christensen T (1996) Phys Rev B 53:2171
31. Swallen SF et al (2007) Science 315:353
32. Still T et al (2008) J Phys Condens Matter 20:404203
33. Still T et al (2008) Phys Rev Lett 100:194301
34. Kippenberg TJ, Vahala KJ (2008) Science 321:1172
35. Damen EPN, Arts AFM, de Wijn HW (1995) Phys Rev Lett 74:4249
36. Cheng W et al (2008) Nano Lett 8:1423
37. Schliesser A et al (2008) New J Phys 10:095015
Chapter 2
Basics and Brillouin Light Scattering
In this chapter, first the basic expressions of elasticity will be introduced, espe-
cially with regard to the propagation of elastic waves in isotropic media and their
description as spherical waves. Then the general principles of light scattering will
be presented and applied to the special case of Brillouin light scattering, including
the discussion of its technical realization.
In this part, fundamental concepts of the theory of elasticity are briefly introduced,
following mostly the notation of the excellent textbook of Landau and Lifschitz [1].
Essential terms are the strain and stress tensors. To introduce them, we first
define the displacement vector u that shifted a point P that can be found by
following the vector r (with the three components x1 = x, x2 = y, and x3 = z in a
Cartesian system) from the origin of the coordinate system to the point P0 with
coordinates given by the vector r0 ; i.e.,
u ¼ r r0 : ð2:1Þ
The distance between any two infinitesimally adjacent points dl is given by
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffi
dl ¼ dx21 þ dx22 þ dx23 ¼ dx2i ; ð2:2Þ
using the Einstein summation convention behind the second equal. After a
deformation it becomes
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffi
dl0 ¼ dx02 02
1 þ dx2 þ dx3 ¼
02 dx02 i ð2:3Þ
The last approximation is valid if second order terms can be neglected. Obviously,
uik is a symmetric tensor. Each symmetric tensor can be diagonalized in any point.
With the diagonal elements u(1), u(2), u(3) the strain in any point can be written as
the sum of three independent terms, which give the strain in three orthogonal main
directions.
For u(i) 1 and if higher order terms are neglected, the relative change of
elongation becomes
dx0i dxi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ 1 þ 2uðiÞ 1 uðiÞ : ð2:8Þ
dxi
With that approximation we can write the relative volume change of an infinite-
simal small volume element dV ! dV 0 as the sum of the diagonal elements of the
strain tensor [1].
dV 0 dV
¼ uii : ð2:9Þ
dV
The resulting force F on any partial volume of an elastic Rbody can be written as
the integral over the forces on any element of the volume FdV: Because of the
identity of actio and reactio for all forces between any two points within the volume
the resulting force
R can also be written as the sum over the integrals of the three
components Fi, Fi dV, which can be translated into an integral over the surface:
Z Z I
orik
Fi dV ¼ dV ¼ rik dfk : ð2:10Þ
oxk
2.1 Elastic Waves in Condensed Matter 11
Here, Fi is expressed as divergence of a second rank tensor, the stress tensor rik:
orik
Fi ¼ ð2:11Þ
oxk
In Eq. 2.10, dfi are the components of a vector that is always oriented in the
direction normal to the surfaces. The nature of the stress tensor is visualized in
Fig. 2.1. When placing the surfaces of the volume element in the principal plane of
the coordinate system (xy, yz, or xz), the component rab of the stress tensor equals
to the a-component of the force that is normal to that xb-axis.
The work dw that is achieved to perform a deformation is
dw ¼ rik duik : ð2:12Þ
For reversible, elastic deformations it follows for the free energy A (A = U - TS,
dU = TdS - dw, S: entropy) that
oA
rik ¼ : ð2:13Þ
ouik T
It can be shown that the free energy after an elastic isothermal deformation of an
isotropic body can be expressed after series expansion and neglecting higher order
terms as
k
A ¼ A0 þ u2ii þ lu2ik ; ð2:14Þ
2
introducing the Lamé coefficients k and l. Note that k and l are force con-
stants. Since it was already shown that the volume change during a deformation
is expressed by the sum of the diagonal elements uii, it is obvious that, if the
term containing the uii becomes zero, the last term expresses a pure shear
deformation. Therefore, l is also called the shear modulus, sometimes denoted
by G.
y
yx
12 2 Basics and Brillouin Light Scattering
with uik = constdik it follows directly that every deformation can be written as a
sum of a pure shear deformation and a homogeneous dilatation. The term in
brackets is surely a pure shear deformation, because the sum of the diagonal
elements vanishes (dii = 3) and the other term is related to the homogeneous
dilatation.
For a perfectly elastic body, Hook’s law can be generalized to state that each
component of the stress tensor is linearly related to each component of the strain
tensor:
due to further symmetry considerations the stiffness matrix in Eq. 2.16 has the
following form [2]:
2 3 2 32 3
r1 C11 C12 C12 0 0 0 u1
6 r2 7 6 C12 C11 C12 0 0 0 7 6 u2 7
6 7 6 76 7
6 r3 7 6 C12 C12 C11 0 0 0 7 6 7
6 7¼6 7 6 u3 7 : ð2:19Þ
6 r4 7 6 0 0 0 C 0 0 7 6 7
6 7 6 44 7 6 u4 7
4 r5 5 4 0 0 0 0 C44 0 5 4 u5 5
r6 0 0 0 0 0 C44 u6
Because the three remaining constants are related by the following relation
M ¼ C11 : ð2:23Þ
The Young’s modulus E, also known as modulus of elasticity or tensile mod-
ulus, is a measure of the stiffness of an isotropic elastic material. It is given by the
ratio of longitudinal stress (which has units of pressure) and the dimensionless
longitudinal strain,
r11 lð3k þ 2lÞ
E¼ ¼ : ð2:24Þ
u11 kþl
The ratio of the lateral strain to the longitudinal strain defines the Poisson’s
ratio,
u22 k
r¼ ¼ : ð2:25Þ
u11 2ðk þ lÞ
14 2 Basics and Brillouin Light Scattering
When a body is deformed, the deformation usually goes along with changes in the
temperature. However, the heat transport is normally much slower compared to
periods of vibrations in the body. Therefore, one can regard the movements as
(quasi)adiabatic. It can be shown that in the adiabatic case the values for the Young’s
modulus and the Poisson’s ratio, given above for the isothermal case, change into
Ta2 Ta2
Eadiabatic ¼ E þ E2 and radiabatic ¼ r þ ð1 þ rÞE ; ð2:26Þ
9Cp 9Cp
where a ¼ ðoV=oTÞp and Cp is the specific heat at constant pressure [1]. In the
following, r and E mean their adiabatic values.
The general equation of movement can be written as
o2 ui orik
. ¼ ð2:27Þ
ot2 oxk
with mass density .: For the isotropic elastic medium the equation of movement
becomes
o2 ui E E
. 2
¼ r2 u þ : ð2:28Þ
ot 2ð1 þ rÞ 2ð1 þ rÞð1 2rÞgrad divu
Regarding a plane elastic wave in x-direction in an infinite medium, i.e.,
the deformation u depends only on the x-coordinate (all derivatives with respect
to y and z become zero), the components of the vector u become
o2 ux 1 o2 ux o2 uy 1 o2 uy o2 uz 1 o2 uz
¼ 0; ¼ 0; ¼ 0: ð2:29Þ
ox2 c2l ot2 ox2 c2t ot2 ox2 c2t ot2
The Eq. 2.29 are one-dimensional wave equations, cl and ct are their velocities
of propagation. Obviously, the propagation in x-direction is different from that in
the other directions. If the displacement ux lies in the direction of propagation of the
wave, the wave is called a longitudinal wave with longitudinal sound velocity cl:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffi
Eð1 rÞ k þ 2l C11
cl ¼ ¼ ¼ : ð2:30Þ
.ð1 þ rÞð1 2rÞ . .
In the other directions, the displacement (uy, uz) lies in a plane normal to the
direction of propagation of the wave. Such a wave is called a transverse wave with
transverse sound velocity ct:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffi sffiffiffiffiffiffiffi
E l C44
ct ¼ ¼ ¼ : ð2:31Þ
2.ð1 þ rÞ . .
2.1 Elastic Waves in Condensed Matter 15
Accordingly, from Eqs. 2.21 and 2.23 it is clear that we can express longitudinal
and shear modulus simply by
M ¼ .c2l ð2:32Þ
and
G ¼ .c2t ; ð2:33Þ
Because this thesis treats in many parts experiments on colloids acting as spherical
scatterers, it is meaningful to introduce briefly the principle solution of the elastic
wave equation for an isotropic medium. The equations presented here will be
required again in the discussion of the vibrational eigenmodes of spherical
particles.
In general, for a harmonic elastic wave of angular frequency x, the displace-
ment vector u can be written as
uðr; tÞ ¼ <½uðrÞ expðixt: ð2:36Þ
Using that one can write the general equation of motion in the following time-
independent form [3]:
where l represents the displacement associated with longitudinal wave and n and m
represent the transverse displacements, which are orthogonal to each other. These
three vectors can be related to scalar functions u; w; and v that are solutions of a
scalar Helmholtz equation (r2 + k2)f = 0:
1
l¼ ru; ð2:39Þ
kl
m ¼ r rw; ð2:40Þ
1
n¼ r r rv: ð2:41Þ
kt
The transverse displacement vectors m and n are herein expressed as the product
of the constant position vector r and the scalar functions mentioned above [4]. In
polar coordinates the solution of the scalar Helmholtz equation is known. It is
flm ðr; h; /Þ ¼ Rl ðkrÞYlm ðh; /Þ: ð2:42Þ
Rl(kr) are n-th order spherical Bessel functions, which represent the radial dis-
placement. Ylm(h, /) are the n-th order spherical harmonics (Legendre functions)
with l = 0, 1, 2, 3, ... and m being an integer -l B m B +l. Knowing this,
the vector solutions of Eq. 2.38 can be written as
1
llm ðR; kl Þ ¼ r½Rl ðkl rÞYlm ð^r Þ; l ¼ 0; 1; 2; 3; . . .; ð2:43Þ
kl
mlm ðR; kt Þ ¼ r ½rRl ðkt rÞYlm ð^r Þ; l ¼ 1; 2; 3; . . .; ð2:44Þ
1
nlm ðR; kt Þ ¼ r r ½rRl ðkt rÞYlm ð^r Þ; l ¼ 1; 2; 3; . . . ð2:45Þ
kt
This section is a brief introduction into the basic principles of light scattering.
While the derivation of scattering theory on the basis of quantum field theory is
possible, herein the scattering medium as well as the light are treated classically,
leading to practically the same results within the scope of this thesis. Of course
there are light scattering effects, as for example in the well-known Raman scat-
tering technique, which deals with rotational and vibrational transitions of single
atoms or molecules, i.e., effects in a quantum length scale, that must be treated
(at least partially) quantum mechanically. However, this thesis deals with the
investigation of phonons by Brillouin light scattering (Sect. 2.3) in condensed
2.2 Light Scattering Basics 17
matter. Thus the investigated phonons are classical waves with wavelengths in the
order of some nanometer up to micrometer, it is fully justified to apply a classic
theory.
The classical theory of light scattering in dense media developed by Einstein [5]
and Smoluchowski considers the sample as divided into small volume elements
large enough to contain many molecules (i.e., following classical physics), but
of linear dimension small compared to the wavelength of light. An incident
light wave induces a dipole moment in each volume element, which becomes
the source of scattered radiation. Provided that the induced polarization is
constant through the medium, the net scattered radiation in all directions but the
forward will be zero due to destructive interference, because the wavelets
scattered from each subregion differ only by a phase factor that depends on the
relative position of the small volumes. Therefore, neglecting small surface
effects, it is possible to pair each small volume with another small volume
whose scattered field is identical in amplitude but opposite in phase, thus they
cancel out [4, 6, 7].
However, in real media there will always be small random fluctuations in the
local dielectric constant due to the thermal motion of the atoms and molecules
in the sample. Because these fluctuations should be uncorrelated from one
volume element to the next, these regions are optically different, and therefore
also the amplitudes of the scattered light are uncorrelated. That means that now
light is also scattered in other directions than forward due to only partial
interference.
The local dielectric constant at a point at position r and time t; ðr; tÞ; is gen-
erally described by the dielectric constant fluctuation tensor dðr; tÞ: It describes
the relation between the local and the average dielectric constant 0 ;
ðr; tÞ ¼ 0 I þ dðr; tÞ: ð2:46Þ
Indeed, the same should be valid for the electric displacement field D and the
magnetizing field H: The solution of the Maxwell equations is lengthy and
18 2 Basics and Brillouin Light Scattering
performed elsewhere [7]. In the end it follows for the component of the scattered
electric field Es(R, t) that
Z
E0
Es ðR; tÞ ¼ exp iks R d3 r
4pR0 V ð2:49Þ
exp iðqr xi t½ns ½ks ðks ðdðr; tÞni Þ;
When—like in most light scattering experiments—ki & ks, i.e., the scattering is
quasi-elastic, the cosine rule becomes applicable and Eq. 2.51 becomes
h
q2 ¼ ki2 þ ks2 2ki ks ¼ 2ki2 ð1 cos hÞ ¼ 4ki2 sin2 : ð2:52Þ
2
Or, with ks = ki = 2pn/k, where the length of the incident wave vector is written
pffiffiffiffi
as function of the refractive index nðn ¼ 0 Þ and the incident wave length k:
h 4pn h
q ¼ 2ki sin ¼ sin : ð2:53Þ
2 k 2
This is the Bragg condition.
(a) (b)
Fig. 2.2 a At the detector, the total radiated field is the sum of all fields radiated from any
infinitesimal volume d3r at position r from the center O of the illuminated volume. The detector is
at position R: b General light scattering setup: incident light of polarization ni ; frequency xi, and
wave vector ki is scattered. Although it is scattered in all directions (not shown for clarity), the
figure shows only the light with ns and wave vector ks that can reach the detector after passing an
analyzer. The wave vector q is shown in gray as ki ks (according to [7])
2.2 Light Scattering Basics 19
ks2 E0
Es ðR; tÞ ¼ exp iðks R xi tÞdis ðq; tÞ; ð2:56Þ
4pR0
where
dis ðq; tÞ ns dðq; tÞni ð2:57Þ
is the component of the dielectric constant fluctuation tensor along the initial and
final polarization direction.
The time-correlated function of Es can then be written as
k4 jE0 j2
Es ðR; 0ÞEs ðR; tÞ ¼ s 2 2 2 hdis ðq; tÞi expðixi tÞ: ð2:58Þ
16p R 0
For a given experiment, the coefficient in Eq. 2.60 becomes a constant, and the
scattering intensity is then only affected by the spectral density of the dielectric
constant fluctuations, i.e., the integral
Z þ1
Iis ðq; xÞ / dt d is ðq; 0Þdis ðq; tÞ : ð2:62Þ
1
The integral of Eq. 2.62 over frequency, the integrated intensity at all fre-
quencies, provides information about the q-dependent mean-square fluctuations :
D E
Iis ðqÞ ¼ jdis ðqÞj2 : ð2:63Þ
Harking back on the definition of the scattering wave vector q (Eq. 2.50) and
the frequency shift x (Eq. 2.61), the scattering event can be considered in terms of
energy and momentum conservation. Most generally, during a scattering process,
the scattered photon sustains an energy change from hxi to hxs and a momentum
change from hki to
hks : This must be related to the creation or annihilation of an
excitation in the scattering medium. It is:
hx ¼
hxs hxi ð2:64Þ
hk ¼
hks
hk i : ð2:65Þ
Note that in these equations the energy and momentum of the excitation can have
both positive and negative sign.
light scattering (opaque) samples and the dispersion relations from the q-depen-
dent BLS spectroscopy on transparent samples is a powerful methodology to
investigate the elastic behavior of nanostructured materials.
This section will introduce the basic principles of BLS following a simple
approach. Further important results obtained by thermodynamical considerations
will be introduced very briefly. Then a description of the BLS setup with
selective attention on the main principles of the tandem Fabry–Pérot interfero-
meter is given.
The main principle of BLS is the scattering of photons on sound waves and the
constructive interference of the multiply reflected light beam as sketched in
Fig. 2.3. A plane elastic wave creates a periodic change of density and hence of
the local dielectric constant in a medium, symbolized by the black/white layers.
The typical velocity of an acoustic wave is between 103 and 104 ms-1, while the
velocity of the probing light c is &3 9 108 ms-1. Because of this great dis-
crepancy the dielectric inhomogeneities can be regarded as a quasi-static (i.e.,
‘frozen’) lattice on which the photons of the probing light are scattered. Thus, it is
justified to treat the medium as a periodic multilayer stack with periodicity K, the
wavelength of the phonon, as shown in Fig. 2.3. The probing laser light is multiply
reflected (under the scattering angle h) on these layers and the reflected light
interferes on the detector, whose distant to the sample R is much larger than the
periodicity of the scattering planes ðR
KÞ: The reflected intensity reaches its
maximum when the interference is constructive, i.e.,
h
2nK sin ¼ k ð2:66Þ
2
with refractive index n and wavelength of light k.
hks
hki ¼
hq: ð2:73Þ
Note that in these equations momentum and energy (frequency) of the acoustic
wave are positively defined and plus- and minus-signs correspond again to anti-
Stokes and Stokes process, respectively.
2.3 Brillouin Light Scattering 23
So far, this easy approach explains satisfactorily the appearance of the doublet
as well as the position of the frequency shifts. However, when performing BLS
experiments a central line always appears that is not explained by the above
considerations. Also the intensity of the signals can not be predicted. To elucidate
these features of a BLS spectra it is necessary to take some ideas from thermo-
dynamics into account.
The total scattering intensity of a Brillouin spectrum was first derived by
Einstein in 1910 [5, 6]. He started with considering as a function of density and
temperature ¼ ð.; TÞ: For . and T are statistically independent, one can write
the total differential
o o
d ¼ d. þ dT: ð2:74Þ
o. T oT .
Then
D E o 2 D E o 2 D E
2 2
ðdÞ ¼ ðd.Þ þ ðdTÞ2 : ð2:75Þ
o. T oT .
and
D E
ðd.Þ2 kB TbT
¼ ð2:77Þ
.2 v
for the mean-square fluctuation in density in the volume element v, where kB is
Boltzmann’s constant and bT is the isothermal compressibility, Eq. 2.75 becomes
D E o 2 D E o 2 k Tb
2 2 B T
ðdÞ ðd.Þ ¼ . : ð2:78Þ
o. T o. T v
Comparison with Eq. 2.63 shows the main result of these calculations,
I / bT ; ð2:79Þ
i.e., the total scattering intensity depends on the isothermal compressibility.
Without going into details, the full vectorial expression for IðqÞ is given by [7]
2
2 o
Iis ðqÞ ¼ ðni ns Þ V.2 kB TbT ð2:80Þ
o. T
with scattering volume V. If V(h) & const, which may be valid depending on the
experimental details, the total scattering intensity is also independent from the
scattering angle.
24 2 Basics and Brillouin Light Scattering
Although Eq. 2.80 gives the total scattering intensity, experiments showed that
only a part of the intensity belongs to the Brillouin doublets while the rest makes
for a central line. These findings could be explained by Landau and Placzek
regarding (S, p) as function of entropy S and pressure p [1]. In analogy to
Eq. 2.75 one can write
D E 2 D E o 2 D E
2 o 2
ðdÞ ¼ ðdSÞ þ ðdpÞ2 ð2:81Þ
oS p op S
It becomes clear that the intensity of the Brillouin doublet 2IB depends on the
adiabatic compressibility bS.
After further simplifications it can be shown that the ratio between total scat-
tering intensity (IC + 2IB) and the intensity of the Brillouin doublet is simply given
by the ratio of isothermal and adiabatic compressibility or of the specific heats at
constant pressure or volume [1, 6, 7]:
IC þ 2IB bT Cp
¼ ¼ : ð2:85Þ
2IB bS CV
In the form
IC b bS Cp CV
¼ T ¼ ¼c1 ð2:86Þ
2IB bS CV
it is known as the Landau–Placzek equation.
2.3 Brillouin Light Scattering 25
It was already pointed out that the relative shift of the photon frequency in BLS
spectroscopy (*108 to 1011 Hz) is quite subtle compared to the initial frequency
of the probing light (1014 Hz) or to the resolution obtained by Raman spectros-
copy, where frequency shifts in the order of 1013 Hz are measured frequently by
diffraction grating spectrometers. In order to achieve such high resolution, Fabry–
Pérot interferometers (FPs) are used in BLS. In combination with an highly
monochromatic laser light source a single FP or, even better, a design composed of
two FPs with multiple light pass can give excellent results. Here the FP principle is
elucidated and the construction details of the used tandem FP as well as the general
features of the whole BLS setup are illuminated.
In principle, a single FP is not much more than an etalon as it is used in laser
resonant cavities. The FP consists of two plane mirrors with reflectivity R mounted
accurately parallel to one another. The spacing between the mirrors is given by d and
can be varied by moving one of the mirrors. When the laser light enters the etalon
through the first mirror, which is usually a plane glass plate with a thin layer of metal
on the inner side, it is reflected numerously between the two mirrors. However, the
light with intensity II is already reflected when it enters the etalon, so that we have
only the intensity I0 = II(1 - R) inside the FP before the first internal reflection.
During the reflections within the FP, the intensity decreases after x reflections to
Ix ¼ I0 Rx ¼ II ð1 RÞRx ; ð2:87Þ
because every time a part of the light is also transmitted; a typical value for BLS
experiments is R & 0.93. Note that Eq. 2.87 is not fully true as it does not take
into account the dissipation of energy, which is usually considered by the
absorptance A, i.e., R + T + A = 1, with transmittance T. Anyhow, it implies that
with increasing reflectivity the number of reflections inside the FP increases
rapidly, and this is important for the working principle of the system. When the
light is reflected many times between the mirrors, the reflected beams interfere.
Only light that fulfills the equation
mk ¼ 2nFP d cos hFP ; ð2:88Þ
with m being an integer and nFP being the refractive index inside the FP, and hFP
being the angle between the light in the FP and the normal to the mirrors, is
transmitted losslessly due to constructive interference. Thus usually the interfer-
ometry is performed by moving one of the mirrors, there is just air between the
mirrors. (The other possibility would be to vary n during a measurement.) With
normal incidence (cos 0 = 1) and n = 1 we can therefore simplify to
2d
k¼ : ð2:89Þ
m
That means that only light with wavelength 2d/m is transmitted, and the trans-
mitted k can be easily scanned when scanning different ds by moving a mirror,
26 2 Basics and Brillouin Light Scattering
Dk Dm
¼ ð2:90Þ
k m
with initial wavelength k. For m ± 1 it follows in the described setup (normal
incidence, nFP = 1)
k2
FSRk ¼ ; ð2:91Þ
2d
or in the frequency domain
m
FSR ¼ ; ð2:92Þ
2d
with m being the velocity of light in vacuum.
The linewidth of the transmitted line depends strongly on R. As denoted above,
the more often the beam is reflected before it transmits the second beam of the FP
(or the first, but this light is uninteresting for the further analysis) the more often it
can interfere with itself. Therefore the destructive interference will be enforced for
wavelengths not satisfying Eq. 2.89 with increasing reflectivity and the function
I(k) becomes more narrow.
Indeed, the phase difference dp between each succeeding reflection is
2p
dp ¼ 2d: ð2:93Þ
k
The transmission function of the etalon T(dp) is found to be (http://en.wikipedia.
org/wiki/Fabry_Perot, April 2009)
ð 1 RÞ 2 1
Tðdp Þ ¼ ¼ ; ð2:94Þ
1 þ R2 2R cos dp 1 þ cF sin2 dp
2
4R
cF : ð2:95Þ
ð1 RÞ2
F is defined as the ratio between FSRk and the full width at half maximum
(FWHMk) of a transmission peak in the T(k)-function, i.e., F gives the relative
separation between nearest transmission peaks (cf. Fig. 2.4) and is also the most
important parameter for the practical resolution of a spectrometer. As it is
approximately
2.3 Brillouin Light Scattering 27
4
FWHMk pffiffiffiffiffi; ð2:96Þ
cF
where d1 and d2 are the mirror distances in FP1 and FP2 and p and q are integers.
The effect of the tandem operation is sketched in Fig. 2.5b. The next order of light
28 2 Basics and Brillouin Light Scattering
(a) (b)
d 2= d 1cos
transmission
FP1
1
FP2
FP2
2
FP1 tandem FP
translation
stage
d1 wavelength
Fig. 2.5 a Basic mechanism of a Sandercock multipass tandem Fabry–Pérot interferometer. The
light passes FP1 and then FP2. Two mirrors (one of FP1 and one of FP2) are moved simulta-
neously on a piezoelectric translation stage. The angle f of the light path between the two FPs is
chosen in that way that the ratio between d1 and d2 is constant. b Effect of the multi-pass through
two FPs: the neighboring orders (viewed from the harmonized central signals) that can pass each
single FP are suppressed. Only the common multiples of Dk1 and Dk2 will be transmitted
gainlessly
angle f between FP1 and FP2 is fixed and the mirror distances have then to be
adjusted in a way that
d2 ¼ d1 cos f: ð2:100Þ
A movement of the translation stage to the right shifts the spacings between the
mirrors simultaneously by Dd1 and Dd1cos f, i.e., the ratio keeps constant.
In my experiments, a six-pass tandem Fabry–Pérot interferometer was used.
The path of the light inside the FP as well as the other details of the applied BLS
setup are sketched in Fig. 2.6. The sample is mounted in the center of a goniometer
(Huber) in a custom-made sample holder with or without oven. A solid state
pumped frequency-doubled Nd:YAG laser (coherence; 150 mW at 532 nm) is
fixed on the goniometer and can be rotated so that scattering angles between 0° and
*160° can be chosen either in transmission or reflection geometry. In difference
to pure BLS backscattering techniques, i.e., h = 180°, this technique has the
advantage that not only the components of q but the whole wave vector is changed
according to Eq. 2.53.
Before the light reaches the sample, it passes a Glan polarizer (extinction ratio
10-5) with vertical polarization (V), i.e., perpendicular to the scattering plane, to
ensure fully polarized incident light. Behind the sample the light scattered in the
direction of the detector is collected by an aperture and focused into the entrance
pinhole of the tandem Fabry–Pérot interferometer (JRS Scientific Instruments) by
some lenses. Before entering the FP, a Glan–Thompson analyzer (extinction ratio
10-8) is passed that selects either vertically (V), i.e., perpendicular to the scat-
tering plane, or horizontally, i.e., parallel to the scattering plane, polarized light
(or, of course, everything in between). After passing the FP the transmitted light is
detected by an avalanche photo diode (APD) and processed by an multi-channel
analyzer with 1,024 channels. The further processing is performed by a computer
software.
The stability of the alignment is greatly enhanced by the use of a reference
beam. Therefore, a small amount of the laser light is diverted via the reflection
interferometer. Electronic
D
ANALZYER
SE
(5%) of a parallel plate on a beam splitter and an optical fiber from the incident
laser beam and introduced as a reference beam that gives the central line via an
optical fiber. Therefore, a mechanical shutter is used, which switches periodically
the entrance to the FP between reference beam and scattered light and excludes the
central elastic line. In order to avoid mechanical disruptions the whole setup is
placed on an optical table with active vibration damping.
In some experiments dealing with temperature-dependent effects, e.g., glass
transition experiments (Sect. 4.3.1 in Chap. 4) or measurements on kinetically
stable organic glasses (Chap. 6), an oven is used in order to control the tem-
perature of the sample in the range between ca. 10 and 200 °C. The oven is a
metal cylinder with filament and coolant tubes in the wall [13]. A small slit at the
hight of the laser and a cylindrical quartz glass insert in order to avoid heat
transfer through the slit allow the light to reach the sample and the detector
afterwards. The temperature is controlled electronically (built in-house) by two
Pt-100 temperature sensors inside the wall and inside the heat chamber, close to
the sample. The temperature can be stabilized within better than ±0.2 K in the
given range.
Most experiments in this thesis deal with samples on a plane substrate, i.e.,
‘films’. When performing the BLS experiment on such samples, there are in
principle two different scattering geometries, the transmission and the reflection
geometry. In transmission geometry the light scattered on the other side of the film
than that of the incident laser beam is investigated, while in reflection geometry
the light scattered on the side of the incident beam is regarded.
The two geometries are sketched in Fig. 2.7 and a full geometrical derivation is
given in Appendix.
ks
n q q
ki
n0=1 qpara
transmission
2.3 Brillouin Light Scattering 31
In the transmission case, it is found that the magnitude of the scattering wave
vector becomes
4pn 1 1 1 1 1
q¼ sin sin sin ðh aÞ þ sin sin a ð2:101Þ
k 2 n n
with the angles a and h given in Fig. 2.7a. It is shown in the appendix that
the length of the component of the scattering wave vector parallel to the film,
qpara, is
2p
qpara ¼ ðsin a þ sin ðh aÞÞ: ð2:102Þ
k
Obviously, in this equation the refractive index n of the sample is eliminated.
A special transmission geometry in which q ¼ qpara exists for h = 2a. In this
case q is given simply by
4p h
q ¼ qpara ¼ sin ðfor h ¼ 2aÞ: ð2:103Þ
k 2
To use this special geometry is advantageous in several ways, indeed nearly all the
experiments in this thesis are performed using it. First of all, it facilitates the
calculation of q, as the refractive index of the sample has not to be taken into
account. Although n of most ‘standard materials’ (e.g., typical polymers like
polystyrene) is well-known, the exact determination of n for unknown materials or,
even more, for composite materials, as they are widely used in this thesis, may
become complicated and time consuming. The second great benefit of this
geometry is that in this case the direction of q is well defined parallel to the film.
When discussing colloidal crystals [14] it is important to know which crystallo-
graphic direction is probed by the BLS experiment.
In the reflection geometry the general expression for q as a function of n, h, and
a becomes
4pn 1 1 1
q¼ cos sin1 sin a þ sin1 sin ðh þ aÞ : ð2:104Þ
k 2 n n
Also for the reflection case there is a special scattering geometry. If a ¼ 1802 h the
scattering wave vector is identical to its component perpendicular to the film, i.e.,
q ¼ qperp in Fig. 2.7b.
In a cylindrical sample, e.g., a liquid in a NMR-tube, q has simply the
magnitude
4pn h
qcylinder ¼ sin : ð2:105Þ
k 2
32 2 Basics and Brillouin Light Scattering
In the foregone section it was pointed out that the result of a BLS experiment is
usually the plot of longitudinal or transverse [or mixed, especially in thin films,
e.g., the so-called Lamb waves (http://en.wikipedia.org/wiki/Lamb_wave, May
2009)] sound waves as a function of the scattering wave vector q; while the
frequencies f(q) are given as a Doppler shift of the probing laser light around its
elastic line. The phase sound velocities are given by the slope of the f(q)-diagram.
Thus, in the case of a nonlinear function f(q), i.e., if the acoustic mode is
dispersive, the result of the BLS experiment is the dispersion relation in the
investigated q-range.
However, the appearance of a dispersion relation is only meaningful if it
is possible to determine q. If strong multiple scattering occurs in a sample,
q becomes ill-defined. In such samples f(q) is no longer accessible. Anyhow, now
the inelastic scattering from localized modes, i.e., the vibrational resonance
modes, can lead to incoherent BLS in analogy to the Raman scattering.
Indeed, in BLS experiments such samples show a spectrum where scattering at
all possible qs between zero degree incident and backscattering case contributes to
the final result, independent from the experimental scattering angle h. Further-
more, also the polarization information is lost.
Nevertheless, the Brillouin spectrum of multiple scattering samples with well
defined shape still gives useful information, which are not accessible by other
techniques. In Chap. 4 of this thesis, several experiments are discussed dealing
with the detection of vibrational resonance modes of spherical colloidal samples.
BLS spectra of such samples deliver many useful information about the individual
colloids and, in case of hybrid materials, about the mechanical properties of the
components, too. This section will give a short introduction into the mathematical
description of the vibrations in spheres, their theoretical calculation, and their
detection by BLS.
Under stress-free boundary conditions, the vibrational modes are usually
called eigenmodes. The eigenmodes for free homogeneous elastic spheres have
been derived by Lamb in the nineteenth century [15]. The modes can be clas-
sified as torsional and spheroidal ones, both labeled by the indices n, l, and m,
which describe the radial (n) and the angular (l, m) dependence of the dis-
placement (cf. Sect. 2.1.3). The torsional modes are fully tangential, i.e., they
involve only shear motions and do not cause changes in the sphere volume—they
do not contribute to the BLS intensity. The spheroidal modes involve usually
both shear and stretching motions, and they can be fully specified by two indices
in analogy to the atomic orbitals as the n-th order radial solution (n = 1, 2, 3,...)
for angular momentum ‘quantum number’ l (l = 0, 1, 2,...) [4, 12, 16]. Only
spheroidal modes with l = 0 have purely radial displacement (breathing modes)
[17].
By solving the wave equation for elastic and isotropic media (Eq. 2.37), Lamb
found the frequencies of the eigenmodes (n, l) to be (in a very general expression):
2.3 Brillouin Light Scattering 33
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi!
Rðcl ; ct Þ
f ðn; lÞ ¼ xðn; lÞ d; ð2:106Þ
.
with x(n, l) being a constant for each individual mode, R(cl, ct) being the rigidity,
and d being the diameter of the sphere. The rigidity is a function of the longitu-
dinal and transverse sound velocities and has the dimension of pressure, thus it is
simply a modulus. Indeed, e.g., for the (1,2)-mode, R is identical with the shear
modulus l(=C44).
The experiments discussed in Chap. 4 are all related to the vibrational modes of
colloidal spheres in air. Because of the large density (and elastic) mismatch
between air and the colloids, the boundary-conditions are quasi that of a stress-
free, undamped vibration, thus these modes are practically real eigenmodes and
the values calculated by Lamb’s theory are a good approximation. However, in the
more general case of a elastic vibrator embedded in an elastic matrix, some
coupling should occur between the eigenmodes and the propagating acoustic
waves in the matrix. The amount of coupling depends on the elastic mismatch
between spheres and matrix. While the limit case of large mismatch corresponds to
the free eigenmodes, where the elastic energy is totally localized in the spheres, the
other limit case is a zero-mismatch, which means that all energy is coupled
between sphere and matrix and an elastic wave can travel the system like a
homogeneous medium.
To calculate the distinct vibrational modes of elastic spheres with radius rs and
density .s in an elastic matrix with density .m , one has to introduce boundary
conditions. At the surface of the sphere, the displacement u has to be the same
inside and outside the sphere [4, 18]:
The same must be fulfilled for the surface traction sðs ¼ rðrÞ r^; with stress tensor
r and outgoing unit vector normal to the sphere surface r^Þ :
X 1
uðrÞ ¼ xLlm r½Rl ðkl rÞYlm ð^
rÞ þ xM
lm r ½rRl ðkt rÞYlm ð^
rÞ
lm
kl
ð2:111Þ
1
þ xNlm r r ½rRl ðkt rÞYlm ð^
rÞ;
kt
where the x replace either the coefficients a inside or b outside the spheres and kl
and kt are the longitudinal wavenumbers (k = x/c) inside (ki) or outside (km)
the sphere, respectively.
With Eq. 2.111 it becomes clear that Eqs. 2.107 and 2.108 are equivalent to
three scalar equations, each, and the boundary conditions can be expressed as a
system of six homogeneous equations with an infinite numbers of unknowns. Due
to the orthonormality over the spherical surface of the Legendre functions Ylm, it is
possible to decompose each of these equations into l equations, but independent
from m. That is also the coefficients a and b depend only on l, leading to a system
of six equations with six unknowns for each l. It can be broken into two smaller
systems, considering the orthogonality of mlm to nlm and llm ; which contain two
equations involving the coefficients aM M
l and bl and four equations containing the
L L N N
coefficients al ; bl ; al ; and bl ; respectively. Non-trivial solutions for both systems
are only achieved if their determinants are zero, leading to a set of discrete modes
with angular frequencies xnl for each l and n-th order, corresponding to the
expression given in Eq. 2.106.
The considerations presented in this section are the fundamental of theoretical
methods dealing with the scattering of plane waves on single and multiple spheres
presented in Sect. 3.4 in Chap. 3.
References
The assembly of colloids into well defined coherent structures commonly occurs
under the influence of external fields, e.g., gravitational sedimentation [1], elec-
trophoretic deposition [2], or vertical deposition either by evaporation [3] or lifting
the substrate [4] (Fig. 3.1).
In the vertical lifting deposition method upon immersion of a hydrophilic
substrate into a colloidal dispersion a meniscus at the substrate is formed. Evap-
oration takes place at the three phase contact lines (air, dispersion, and substrate),
which causes a solvent flux towards the meniscus, as shown in Fig. 3.2. Thus,
colloidal particles are constantly transported with the liquid to the crystallization
front. An interplay of long-range attractive and short-range repulsive forces causes
the self-assembly of the colloidal material into a face-centered cubic (fcc) or
hexagonally close packed (hcp) crystal [5, 6]. In order to control the thickness of
such a colloidal crystal—a critical parameter for further use in ensuing applica-
tions—the substrate is withdrawn from the dispersion at a certain speed at given
environmental parameters such as temperature and humidity. Besides the fabri-
cation of such ‘simple’ colloidal crystals also more complex systems like binary
[7, 8] and ternary colloidal crystals [9] were demonstrated. The colloidal crystals
that are topic of this article are produced using an enhanced vertical lifting
apparatus by Retsch [10].
Alongside with this research, also the counterpart to the highly ordered crystals—
colloidal glasses [11, 12]—comprising of two distinct, monodisperse latex particles
have already been developed [13]. Doping of the colloidal crystal with removable
moieties (sacrificial templates) opens a pathway to the designed introduction of
defects, which are highly interesting with respect to their contribution to phononic
properties. At the same time, colloidal crystals and glasses serve as templates for the
fabrication of so-called inverse opals [16]. These materials exhibit an interconnected
3D network with high surface area, since the constituent spheres of the colloidal
crystals have been removed. Inverse opals have attracted strong interest in photonic
crystal research, due to their full photonic band gap [16, 17], and are also promising
but unexplored materials for phononic experiments.
Complementary to the vertical deposition method a technique for the fabrica-
tion of large area colloidal monolayers has been recently established. This new
method itself gives access to highly interesting new materials. Multiple stacks of
colloids of various diameters as well as deposition on uneven or curved substrates
have been demonstrated. Controlled etching of such a colloidal monolayer allows
for the first time the fabrication of large areas of non-close-packed structures [18].
Finally, colloidal monolayers are well known and widely used for surface
patterning. Vertical lifting deposition, colloidal monolayer fabrication and the
3.1 Vertical Lifting Deposition and Colloidal Crystals 37
combination of both are powerful tools in order to rationally design new periodic
functional materials with generic possibilities for defect tuning, creation of multi-
layers, colloidal glasses, and inverse opals.
Vertical lifting deposition is a method that delivers high quality colloidal crystals
several tens of microns thick and with cm2 areas. Although, in principle it would
be possible to upscale the method, it would be unpractical and slow to produce
much larger and thicker samples by that. A technique that is much better suited if
fast and large scale hybrid material films of thickness up to the millimeter range is
needed is the melt compression (or compression molding) technique that is
schematically shown in Fig. 3.1f and in more detail in Fig. 3.3a.
Melt compression can be utilized to prepare films out of hard core/soft shell
core–shell particles, where the soft shell becomes the matrix and the hard cores
form a crystalline (usually fcc) lattice, whose spacing depends on the initial vol-
ume ratio of core and shell materials. The preparation starts with the coagulation of
the latex dispersion in a rubbery mass. When the soft shell is an elastomer with low
Tg, e.g., poly(ethyl acrylate) (PEA), the shells form a continuous matrix in which
the cores, e.g., harder polymers like polystyrene or oxides like silica, are dispersed.
At DKI the mass is shaped into a cylindrical sample, which is uniaxially com-
pressed at 170 °C in a Collin 300 press, with an initial pressure of 1 bar and a final
pressure of 50 bar [19]. In the beginning of the pressing process, the first crys-
talline layers appear on the plates of the press. As shown in Fig. 3.3a, the pressure
creates a horizontal flow perpendicular to the direction of compression. The order
increases with pressure by the sequent formation of new crystalline layers above
the already existing layers in order to achieve the densest packing. Finally, the
whole film should be crystalline. It could be shown only recently that in such films
the direction corresponding to the (111) plane of the fcc lattice is radial from the
center of the round press cylinder, i.e., multi-domain ordering must be assumed,
however, the crystallographic directions are well defined [20].
Fig. 3.3 a The flowing melt crystallizes along the plates of the press under uniaxial compression
[19]. b Melt compressed polymer opal film with PS core and PEA shell [20]
38 3 Methods
Figure 3.3 shows a typical film prepared consisting of PS cores and PEA
matrix. It is obvious that in this case the particle size is chosen that way the film
acts as a photonic crystal, since against the dark background the film reflects the
green light under Bragg conditions. Where it is bended, the angle to the light
changes and it is reflected on another crystallographic plane, hence the color
changes.
The aim of this section is to introduce very briefly the different auxiliary tech-
niques applied in this thesis to characterize polymer and colloid samples, including
spectroscopy, electron microscopy, calorimetry, density determination, and size
exclusion methods.
The photon correlation spectroscopy (PCS), also known as dynamic light scat-
tering, is a light scattering technique that can be used to obtain the size distribution
of small particles (d \ k) in diluted solution. When laser light hits small particles,
Rayleigh scattering occurs in all directions. Because the small particles undergo
Brownian motion in the liquid, the relative position of the scatterers within the
scattering volume is constantly changing. This results in a time-dependent scat-
tering intensity fluctuation, because the monochromatic and coherent laser light is
scattered on the moving scatterers. The light scattered from the individual scat-
terers interferes with the light scattered on all the other particles. By using a
correlator at a given scattering angle h, i.e., at a given wave vector q, one can
record the time–correlation function
Z
1 1
Gðq; sÞ ¼ hIðq; tÞIðq; t þ sÞi ¼ lim Iðq; tÞIðq; t þ sÞdt ð3:1Þ
T!1 T 0
with I(q, t) being the scattering intensity at time t and wave vector of magnitude q.
Thus I(q, t) is randomly fluctuating with time due to particle motion, G(q, s) is
decreasing exponentially
from its initial value G(q, 0) to zero for s ! 1—or, to
be exact, from I 2 i to I 2 : The so-called normalized second order autocorrelation
curve is generated from the intensity trace as follows:
hIðtÞIðt þ sÞi
g2 ðq; sÞ ¼ : ð3:2Þ
hIðtÞi2
When the viscosity g of the liquid is known, D gives the radius of the spherical
particles r by the Stokes–Einstein relation
kB T
D¼ : ð3:3Þ
6pgr
All PCS experiments in this thesis were performed by myself on a setup built
in-house at MPIP.
A broader beam is illuminating a part of the sample and the transmitted electron
intensity is imaged onto a fluorescent viewing screen and recorded by a CCD
camera for further computing.
Electron microscope pictures shown in this thesis are recorded by Markus
Retsch, Gabrielle Schaefer, Gunnar Glasser, or Maren Mueller at different setups
at MPIP.
In some cases the exact density of a sample is needed. While for polymers or
colloids the direct measurement of volume and weight might be sophisticated, the
use of a density gradient column is a popular method to measure a sample’s mass
density [21]. The column is a long vertically standing glass, which is filled with a
liquid whose density is increasing gradually from top to bottom. This can, e.g., be
realized by using a special double-flask setup in which two aqueous solutions of a
salt with two different concentrations (and therefore different densities) are mixed
to a solution with continuously changing concentration and density, which is
carefully poured in the column [23, 24]. The density gradient is then calibrated
using several floaters of well known density. Drawing the calibration curve of
floater density against floating hight in the column, the density of other samples
can be determined by flotation.
3.3 Polymer and Colloid Characterization Techniques 41
The density gradient column technique is often used to measure the crystallinity
of a polymer. The degree of crystallization wc is defined as
.c . .a
wc ¼ ð3:5Þ
. .c .a
with ., .c , and .a being the density of the probed sample, the purely crystalline
polymer, and the purely amorphous polymer, respectively. .c is accessible from
the crystal structure, . and .a can be obtained by flotation in the density gradient
column. To obtain a completely amorphous sample, the polymer has to be molten
and quenched to low temperatures with effectual celerity.
Another method to determine the degree of crystallinity is the wide angle X-ray
scattering technique. X-ray structure analysis is in general the most common
method to measure crystal structures. When we took the atoms in the crystal as
points in a real space lattice and transform the lattice into reciprocal space, the
distance between two atoms in real space is now the distance it d between lattice
planes. When a monodisperse, coherent X-ray beam is scattered on parallel lattice
planes, the scattered beams interfere constructively when
2d sin / ¼ nk ðn ¼ 1; 2; 3; . . .Þ: ð3:6Þ
/ is the glancing angle between the incident beam and the scattering plane and
k the wave length of the X-ray beam. Equation 3.6 is the Bragg condition for
X-ray scattering. When scanning a range of scattering angles at some distinct
angles, where the Bragg condition is fulfilled, the constructive interferences occur
as sharp crystalline peaks in the scattering spectrum. In partially crystalline
samples the X-ray spectrum at wide angles will consist of some distinct sharp
peaks but also of an underlaying broad halo, which originates from the amorphous
parts of the sample. Integrating these two areas in the spectrum gives access to the
degree of crystallization, which is
Ac
wc ¼ : ð3:7Þ
Ac þ Aa
With Ac and Aa representing the areas under the curve of the crystalline signals and
the amorphous halo. The discrimination of both parts is highly subjective, which
makes the results less accurate [21].
200 and 1000 nm. In this region of the electromagnetic spectrum, molecules
undergo electronic transitions. The use of UV/VIS spectroscopy in polymer
science lies mostly in the measurement of the concentration of a polymer in
solution. If the solution is diluted, the Lambert–Beer law is valid:
A ¼ logðI0 =IÞ ¼ cL ð3:8Þ
A is the measured absorbance, I0 is the intensity of the incident and I the intensity
of the transmitted light. is the extinction coefficient, a material constant, L is the
pathlength through the sample and c the concentration.
Most UV/VIS spectrometers work with two cuvettes, one with the sample, the
other one as reference containing only the pure solvent, which are compared for
every wavelength. All spectrometers consist of a light source, a dispersive element
(lattice, prism, monochromator filters), the sample, and a detector. Mostly ava-
lanche photo diodes are used. To obtain the spectrum the wavelength range of
interest is scanned and absorbance (or transmittance) is plotted against wavelength.
For structures with periodic pattern of the dielectric constant, e.g., colloidal
crystals, light is diffracted on these pattern with subsequent constructive and
destructive interference. That is electromagnetic waves of distinct wavelengths
cannot pass such structures in distinct crystallographic directions, dependent from
the structure’s periodicity length. This effect is referred to as photonic Bragg gap.
The corresponding wavelength can be calculated as
kBragg ¼ 2sneff sin H; ð3:9Þ
with
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
neff ¼ /s n2s þ /m n2m ; ð3:11Þ
where /s and ns, or /m and nm are the volume fraction and the refractive index of
the spheres or the surrounding matrix, respectively [24].
In this thesis UV/VIS is used to characterize colloidal mixtures. The mea-
surements have been performed by Retsch on the setup at MPIP.
where Kg and ag are parameters depending on the polymer, its shape, and the
solvent. The details of the sophisticated universal calibration is given in standard
textbooks of macromolecular chemistry and physics [21].
single spherical scatterer. From that I will describe the related plane-wave method
to calculate the band structure of periodic composites. Anyhow, the more powerful
method to do the latter and more important for this thesis is the (layer) multiple-
scattering approach, which will be presented concludingly. Finally, the finite
element approach is discussed very briefly.
To calculate the scattering cross-section of a single sphere [25], one can calculate
the scattering of an incident plane longitudinal wave of the form
X 1
inc L
u ðrÞ ¼ clm r½Rl ðkl rÞYlm ð^rÞ ð3:14Þ
lm
ql
The boundary conditions at the surface of the sphere, given above in Eqs. 2.107
and 2.108, must still be fulfilled and can be written now as
and
with
X
usc ðrÞ ¼ L
dlm llm ðR; kli Þ þ dlm
M
mlm ðR; kti Þ þ dlm
N
nlm ðR; kti Þ: ð3:19Þ
lm
It is shown that the coefficients for the incident wave cxlm (with x = L, M, N) are
known a priori [25]. As discussed in Sect. 2.3.3 in Chap. 2, it is possible to
decompose this system into six scalar equations that can be separated into an
infinite number of equations that are functions of lm. When the coefficients cxlm are
known, there remain only the coefficients axlm, given in Eq. 2.109, and dxlm, given in
3.4 Theoretical Calculations 45
Eq. 3.19 for the field inside the spheres or the scattered field, respectively.
Therefore exists for each lm a system of six equations with six unknowns that can
be solved. The detailed algebraic analysis is done for example in Ref. [25].
The scattering cross-section r is a measure for the likelihood of the physical
interaction between an incident particle and another particle—in this case, the
scattering of a wave by the single scatterer. It is defined as the ratio of the scattered
energy flux to the incident energy flux per unit surface [26].
For the discussed case of an incident longitudinal wave (subscript l) the
dimensionless scattering cross-section for a sphere with radius r is shown to
become [27]:
" m N 2#
rl X1 L 2
jdlm j k r jdlm j
¼ 4ð2l þ 1Þ m 2 þ lðl þ 1Þ lm : ð3:20Þ
pr 2
l¼0 jkl rj kt r jktm rj2
As pointed out before (and more widely discussed in Chap. 5), in phononic sys-
tems consisting of many elastic scatterers in a matrix with different elastic prop-
erties, band gaps may occur in the band diagram, originating from either the
periodicity of scattering layers (Bragg gap, BG) or from localized states in indi-
vidual scatterers. The localized modes in the individual scatterers are given by the
maxima in the scattering cross-section diagram discussed in Sect. 3.4.1, and it
seems plausible to extend the method presented for the scattering cross-section
to calculate the band diagram of phononic systems by assuming a plane wave
46 3 Methods
scattered now by many scatterers. In order to obtain analytical solutions, the plane
wave (PW) method is restricted to periodic systems, i.e., phononic crystals [28].
The initial point of the PW approach is the general wave equation for a medium
locally isotropic
i
o2 u i 1 o oul o ou oul
¼ k þ l þ i; l ¼ 1; 2; 3: ð3:21Þ
ot2 . oxi oxl oxl oxl oxi
Here, ui,l and xi,l are the Cartesian components of the displacement vector uðrÞ and
the position vector r; respectively; kðrÞ; lðrÞ; and .ðrÞ are the Lamé coefficients
and the local mass density, respectively, which are periodic functions of r with
periodicity lattice vector R:
f ðr þ RÞ ¼ f ðrÞ: ð3:22Þ
As a result of the common periodicity of the three coefficients in Eq. 3.21, its
solutions can be chosen to satisfy Bloch’s relation
where k is a vector in the reciprocal lattice restricted within the first Brillouin zone
(BZ) and uk ðrÞ is a periodic function. Hence, it is possible to expand f ðrÞ in a
three-dimensional Fourier series:
X
f ðrÞ ¼ fG eiGr ; ð3:24Þ
G
where the summation is over all reciprocal vectors G; which can be written as sum
with integer coefficients of the orthonormalized vectors that span the three-
dimensional vector space in which R is the sum of the basis vectors.
With that one can rewrite Eq. 3.23 after expanding uk in Fourier series as
X iðkþGÞr
uðrÞ ¼ uk þ Ge : ð3:25Þ
G
The substitution of Eq. 3.24 (with f ¼ k; l; .1 ) and Eq. 3.25 into Eq. 3.21
delivers finally the expression:
"
X X
2 i
x cukþG ¼ .1 0 00
GG00 kG G ðk þ G Þl ðk þ G Þi
00 0
G0 l;G
þlG00 G0 ðk þ G0 Þi ðk þ G00 Þl ulkþG0 ð3:26Þ
! #
X Xh i
1 0 00 i
þ .GG00 lG00 G0 ðk þ G Þj ðk þ G Þj ukþG0 :
G00 j
If the Fourier series in Eq. 3.25 is performed over M reciprocal vectors (i.e., for
M scatterers in the periodic medium), Eq. 3.26 is reduced to a 3M 9 3M matrix
3.4 Theoretical Calculations 47
Following the same considerations as for the (general) solid case one obtains for
M terms in the Fourier sum a M 9 M system.
PW is a fast and easy-to-apply method to calculate the band diagram in fluid/
fluid or solid/solid systems, i.e., fluid or solid scatterers in matrices of the same
aggregate state. However, it fails when dealing with solid/fluid systems, which are
mostly discussed in this thesis.
When a solid scatterer is embedded in a liquid host, transverse waves cannot
propagate in the matrix and Eq. 3.27 becomes the appropriate elastic wave
equation. In that case the M eigenmodes for M scatterers correspond to purely
longitudinal waves. However, it is known that even for a longitudinal incident
wave, the field inside the scatterer will be both longitudinal and transverse,
i.e., there are localized transverse modes inside the scatterer that cannot propagate
[27, 29].
The approach that uses the full (including l 6¼ 0) wave equation (Eq. 3.21) fails
because of the non-propagating character of these modes and leads to no con-
vergence at all. Using Eq. 3.27 would ignore completely the transverse component
of the wave within the scatterer as it would de facto replace the solid scatterer by
the fluid scatterers of the same k and l. Although it leads to mathematically
reasonable results, it was shown that these results are not suited to describe the
reality [27].
Another limitation of the PW method is that it can be only applied on infinite
periodic samples. This also means that it is unable to calculate the transmission
properties. An approach to overcome these limitation is the multiple-scattering
method that will be discussed in the next section.
The main idea behind the multiple scattering method is described schematically in
Fig. 3.5. A plane wave in a homogeneous medium impinges on a system of N non-
overlapping scatterers (n = 1, 2, N). The wave that impinges on each scatterer
consists of N contributions, the incident wave and the outgoing waves from any of
the N - 1 other scatterers.
With the coefficient notation used in the sections above, for a single scatterer
the coefficients for the scattered wave field, dxlm, are related to the given coeffi-
cients (i.e., amplitudes) of the incident plane wave, cxlm:
48 3 Methods
x2
x dlm
c lm x3
dlm
t,x1
clm
x4
x6 dlm
dlm
incident
x5
plane wave dlm (x=L,M,N)
Fig. 3.5 Schematic draw of the multiple scattering mechanism. A plane incident wave (with
coefficients cxlm) is impinging an ensemble of scatterers, which can have arbitrary shape. The total
wave field acting on the n-th individual scatterer (with coefficients ct,xn
lm ), e.g., the scatterer 1 in the
middle, is a summation over the impinging plane wave and the scattered waves from the other
scatterers (with coefficients dxn
lm; in our example for n = 1)
X
x
dlm ¼ Txlm;x0 l0 m0 cxlm : ð3:28Þ
x0 ;l0 m0
That is
X
uinc;n ðr rn Þ ¼ ct;Ln t;Mn t;Nn
lm llm ðR; kl Þ þ clm mlm ðR; kt Þ þ clm nlm ðR; kt Þ
lm
X
¼ cLlm llm ðR; kl Þ þ cM N
lm mlm ðR; kt Þ þ clm nlm ðR; kt Þ ð3:30Þ
lm
X L;n0
M;n 0
N;n 0
þ dlm llm ðR; kl Þ þ dlm mlm ðR; kt Þ þ dlm nlm ðR; kt Þ
n0 6¼n;lm
or
3.4 Theoretical Calculations 49
X 1
uinc;n ðr rn Þ ¼ cLlm r½Rl ðkl rÞYlm ðr rn Þ
lm
kl
þ cM lm r ½ðr rn ÞRl ðkt rÞYlm ðr rn Þ
N 1
þ clm r r ½ðr rn ÞRl ðkt rÞYlm ðr rn Þ
kt
X L;n0 1 ð3:31Þ
þ dlm r½Rl ðkl rÞYlm ðrn rn0 Þ
lm;n0
kl
0
M;n
þ dlm r ½ðrn rn0 ÞRl ðkt rÞYlm ðrn rn0 Þ
0 1
N;n
þ dlm r r ½ðrn rn0 ÞRl ðkt rÞYlm ðrn rn0 Þ :
kt
In the last two equations the two sums on the right side belong to the incident plane
wave (with coefficients cxlm, x = L, M, N) and to the scattered waves from the
x;n0
other scatterers (with coefficients dlm ).
For the resulting total incident field the coefficients ct,xn
lm are introduced in
Eq. 3.30. They are related to the scattering coefficients of the n-th scatterer dx,n
lm by
the T-matrix (Eq. 3.28). hP
lmax
Doing so for all N scatterers, one obtains a system of N l¼0 ð2l þ 1Þ þ
P max i
2 ll¼1 ð2l þ 1Þ algebraic equations, in which the only known amplitudes are
those for of the incident plane wave. With the assumption that terms with l [ lmax
do not contribute significantly to the spherical wave expansion, this system can be
x;n
solved numerically to determine all dlm :
So far, the method is not restricted to an equal shape of all scatterers (however,
for different shapes the T-matrix of all individual scatterers would have to be
known) or to any periodicity, as the relative position of each pair of scatterers is
considered in Eq. 3.31.
However, as the aim is to calculate the band structure in phononic crystals as
function of the reduced wave vector k in the first Brillouin zone, the introduction
of the crystal periodicity is needed. It was shown that in this case the phononic
band structure can be calculated in analogy to the theory of Korringa [31], Kohn
and Rostoker [32] (KKR theory) developed to solve the Schrödinger equation for
electromagnetic waves in periodic lattices [33].
After introducing the pressure field
pðrÞ ¼ kr uðrÞ; ð3:32Þ
it is possible to write the wave equation for a periodic medium of solid scatterers in
a fluid matrix as follows [33]:
2 x2 2 1 1 1
r pðrÞ þ pðrÞ þ x 2 pðrÞ þ .ðrÞ r rpðrÞ ¼ 0: ð3:33Þ
cm c ðrÞ c2m .ðrÞ
50 3 Methods
where Hm ðrÞpðrÞ ¼ 0 represents the wave equation for the matrix without scatterer
(Hm ðrÞ ¼ r2 þ x2 =c2m ). In a periodic system a Green’s function approach can be
chosen to reformulate Eq. 3.33.
Z
pðrÞ ¼ Gðr r0 ÞVr0 pr0 dr0 ð3:35Þ
V
and
0
1 eijrr jx=cm
G0 ðr r0 Þ ¼ : ð3:37Þ
4p jr r0 j
Rn is a Bloch’s vector, so that
The next step is the expansion of Gðr r0 Þ and pðr0 Þ into spherical functions of
r and r0 : These calculations are performed in detail in Ref. [33]. In the end the
final multiple-scattering equation appears to be
X x
¼ Alml0 m0 =ðdl10 Þd ll mm al0 m0 ¼ 0:
0 d 0 ð3:40Þ
l0 m0
cm
The coefficients Alml0 m0 are the so-called structure constants as they depend on k; x,
and the periodic lattice structure, the coefficients dl0 relate the incident to the
scattered field.
Equation 3.40 represents a linear homogeneous algebraic system. Its nontrivial
solutions give the eigenfrequencies of the periodic system and hence the dispersion
relation.
An enhanced variant of the multiple-scattering method calculates the dispersion
relation for samples consisting of different composites with two-dimensional
periodicity and is called layer-multiple-scattering approach (LMS) [31, 35, 36].
For a slab parallel to a distinct crystallographic plane, the reduced vector parallel to
this plane, kk ; is usually a conserved quantity. Therefore, LMS searches in each
3.4 Theoretical Calculations 51
individual slab propagating Bloch waves for given x and kk ; which are the
eigenmodes of the elastic field in that slab. It is an on-shell method since it operates
at a given frequency. It was shown that LMS is a powerful method to determine the
dispersion relation as well as the transmittance of three dimensionally structured
systems. Example 3D phononic crystals can be regarded as a succession of planes of
scatterers parallel to a chosen crystallographic plane. As already mentioned for the
general MS approach, the LMS technique takes the full vector nature of the acoustic
field into account and is therefore not limited to certain fluid/solid combinations, as
for example PW is.
Note that nearly all calculations in this thesis using the single-sphere scattering
cross-section and the multiple scattering algorithms have been performed by my
co-worker Revekka Sainidou (University of Le Havre).
References
3. Jiang P, Bertone JF, Hwang KS, Colvin VL (1999) Chem Mater 11:2132
4. Gu ZZ, Fujishima A, Sato O (2002) Chem Mater 14:760
5. Fustin CA, Glasser G, Spiess HW, Jonas U (2004) Langmuir 20:9114
6. Denkov N et al (1992) Langmuir 8:3183
7. Wang L, Baowen L (2008) Phys World 21:27
8. Tommaseo G et al (2007) J Chem Phys 126:014707
9. Wang J, Li Q, Knoll W, Jonas U (2006) J Am Chem Soc 128:15606
10. Fustin C-A, Glasser G, Spiess HW, Jonas U (2003) Adv Mater 15:1025
11. Pusey PN, van Megen W (1987) Phys Rev Lett 59:2083
12. García P, Sapienza R, Blanco A, López C (2007) Adv Mater 19:2597
13. Li Q et al (2008) Porous networks through colloidal templates topics in current chemistry.
Springer, Heidelberg
14. Still T et al (2008) Phys Rev Lett 100:194301
15. Zakhidov AA et al (1998) Science 282:897
16. Li Z-Y, Zhang Z-Q (2000) Phys Rev B 62:1516
17. Li ZY, Zhang ZQ (2001) Adv Mater 13:433
18. Haginoya C, Ishibashi M, Koike K (1997) Appl Phys Lett 71:2934
19. Ruhl T, Spahn P, Hellmann GP (2003) Polymer 44:7625
20. Spahn P (2008) PhD thesis
21. Pursiainen OLJ et al (2008) Adv Mater 20:1484
22. Lechner MD, Gehrke K, Nordmeier EH (2003) Makromolekulare Chemie, 3rd edn.
Birkhäuser, Basel
23. Linderstrom-Lang K (1937) Nature 139:713
24. Toth H, Fehlauer H (2004) Phys Unserer Zeit 35:76
25. Gaillot DP, Graugnard E, King JS, Summers CJ (2007) J Opt Soc Am B 24:990
26. Psarobas IE, Stefanou N, Modinos A (2000) Phys Rev B 62:278
27. Cheng W (2007) Hypersonic elastic excitations in soft mesoscopic structures. PhD thesis
28. Kafesaki M, Economou EN (1995) Phys Rev B 52:13317
29. Economou EN, Sigalas M (1994) J Acoust Soc Am 95:1734
30. Einspruch N, Truell R (1960) J Appl Phys 31:806
31. Sainidou R, Stefanou N, Psarobas IE, Modinos A (2005) Comp Phys Commun 166:197
32. Korringa J (1947) Physica 13:392
33. Kohn W, Rostoker N (1954) Phys Rev 94:1111
34. Kafesaki M, Economou EN (1999) Phys Rev B 60:11993
35. Liu ZY et al (2000) Phys Rev B 62:2446
36. Sainidou R, Stefanou N, Psarobas IE, Modinos A (2005) Z Kristallogr 220:848
Chapter 4
The Vibrations of Individual Colloids
4.1 Introduction
Nowadays, colloidal nano- and mesoscale particles with dimensions from few
nanometers up to 1 lm are used in a growing number of applications, e.g., as
fillers in polymer thin films to enhance thermo-mechanical properties [1], to
improve coatings performance and as components in nanocomposites operating as
photonic [2], plasmonic [3], and phononic structures [4]. For a wide range of such
applications, information on the mechanical properties and the stability of these
colloidal composite materials are of paramount importance. Conventional rheo-
logical measurements on the macroscopic system are often not sufficient to elu-
cidate the specific contributions of the nanostructured components. At the
nanoscale, forces negligible in macroscopic systems, such as depletion, interfacial,
and confinement effects often become significant, and the behavior of the same
materials in nanoscopic systems can considerably deviate from the bulk. A fun-
damental understanding of transport and thermomechanical properties of nano-
structured materials is a precondition to address a specific need by structural
engineering. In the case of colloidal composite materials, the vibrational modes
confined to the individual particles result from the elastic motion at the nanoscale
and should sensitively depend on the geometrical, architectural, interfacial, and
mechanical characteristics of the particles. However, there is a paucity of non-
destructive experimental techniques to probe this ‘music’ of particle vibrations
since both high frequency resolution and sensitivity are required to detect the
numerous eigenmodes. Raman scattering [5, 6] has been utilized to measure few
eigenfrequencies of nanoparticles with dimension below 10 nm, whereas Brillouin
light scattering (BLS) [7–9] and optical pulse-probe techniques [10–12] can probe,
respectively, the spontaneous and stimulated vibrations confined in sub-microm-
eter particles. In the latter technique, the excited acoustic oscillations are observed
in the form of modulations of the transient reflectivity of the probe laser, and hence
the particles must possess good reflectance, e.g., by introduction of gold shells. In
BLS, light is scattered inelastically by the density fluctuations (phonons)
associated with these particle localized modes at thermal equilibrium and there are
no further stringent conditions.
Self-assembly of colloidal particles in periodic structures [13, 14] (cf. Sect. 3.1
in Chap. 3) has received special attention as the resulting photonic and phononic
crystals have revealed the potential of manipulating the propagation of electro-
magnetic and elastic waves. Their propagation is forbidden at ‘Bragg’ frequencies
or wavelengths commensurate with the lattice constant, which for sub-micrometer
particles is comparable with the wavelength of the visible light. Synthetic opals
from these particles can exhibit dual, i.e., hypersonic phononic and photonic
[14, 15] band gaps allowing for acousto-optical interactions [16]. Moreover, the
design of sub-micron particles that can act as strong localized resonant elements in
an appropriate matrix provides the possibility for additional gaps well below the
Bragg frequency, termed hybridization gaps (cf. Chap. 5) [17, 18]. The opening of
band gaps perturbs the phononic density of states which impacts physical quan-
tities such as group velocity, heat capacity and heat conductivity in dielectrics
being potentially useful for thermoelectric devices [19, 20]. One of the pivotal
concerns for these systems is the phonon dispersion, which is essentially defined
by the elastic parameters of the constituent components and the spatial architecture
of the composite system. Colloid science can create novel materials that possess
spatial variation of density and elastic constants at the nanoscale but their
mechanical characterization remains difficult.
If vibrations are excited in a finite elastic body, it can act as an elastic resonator.
The elastic standing waves in stress-free boundary are referred to vibration
eigenmodes. In the case of elastic spheres, the analytical solutions were first
derived by Lamb [21]. The eigenmodes of spheres can be classified as torsional
and spheroidal modes, both labeled by three indices (nlm), which describe the
angular (lm) and radial (n) dependence of the displacement. Spheroidal resonance
modes are fully characterized by the angular momentum l, imposed by the
spherical symmetry of the particle, and n, where n denotes the n-th order solution
for a given l (cf. Sect. 2.3.3 in Chap. 2) [21, 22]. The use of inelastic light
scattering to measure vibration eigenmodes of small spherical particles was first
performed experimentally by low-frequency Raman scattering (RS) [22]. Due to
selection rules only two distinct vibration eigenmodes contribute to the RS of
spherical particles with diameter (d) much smaller than the wavelength of the
probing light (d k) [23]. For bigger spheres with d k, Brillouin light
scattering (BLS) in the GHz-range becomes the technique of choice [7, 8, 24]. Due
to the consideration of higher-order terms in the electric multipole expansion and
of retardation effects [25, 26], BLS can resolve a multitude of eigenmodes.
For homogeneous transparent systems, BLS measures the spectrum of light
inelastically scattered by the acoustic phonons with a selected polarization
[longitudinal (l) or transverse (t)] and wave vector q, leading to spectra consisting
of doublets at xB = ±cq around the central elastic Rayleigh line. Since c in soft
materials like polymers is of the order of 103 ms-1 and q is in the range
1-30 lm-1, the frequencies f = xB/2p fall into the GHz range. This yields the
4.1 Introduction 55
two elastic constants C11 ¼ .c2l and C44 ¼ .c2t with cl and ct being the two phase
sound velocities and . being the mass density. For inhomogeneous turbid systems,
e.g., powder of mesoscopic (d k) particles, q is ill-defined in the BLS experiment
due to strong multiple scattering. BLS can measure only localized in space (and
hence q-independent) vibrational modes. Each resonance mode appearing at
frequency f(n, l) is characterized by the angular momentum l of the n-th order. The
frequencies of the individual vibrational modes depend on their rigidity, mass
density, and size dimensions of the particles. For the case of homogeneous
spherical particles the frequencies are given by Lamb in Eq. 2.106.
The frequencies can be theoretically obtained from the calculated density of
states (DOS) spectra of a single sphere as a function of the two elastic constants
and the inverse diameter. For polystyrene spheres, e.g., the constant in Eq. 2.106
becomes x(1, 2) & 0.85 and R(cl, ct) = C44 with no adjustable parameter [8].
In this chapter the state-of-the-art for BLS measurements on homogeneous
mesoscopic spherical particles is briefly summarized. Novel studies dealing with
mixtures (‘hybrids’) of different kind of spheres are presented as well as a study on
spheres prepared as copolymers with different compositions to elucidate the
influence of the next neighbors and of the rigidity on the mechanic vibrations in
the mesoscale. The second part of the chapter extends the scope on nanostructured
colloids. Hybrid material spherical core–shell particles are investigated as model
systems. Therein, especially the influence of heat on polymer cores contained in a
hard silica-shell and the influence of the composition in the vice versa case of
silica–PMMA core–shell particles with different ratio of core size to total diameter
on the vibrational eigenmodes is of interest.
Due to some theoretical support, these studies give an unprecedentedly com-
prehensive picture of the elastic properties of individual colloids, although still
numerous questions have to stay open—e.g., when dealing with structures going
beyond the investigated ‘easy’ spherical model systems.
As pointed out in Sect. 2.3.3 in Chap. 2 and in the Introduction of this chapter,
BLS can be utilized to measure the resonance modes of dry non-transparent col-
loidal crystals. Due to the strong elastic form factor of the individual spheres and
the large elastical contrast with the surrounding air, the opals show strong multiple
scattering. In such samples the inelastic scattering from localized modes leads to
incoherent BLS in analogy to the Raman scattering.
Thus BLS can be utilized to analyze the particle eigenfrequencies, describing
the spheroidal (n, l)-modes, with n as the n-th mode of the l-th spherical harmonic.
The first demonstration of the feasibility of the BLS experiment was shown by
Penciu et al. in the case of dilute suspensions of giant core–shell micelles [9]. Few
56 4 The Vibrations of Individual Colloids
years later Kuok et al. have extended this application to closely packed mono-
disperse silica nanospheres in air [7, 27]. Up to six localized particle eigenmodes
have been resolved out of the numerous possible modes, probably due to the weak
scattering of moderately compressible silica. In a subsequent study, artificial soft
colloidal crystals, composed of monodisperse submicrometer polystyrene (PS)
spheres with diameter d between 170 and 856 nm have been investigated [8]. Up
to 21 q-independent eigenmodes have been resolved.
Figure 4.1 shows exemplary three BLS spectra taken from colloidal PS opals
with diameters 180, 360 and 550 nm [24]. The resonance frequencies scale with
1/d (Fig. 4.2), which is in perfect agreement with Lamb’s theory [21] and the the-
oretical predictions based on single-phonon scattering cross-section calculations
(cf. Sect. 3.4.1 in Chap. 3) [28, 29]. In the computations, a plane sound wave
propagating in air and impinging upon a single PS sphere was considered and after
subtracting the scattering amplitude for a rigid sphere of equal size, the sphere
eigenmodes appear as resonance peaks in the plot of scattering cross-section
versus frequency. Thereby the resonance frequencies f(n, l) can be identified as
mode with angular momentum quantum number l of n-th order.
Using the experimental values for the longitudinal sound velocity
cl = 2350 ms-1, the transverse sound velocity ct = 1210 ms-1 and mass density
. ¼ 1050 kg m3 of bulk PS, all resolved frequencies are quantitatively captured
within 3% with no adjustable parameter. The product of frequency and diameter is
Fig. 4.1 Left SEM—images of colloidal PS crystals with d = 180, 360 and 550 nm (top to
bottom); right corresponding q-independent BLS eigenmode spectra of the PS opals. To capture
all possible vibrations, two spectra recorded at two different free spectral ranges are superim-
posed. Spectra are recorded at 20 (q = 0.0041 nm-1). For the thickest particles (d = 550 nm,
bottom) there is a clear cut-off bump around 15 GHz, which is the frequency of the acoustic
phonon in PS at backscattering geometry [24]
4.2 Elastic Vibrations in Homogeneous Polymer Colloids 57
Fig. 4.3 FEM simultation for the (1,2)-eigenmode of a PS sphere with d = 540 nm at
1.87 GHz; a total displacement (color scale) and deformation of the surface, b–d displacement
and deformation of intersecting planes in the three planes of a Cartesian coordinate system.
Deformations are amplified by a factor of 20
Fig. 4.4 FEM simulation for the higher l eigenmodes (n = 1). Deformations are amplified by a
factor of 20. Top total displacement (color scale) and deformation of a the surface and b–d the
intersecting planes of the (1,3) spheroidal mode. Bottom displacement and deformation of the
surface for modes with l = 4–7. For a PS sphere with d = 540 nm the frequencies are found to be
2.79 (1,3), 3.58 (1,4), 4.32 (1,5), 5.03 (1,6), and 5.74 GHz (1,7) in good agreement with the
experimental data (Fig. 4.2)
found by FEM, however, with increasing order the assignment becomes more
and more complicated, since in the calculation mostly asymmetric mixed modes
appear. Theoretically, with much finer mesh, all modes can be assigned clearly.
Anyhow, even the six modes shown here (in perfect accordance with another
calculation method) proof that there is no limitation on even ls as claimed by Li
et al. [26], since the assignment of the (1,3) and the (1,5)-mode with no fitting
parameter is clear and without any alternative. On the other hand, the FEM
simulation of the eigenvibrations of silica spheres with d = 360 nm, whose
eigenmode spectrum is shown in Ref. [26], allows other assignments for the first
four modes than that chosen by the authors, who restricted themselves on even
ls, utilizing the elastic parameters given by the same authors experimentally
(ct = 2520 ms-1, cl = 3960 ms-1, . = 1960 kg m-3) [31]. The mode at
8.76 GHz [assigned as (1,0)] could be the (1,3)-mode (8.60 GHz by FEM), the
mode claimed to be the (1,6)-mode at 14.4 GHz could be a double signal
containing the modes (1,5) and (1,6), found by FEM to appear at 13.14 and
15.29 GHz, respectively. For higher orders, the assignment is more difficult,
however, the mode at 17.65 GHz [claimed to be (2,4) or (3,2)] could be
identified as the (1,7)-mode, calculated to appear at 17.4 GHz.
It should be noted that deviating from Ref. [24] in Fig. 4.2 also the breathing
mode (1,0) is shown as a theoretical fit based on FEM calculations, again in perfect
agreement with the scattering cross-section method [32]. It is found that for the
case of PS spheres the (1,0)-mode is not far away from the (1,4). In fact, the third
signal (from low to high frequency) for the PS spheres in Fig. 4.1 seems slightly
broadened, which might be related to a weaker signal from the (1,0)-mode.
4.2 Elastic Vibrations in Homogeneous Polymer Colloids 61
The assignment of the two phononic gaps is also corroborated by their sensitivity on
the disorder. In this context, non-crystalline colloidal films were prepared by vertical
lifting deposition of ‘hybrids’, binary mixtures consisting of an equal number of two
PS spheres with different diameter (d = 300 and 360 nm). The size polydispersity is
then artificially increased and no crystallization takes place, which is affirmed by
SEM-pictures. The eigenmode acoustic spectra of these dry hybrid films and the
dispersion relations in their infiltrated counterparts were measured by BLS. The
eigenmode spectrum of the 1:1 300:360 nm hybrid is shown in Fig. 4.5 along with
the eigenmode spectra of the individual one component opals. Interestingly, the
spectrum of the hybrid is a superposition of the individual opals as is indicated by the
vertical lines denoting some exemplary resonance frequencies either from the small
or the big spheres. Moreover, hybrids of different relative ratios of the constituent
spheres were prepared. The intensity of the signals originating from a particular
particle size relates to its composition in the mixture. In Fig. 4.5, this linear
composition dependence is demonstrated by superimposing the spectrum of the 1:1
300:360 nm hybrid with the spectrum of a 3:1 300:360 nm hybrid, normalized to the
peak intensity in the 360 nm spheres. For clarity a baseline correction was
introduced into the Stokes-side of the hybrids (inset to Fig. 4.5).
Fig. 4.5 Left eigenmode acoustic spectra of the PS opals with d = 300 nm (blue), d = 360 nm
(red) and the symmetric PS hybrids 1:1/3:1 300:360 nm (black/green)(bottom to top). The ver-
tical lines (dotted for d = 300 nm, dashed for d = 360 nm) denote some exemplary eigenmodes
appearing in the opals and in the hybrids. In the top right corner the Stokes-sides of the hybrids’
spectra are compared after baseline correction to better visualize the influence of the composition
on the relative intensity of the individual signals. Right corresponding SEM-images of opals and
hybrids shown on the left side [24]
62 4 The Vibrations of Individual Colloids
Fig. 4.6 Left comparison between the eigenmode acoustic spectra of the 550 nm PS single opal
and the highly asymmetric 9:1 180:550 nm PS hybrid. Right SEM-image of the 9:1 180:550 nm
PS hybrid [24]
4.2 Elastic Vibrations in Homogeneous Polymer Colloids 63
In summary, the number of next neighbors does not influence significantly the
eigenvibrations of the individual spheres. The effect of the disorder on the
dispersion relation of our PS hybrids after infiltration is discussed in Sect. 5.1.3
(in Chap. 5).
While there are several BLS studies dealing with the size-dependence of the
eigenmodes of homogeneous spheres from different materials (Sect. 4.2.1) [7, 8, 24],
where the elastic properties are calculated from the experimental results, as well as
two more elaborate studies on core–shell particles, presented also in this thesis
(Sect. 4.3) [34, 35], a systematic experimental investigation of the influence of the
particle rigidity on the eigenvibrations of mesoscopic copolymers is still missing.
After Lamb (Eq. 2.106)
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi!
Rðcl ; ct Þ
f ðn; lÞ ¼ xðn; lÞ d;
.
very similar r values (nBA: r = 0.94; MMA: r = 0.91) not very different from
r = 1.
Differential scanning calorimetry (DSC) was used to determine the glass
transition temperature Tg. The results of the heating period of the second run, i.e.,
after heating initially to 200 C (and therefore bulk properties), are shown in
Fig. 4.7. Tg decreases rapidly with increasing amount of nBA. Using the value
measured for pure PMMA and the literature value for nBA (-43 C) [38], allows
an excellent representation of the glass transition temperature as a function of the
composition by the Fox–Flory equation [39]
1 / /
¼ PMMA þ PnBA ð4:1Þ
Tg Tg;PMMA Tg;PnBA
with no fit parameter (inset). The Gordon–Taylor fit also shown in the inset is
borrowed from Ref. [38], including the empirical parameter K = 0.82, however,
the small deviations originate partially in different values measured for pure
PMMA. Our Tgs could be perfectly represented with Tg(PMMA) = 113 C and
K = 0.65. In addition to the single Tg, dielectric spectroscopy has been performed
by my colleague Mpoukouvalas, which showed a single a-relaxation for each
sample, i.e., both kind of segments feel the same energy landscape.
The density . of the copolymers was determined by a density gradient column
using aqueous calcium nitrate solutions [40]. The experimental results shown in
Fig. 4.8 conform to a weight average density, using for the density of the
homopolymers .PnBA ¼ 1:035 g cm3 and .PMMA ¼ 1:195 g cm3 .
The size of the spheres is ascertained by scanning electron microscopy (SEM).
The average molecular weight Mn and the polydispersity index PDI of the
polymers were measured relative to PMMA standards by gel permeation
chromatography (GPC). Crystallinity was excluded by small angle X-ray
scattering measurements. The particle properties are itemized in Table 4.1.
4.2 Elastic Vibrations in Homogeneous Polymer Colloids 65
Table 4.1 Particle and material properties of the investigated spheres obtained by SEM (d), GPC
(Mn, PDI), DSC (Tg) and density gradient column(.)
nBA (wt%) d (nm) Mn (g mol-1) PDI Tg (C) . (g cm-3)
0 232 92,500 4.2 113 1.195
10 214 89,500 4.2 92 1.179
20 200 84,200 3.2 68 1.163
30 204 110,000 3.5 50 1.147
mesospheres. The polarized spectra delivered cl, but the intensity of the theoretically
accessible ct in the depolarized spectra was not strong enough to determine a single
signal. The measured cls are presented in Table 4.2, together with the value of a bulk
PnBA sample.
Previous studies on homogeneous mesoscopic polymer and silica beads show
that there should be no significant change in the longitudinal sound velocity that
has to be taken into account in the colloids compared to that in the bulk material
[8, 24, 31]. Since the rigidity R for the lower frequency eigenmodes is found to be
a much stronger function of ct, i.e., shear modulus G ¼ .c2t , than of cl, i.e.,
longitudinal modulus M ¼ .c2l (cf. Sect. 2.1.2 in Chap. 2), it is meaningful to take
the bulk longitudinal sound velocities also as given in the description of the
vibrational modes.
Table 4.2 shows the decrease of cl, i.e., softening, in the copolymers with
increasing amount of nBA. However, this decrease is small compared to the
experimental value of cl in the two homopolymers. In Sects. 5.1.2 and 5.2 in
Chap. 5 of this thesis, several approaches to describe the sound propagation in an
effective medium composed of (at least) two mechanically different materials are
discussed. It turns out that the simple expression
Table 4.2 Experimental and theoretical bulk cl for homopolymers and copolymers at room
temperature
nBA (wt%) cl(BLS) (ms-1) cl(Wood) (ms-1) cl(Wood) (ms-1)
cl,PnBA = 2200 ms-1
0 2,755 2,755 2,755
10 2,678 2,569 2,663
20 2,615 2,421 2,584
30 2,498 2,300 2,515
100 1,835 1,835 2,200
4.2 Elastic Vibrations in Homogeneous Polymer Colloids 67
1 / 1 /1
¼ 1þ ; ð4:2Þ
Meff M1 M2
which is known as Wood’s law [43], is in many cases a good approximation to
give an effective modulus Meff as a function of the components individual moduli
and the composition. If one assumes Wood’s law to be, in principle, a good
approximation also for copolymers, the theoretically expected values for cl can be
easily calculated, as shown in the second column of Table 4.2. Obviously, the
presented approximation strongly overestimates the softening due to the nBA in
the copolymers, leading to a deviation of up to 8% for 30 wt% nBA. It will be
shown in Sect. 4.3.1 that the mechanical moduli are temperature dependent,
strongly decreasing at temperatures above Tg. At room temperature, all sound
velocities except for bulk PnBA are measured in the glassy state, i.e., for purely
elastic response. In contrast, bulk PnBA at room temperature is in the rubbery
regime, i.e., it behaves viscoelastically and hence its cl is lower than for a glassy
PnBA. In other words, the nBA segments in the three glassy PMMA–r-PnBA
copolymers assume a dense packing and therefore should display elastic response
due to the very slow (essentially frozen) dynamics. The success of Eq. 4.2 in
representing the experimental cl values of the copolymers is optimized if a fictive
cl = 2200 ms-1 is used for the bulk PnBA, i.e., by introducing an ‘effective glassy
longitudinal sound velocity’. The third column in Table 4.2 shows the good
representation by Wood’s law, using this artificially increased value.
Returning to the eigenmode spectra in Fig. 4.9, the fixation of d, ., and cl to
their experimental values reduces the number of adjustable parameters in the
single phonon scattering calculations (Sect. 3.4.1 in Chap. 3) or FEM (Sect. 3.4.4
in Chap. 3) into one (ct).
Scattering cross-section calculations were performed by my coworker Sainidou
[29, 44]. Utilizing reasonable values for ct, the first three modes could be clearly
assigned as the (1,2), (1,3), and (1,4)-mode, in analogy to the results found in
polystyrene and silica samples in Sect. 4.2.1. However, the fourth signal, which is
much broader than the others, cannot be assigned to the expected (1,5)-mode, since
it appears at higher frequencies. Indeed, its frequency is found between those of
the (1,5) and the (1,6)-mode, indicating that the signal is a superposition of both
modes. This explains also the broad appearance of the signal, although a fit with
two Lorentzians does not converge. Therefore, instead of pretending a meaningful
fit of two individual modes, one must be content to show the signal in between
these two modes. The ‘fifth’ mode, on the other hand, can be assigned as the
(1,7)-mode with acceptable accuracy, by choosing ct given by the mean square fit
of only the first three modes. In fact, for this last mode, the deviation from the
theory decreases with increasing amount of nBA in the copolymer.
All experimental modes are shown in Fig. 4.10 together with the fitted values
for fd, using ct as the only floating parameter. The linear fits have no theoretical
meaning, however, they follow the trend of most signals quite well, although the
change between 0 and 10 wt% does not follow the linear trend for the first two
modes. For the (1,7)-mode (‘mode 5’), however, the change between 20 and
68 4 The Vibrations of Individual Colloids
30 wt% is quite large regarding the virtually unaltered fd-values between pure
PMMA and 20 wt% nBA. Correspondingly, the relative deviation between
experiment and theory is largest (&4%) for the (1,7)-mode and pure PMMA. On
the other hand, a linear fit through only the 20 and 30 wt% points captures the
calculated points with a deviation \2%.
From the experimentally obtained material parameters, in combination with the
fitted values for ct, it is possible to calculate the elastic moduli of the mesospheres
and by that to quantify their softening due to the addition of nBA. The calculated
values for the elastic moduli as well as for Poisson’s ratio can be found in
Table 4.3.
As expected, all moduli go down with increasing amount of nBA, indicating the
strong softening effect (by &20% between pure PMMA and the softest copolymer).
The more striking feature is the constance of Poisson’s ratio r. In a copolymer with
increasing amount of a rubber-forming monomer (and consequently strongly
decreasing Tg), one could expect a more rubber-like behavior of the copolymer, i.e.,
an increasing r. At room temperature, i.e., about 30 C below the lowest Tg,
however, this is not the case. This finding means that the compressibility
Table 4.3 Fitted ct, elastic moduli (Young’s, longitudinal, and shear modulus), and Poisson’s
ratio
nBA (wt%) ct (ms-1) E (GPa) M (GPa) G (GPa) r
0 1,530 7.14 9.07 2.80 0.277
10 1,520 6.88 8.46 2.72 0.262
20 1,466 6.35 7.95 2.50 0.271
30 1,394 5.68 7.15 2.23 0.274
4.2 Elastic Vibrations in Homogeneous Polymer Colloids 69
In Sect. 4.2.1, I pointed out that, theoretically, the scattering intensity of the
individual modes should be strongly q-dependent, however, due to the strong
multiple scattering all qs corresponding to values between zero and backscattering
contribute equally to the spectrum and hence the intensity distribution is the same
independent from the scattering angle h. Based on symmetry arguments, Montagna
showed that for a given n the BLS intensity of modes with increasing l are only
active for increasing values of qd [25]. In fact, for a given system, the intensity
maximum of each mode (n, l) should appear at a different q, i.e., for a transparent
system that allows the observation of eigenmodes at different scattering angles h,
while the angle of maximum intensity hmax increases with l.
It was shown first in 2003 by Penciu that BLS can be used to measure the
eigenmodes of silica colloids suspensions in a refractive index matching liquid
(cyclohexane/decalin) [29, 45]. By measuring the suspension in an NMR-tube the
sample can be regarded as quite thick, which allows to see also eigenmodes that
are not resolved anymore in the systems presented in Chap. 5. It is shown that the
appearance of the eigenmodes depends on the filling fraction of the silica spheres
in the suspension. In fact, quite high filling fraction are needed to resolve the
eigenmodes at all. Those have been achieved by centrifugation of the suspension at
the bottom of the NMR-tube. The graphs in Fig. 7 of Ref. [29] show the resolved
modes as a function of qd for two different silica particles with diameter
d = 328 nm and 250 nm, respectively, at three different filling fractions (/ = 0.3,
0.62, and 0.67). For all systems, an effective medium acoustic branch is observed
that will not be further regarded in this section. Besides that, there are up to four
modes that appear at the same frequency for each qd, resonance modes of the silica
spheres. While at / = 0.3 only one such mode is found, in the polycrystalline or
glassy samples with significantly higher / more modes can be distinguished.
Although in Ref. [29] some exemplary spectra at different qd give evidence that
the q-dependence of the BLS intensity can be shown by these samples, a sys-
tematic investigation is still missing. Therefore, in this section the results of BLS
measurements on silica particles with d = 375 nm in an index matching matrix of
70 4 The Vibrations of Individual Colloids
(b)
4.2 Elastic Vibrations in Homogeneous Polymer Colloids 71
(silica) spheres [7, 8, 24]. Due to the small expansion coefficient of fused silica
(5.5 9 10-7 K-1), the diameter d in Eq. 2.106 is virtually temperature
independent.
The heating experiment was performed at a non-arbitrative scattering angle of
50 while the sample was encased in a glass windowed oven (cf. Fig. 2(b) in
Ref. [51]). The sample was heated from room temperature to 165 C in steps
between 30 (low T) and 10 K (high T) within a few minutes for each step. After
equilibrating the sample for 15 min at each temperature, the eigenmode spectrum
was recorded for about 15 min probing always the same spot ( 50 lm). This
ensures that even the intensities of the main modes can be traced as a function of
temperature. Figure 4.14a shows eigenmode spectra at different temperatures upon
heating and after cooling over a narrower frequency range than in Fig. 4.13, in
order to further boost the resolution. All recorded spectra have been represented in
the range of the (1,2)-signal by a Lorentzian using its frequency, line width, and
amplitude as adjustable parameters. The (1,2)-mode is red shifted by &0.2 GHz at
165 C relative to its frequency at room temperature either before (bottom) or after
(top) heating, indicating a small decrease of shear modulus (C44). Note that the
same mode undergoes a much smaller shift at 70 C.
The experimental frequencies f(1,2) shown as a function of temperature in
Fig. 4.15 are well represented by two distinct straight lines, expectedly both lines
with negative slopes due to decreasing rigidity. The characteristic kink in the f-T
diagram signifying the transformation of a glassy into a rubbery state occurs at
107 ± 3 C, which is slightly higher than the Tg of bulk amorphous PS (100 C)
due to the crosslinking of the PS core with divinylbenzene (DVB) [54]. The
presence of a single Tg in Fig. 4.15 indicates a homogeneous, over few hundreds
nanometers, core environment and hence this value of Tg is further confirmed by
differential scanning calorimetry (DSC) measurements on the bare PS cores and
the PS–SiO2 particles. In both systems Tg was found to amount to 107.5 ± 3 C,
whereas in the absence of crosslinking Tg lowers to &100 C (see Fig. 4.15b), in
74 4 The Vibrations of Individual Colloids
(a) (b)
Fig. 4.14 a Exemplary eigenmode spectra during heating and after cooling (side arrows). The
solid lines in the Stokes side of the spectra indicate the representation of the two low frequency
modes by Lorentzian spectral lines (individual modes and sum). The dashed line indicates the
position of the (1,2)-mode at room temperature. b SEM pictures of the core–shell particles before
and after heating to 165 C and cooling back to room temperature [35]
agreement with DSC. The clear detection of the glass transition affirms the validity
of the obtained findings.
The relative drop of f(1,2) between room temperature and Tg is less than 2% in
the PS–SiO2 particles and more than 6% in the bare PS cores (cf. Fig. 4.16). In the
rubbery state of the core, f(1,2) at the highest examined temperature (165 C) is
decreased relatively to its value at Tg only by about 6%. Assuming a constant
density, this f(1,2) value corresponds to a shear modulus c44 & 1.45 GPa in the
confined polymer melt at about Tg + 60 K. For comparison, a corresponding
decrease in the transverse sound velocity of PS films is about 40% (i.e.,
c44 \ 1 GPa) [51]. This remarkable small temperature effect on the rigidity of the
present three-dimensionally confined PS–SiO2 particles as a result of the increased
pressure in the core is unprecedented.
Even at temperatures far above Tg, the spectra (Fig. 4.14a at 165 C) still
possess the typical shape for T \ Tg with minor changes in the intensity. This
4.3 Elastic Vibrations in Nanostructured Colloids 75
(b)
proofs that even in the rubbery regime the PS core keeps its spherical shape. The
thermal expansion of the PS core is expected to increase by more than a factor two
above Tg and hence the pressure on the silica shell [55]. Especially at temperatures
well above Tg, it is this pressure that could, in principle, break the shell. However,
when heating up to 165 C, the intensity of the observed mode changes only very
little within the experimental error. After cooling back to room temperature, not
only the frequency (dashed line in Fig. 4.14) but also the intensity and the line-
width (&0.39 GHz full width half maximum) assume their initial values before
heating. Thus the thin silica shell can stand the higher pressure under these con-
ditions and prevents the PS core to change its shape. This notion is corroborated by
the scanning electron microscope (SEM) photographs of the core–shell particles
before and after the heating cycle to 165 C shown in Fig. 4.14b. The lack of
visible defects after the heating procedure confirms the results obtained by the
BLS. Actually, the silica shell acts as an armor for the PS core.
In Fig. 4.16, the corresponding BLS experiment on bare PS particles is pre-
sented. These eigenmode spectra are very sensitive to the particle shape. For the
bare PS particles f(1,2) displays stronger decrease with temperature than for the
76 4 The Vibrations of Individual Colloids
PS–SiO2 core–shell particles (inset to Fig. 4.16) due to the absence of the
confinementand partially to the now unsuppressed expansion of the PS spheres,
i.e., decrease of c44 and increase of d in Eq. 2.106. Already below Tg the line shape
of the (1,2)-mode changes and above Tg the signal vanishes completely within a
few minutes. Starting at room temperature, for increasing temperature a shift to
lower frequencies and a broadening/splitting of the signal is observed. The spectral
shape of the (1,2)-mode (around 2.5 GHz) in the BLS spectra of the bare PS core
in Fig. 4.16 severely changes as the temperature increases towards Tg. At
temperatures above 85 C, the single Lorentzian shape initially splits into two
through the appearance of a second peak at the high frequency side. Finally, above
103 C the intensity decreases rapidly and eventually vanishes at 108 C within a
few minutes. Apparently, shape alterations start already below Tg because of
surface melting and breaking the particles spherical symmetry. With the onward
and patchy destruction of the spherical shape, the resonance modes of the unarmed
PS particles finally disappear, and the destruction of the spheres is, of course,
irreversible. SEM pictures show only an undefined surface. Probing the shape of
the PS mesospheres on the average during the melting process is another inter-
esting but demanding task. In the context of the present study, it would require
finite element modelling of all eigenfrequencies supported by a temperature
dependent SEM or atomic force microscopy.
4.3 Elastic Vibrations in Nanostructured Colloids 77
Colloid science can create novel materials that possess spatial variation of density
and elastic constants at the nanoscale but their mechanical characterization
remains difficult. Similarly, the influence of the particle architecture (core–shell
spheres or hollow capsules), size, and shape on the eigenfrequencies, and the
localization of elastic energy in specific regions of an individual particle is not
known. In addition to the characterization on the nanomechanical properties, tai-
lored acoustic confinement will be important for precise phonon management by
structural engineering.
In this section, BLS is employed for the first measurement of the resonant
modes (the ‘music’) in sub-micron core–shell spheres [silica-poly(methyl meth-
acrylate)—SiO2–PMMA] having constant core radius and varying shell thickness
and in the corresponding spherical PMMA hollow nanoshells after dissolving the
silica core. We observed up to nine particle vibrational frequencies with increasing
size of the core–shell spheres and revealed the strong impact of the empty core of
the hollow capsules on the size dependence of the resonance frequencies. The
observed vibration eigenfrequencies are identified by detailed and thorough
numerical calculations as the resonance eigenmodes of the individual spheres. The
good overall agreement with the experiment allowed to determine the core density
and the two elastic constants of both constituents in the hybrid particles. We found
a significant deviation of these material properties in the nanostructured hybrid
spheres from their values in the macroscopic bulk systems, which underlines the
importance of such measurements at the nanoscale. The simulation of the
displacement fields of the different elastic modes allowed to visualize their
localization in different regions of the hybrid particles and provides a deeper
insight into their origin. These first findings illustrate qualitatively general
features of the localization of the elastic energy in nanostructured colloids beyond
core–shell particles.
The SiO2–PMMA particles with rigid silica cores and softer PMMA shells were
prepared by my coworker Dr. Spahn (DKI, Darmstad) in a two-step process
starting with the Stöber synthesis of the core followed by emulsion polymerization
78 4 The Vibrations of Individual Colloids
of the shell [56]. The silica core, with a diameter of 181 ± 3 nm, was coated with
PMMA spherical shells of three different thicknesses (in average 25, 57, and
112 nm), leading to core–shell particles of final (outer) diameter ranging from
d = 232–405 nm. All samples were characterized by scanning electron micros-
copy (SEM), using a 1530 Gemini SEM by LEO with acceleration voltages setup
between 0.2 and 1.0 kV. Figure 4.17a shows exemplary details for the uncoated
silica cores and the three core–shell particles. Although there is some degeneration
of the PMMA by the electron beam at 1 kV or some blurring using lower voltage
(200 V), the particles’ diameters d can be determined by averaging the software-
aided gauged diameters over about 100 spheres for each sample. The size poly-
dispersity is about 5%, which is confirmed by the formation of crystalline films
(Fig. 4.17a) after vertical lifting deposition from the particle suspensions (cf. Sect.
3.1 in Chap. 3) [57]. Figure 4.17b shows SEM images of the spherical hollow
capsules with d = 232 and 294 nm. After dissolution of the silica core with
aqueous hydrofluoric acid, the diameters are found to be unchanged within the
experimental error. For the smaller hollow capsules (bottom left) a considerable
fraction (approximately 50%) of the particles possess holes, while the 294 (top
right) and the 405 nm (not shown) hollow shells are nearly defect free due to their
thicker shells. Notably these defects in the 232 nm hollow capsules lead to a
broadening of the spectral lines, but do not significantly change their peak posi-
tions in the BLS spectra (see Fig. 4.20a below). The sizes obtained from the SEM
were confirmed by the hydrodynamic radii, Rh, of the core–shell particles in dilute
suspension measured by photon correlation spectroscopy (Sect. 3.3.1 in Chap. 3).
Within 3%, Rh amounts to 214 nm for bare silica and 262, 352 and 504 nm,
respectively, for the three core–shell particles. These values are expectedly higher
than the geometric radii R measured by SEM since for homogeneous spheres the
ratio R/Rh assumes the value of 0.78 [58]. Experimentally, R/Rh is found to be
slightly higher, varying between 0.80 and 0.85. For the BLS experiment, films of
all seven particles were prepared on a thin glass substrate using the vertical lifting
technique [4].
Fig. 4.17 SEM-images of a the bare silica (left, d = 181 nm) and silica–PMMA core–shell
particles (from right: d = 405, 294 and 232 nm), and b the hollow PMMA capsules (bottom left:
d = 232 nm, top right: d = 294 nm). For the smaller particles in b, holes in the spheres are
observed, whereas there are nearly no defects in the thicker shells [34]
4.3 Elastic Vibrations in Nanostructured Colloids 79
In analogy to the experiments in the previous sections, Fig. 4.18a shows the
q-independent BLS spectra of the bare silica particles and the three core–shell
particles. The spectra can be well represented by up to nine Lorentzian line shapes
as shown by the solid lines on the Stokes side of the BLS spectra. The peak
position of the spectral lines yields the resonance frequencies of the eigenmodes
indicated by solid circles in Fig. 4.18b–e for the four particles. Advantageously,
BLS can, in principle, record all thermally excited modes within one measurement,
which is not possible, e.g., in the pump-probe technique. For the bare silica
spheres, only two resonance frequencies at about 13 and 19 GHz can be resolved.
However, with increasing the PMMA-shell thickness of the core–shell particles
(d = 232, 294 and 405 nm), the BLS spectra become richer as it was observed for
pure polystyrene colloidal particles [8]. For the particle with the thinnest shell a
third weak peak is discernible in the spectrum in the inset of Fig. 4.18c, whereas
five modes are observed in medium-thickness shell (Fig. 4.18d), and even nine
modes are resolved in the BLS spectrum of the thickest shell (Fig. 4.18e).
The increased number of the resolved modes in the BLS with increasing particle
size relates to the intensity of the resonance signals which depends on qd. In the
present case of strong multiple scattering q B 2ki (the backscattering vector),
(d) (e)
(f)
Fig. 4.18 a BLS eigenmode spectra of bare silica (top) and the three core–shell particle films. The
representation of the BLS spectra with up to nine Lorentzians (solid lines) is shown on the Stokes-
side. (Note the different frequency scale for the spectrum in the top.) b–e Enlarged Stokes-sides of
the spectra from a (b pure silica, c–e core–shell). The experimental frequencies are accented by
orange spheres, the small vertical lines denote the corresponding calculated resonance frequencies,
each of them characterized by its angular momentum, l, shown at the top of the lines. All experi-
mental and theoretical (solid bars) values are summarized in f, with the dotted lines connecting
modes of the same angular momentum l; n = 1 for all observed modes [34]
80 4 The Vibrations of Individual Colloids
and the number of resolved modes increases with the 2kid as discussed in
Sect. 4.2.1 [25].
The elastic parameters (longitudinal and transverse velocities, cl and ct) of the
two constituents (core- and shell-materials) are not a priori known for such
nanostructured systems. An access to these material properties at these length scales
and high frequencies is important since they can considerably differ from their
values in macroscopic systems. The elastic constants are frequency dependent, and
BLS specifically yields their limiting high frequency values, which relate to local
packing and interactions, as well as, the glass transition temperature [51]. The
detection of more than two particle elastic excitations in the experimental (BLS)
spectra of Fig. 4.18 allows for an unambiguous determination of the elastic moduli,
shear modulus G ¼ .c2t and Young modulus E ¼ .c2l ð1 þ rÞð1 2rÞ=ð1 rÞ, with
r ¼ ðc2l 2c2t Þ=½2ðc2l c2t Þ being the Poisson ratio and . being the mass density.
The experimental values of the resonance frequencies are compared with the
resonance frequencies obtained from the calculated density of states (DOS) spectra
of a single constituent sphere of the experimental systems. The theoretical
computations were performed by Dr. Sainidou (University of le Havre) using a
formalism, appropriately developed for this case and presented elsewhere [28, 59].
Each resonance mode appearing at frequency f(n, l) in these DOS spectra is
characterized by the angular momentum l, imposed by the spherical symmetry of the
particle, where n denotes the n-th order solution for a given l. All the shell-localized
modes reported in this section have n = 1. The materials elastic parameters (cl, ct)
and densities are used as adjustable parameters in order to achieve the least deviation
between theoretical and experimental eigenfrequencies. Obviously, in the theoret-
ical calculations the constituent spheres are considered as homogeneous and
isotropic, and their elastic coefficients are frequency-independent.
First, the bare silica particles are considered (Fig. 4.18b). The two sound
velocities treated as adjustable parameters are obtained from representation of the
two experimental frequencies (solid circles) by the calculated resonance frequencies
(small vertical lines in Fig. 4.18b). The obtained sound velocities for the bare-silica
particles, cl = 4420 ms-1 and ct = 2780 ms-1, are significantly lower (&25%)
than the values of dense bulk amorphous silica (cl = 5970 ms-1, ct = 3760 ms-1,
. ¼ 2200 kg m3 ), indicating the presence of porosity in these particles [31]. It also
underlines the necessity of the BLS experiment to determine these values avoiding
erroneous assumptions. The mass density does not sensitively affect the DOS
spectra, due to the huge impedance difference between silica and air. Nevertheless,
these porous silica spheres should be less dense than bulk silica (see below).
Next, the PMMA coated silica particles consisting of same silica cores are
considered. For a first description of the DOS spectra, representing the experi-
mental frequencies seen in the BLS spectra of Fig. 4.18c–e (solid circles), the
elastic constants measured for the bare silica particles were used and the values for
the sound velocity and density of the PMMA shell were fixed on the values for
bulk PMMA (cl = 2800 ms-1, ct = 1400 ms-1, . ¼ 1190 kg m3 ) [17]. However,
this choice for the set of the elastic parameters of both the silica core and the
4.3 Elastic Vibrations in Nanostructured Colloids 81
PMMA does not quantitatively represent the experimental resonances in the BLS
spectra of Fig. 4.18c–e.
A systematic theoretical analysis based on the DOS spectra of the three core–
shell spheres has shown that changes in the elastic parameters of both materials of
the hybrid particles are required. Notably, it turned out that for the silica core one
must assume sound velocities, which are about 3% higher than in the bare silica
particle, i.e., cl,c = 4540 ms-1 and ct,c = 2860 ms-1. This hardening of the silica
core is probably due to a partial infiltration of methyl methacrylate in the pores and
subsequent polymerization to PMMA during the formation of the shell. Due to the
reduced impedance contrast between the silica core and the PMMA shell relative
to the bare silica spheres vs. air, the core mass density .c has now a substantial
influence on the DOS calculations. For a given density .c ¼ 1900 kg m3 , the
sound velocities in the PMMA shell are cl,s = 3080 ms-1 and ct,s = 1540 ms-1,
i.e., about 10% higher than in bulk PMMA; the PMMA mass density was kept at
the bulk value. Therewith, nearly all measured signals can be identified as
spherical eigenmodes with angular momentum l and can be captured quantitatively
by the theory within about 3%. The calculated resonance frequencies are shown by
small vertical lines in Fig. 4.18c–e and summarized in Fig. 4.18f along with the
corresponding experimental values. The Young modulus E and the shear modulus
G are directly accessible for both the core and the shell components.
In order to obtain an insight into the nature of the experimentally observed
modes (Fig. 4.18f), the elastic field at the resonance frequencies in the region of
the sphere was calculated, assuming a longitudinal acoustic plane wave of the
same frequency, impinging on the sphere (cf. Sect. 3.4.2 in Chap. 3). An example
of elastic-field intensity plots is given in Fig. 4.19 for the case of the thickest
PMMA shell sample (d = 405 nm). Their topology is that of a field-intensity
having 2l maxima on the internal interface of the outer circumference of the core–
shell particle: the two of them are strong (global) maxima along the direction of
the incident field, while the rest 2(l - 1) are weaker and distributed equidistantly
along the circumference. The vibration of the observed modes is radial at the
maxima changing alternatively direction (outwards or inwards) from maximum to
maximum. In the regions of minima the field is directed tangentially. Based on
Fig. 4.19, the experimentally observed modes (Fig. 2f) are shell-localized with
l = 2-4. Core-localized modes either do not exist within the considered fre-
quency range (films with d = 232 and 294 nm), or they are not observable in the
experimental BLS spectra (thickest-shell case, Fig. 4.18e), probably due to the
very strong localization (virtually delta-functions in the DOS spectra). It is worth
mentioning, however, that the elastic parameters affect the frequencies of both
core- and shell-like eigenmodes.
Figure 4.20a shows the BLS spectra of two hollow capsules (d = 232 and
294 nm), akin to those of double-shelled hollow carbon microspheres published
recently [60]. By comparison of the two spectra, it is apparent that the thicker
hollow capsules display significantly sharper peaks than the thinner hollow
capsules. For solid spheres, the line width of the spectra is associated with the size
82 4 The Vibrations of Individual Colloids
Low
(a) (b) (c)
High
Low
(d) (e) (f)
High
Fig. 4.19 a–c Elastic-field distribution for the first three resonances (n = 1) of the 405 nm core–
shell particle (Fig. 4.18e) in a cross section through the center of the sphere. Plane wave inci-
dence is assumed along the horizontal axis, from the left. d An example of core-localized mode
with l = 1 (n = 2) at 9.3 GHz, which is not observed experimentally. Plots e and f give,
respectively, the real part of the radial and polar component of the elastic field for the case shown
in c, with the arrows visualizing the vibrational mode of the shell [34]
polydispersity [8]. For the thin hollow spheres (d = 232 nm), however, an
additional source for the line broadening is the significant fraction of particles with
holes as seen in Fig. 4.17b. These defects are neither uniform in size nor in shape
and hence cause a distribution of the elastic properties that further broadens the
experimental spectra (Fig. 4.20a). In order to identify the nature of the experi-
mental eigenmodes in the two hollow capsules, the resonance frequencies of shell-
localized modes were computed, adopting for the PMMA shell the values of the
elastic moduli obtained from the representation of the eigenfrequencies of the
core–shell particles (Fig. 4.18f). Figure 4.20b shows the calculated (solid lines)
eigenfrequencies for the modes characterized by the angular momentum l = 2–4
for hollow PMMA nanoshells with constant inner diameter (181 nm) as a function
of the (outer) particle diameter d. The comparison with the experimental fre-
quencies (orange spheres) identifies the observed frequencies in the three hollow
PMMA spheres with the two lowest modes with l = 2 and l = 3. Interestingly, the
frequencies of these two modes vary very little with d due to two competing
effects. At a constant particle diameter, the resonance frequencies of hollow
particles increase with shell thickness [61], whereas the resonance frequencies of
filled spheres decrease with d since f ðn; lÞ 1=d [7–9]. Since in our case both shell
thickness and total diameter increase simultaneously, the net effect is essentially
the apparent insensitivity of the eigenfrequencies to the d variation of the branches
with l = 2 and 3 seen in Fig. 4.20b. Notably, the elastic parameters of the PMMA
were not affected by the core etching.
4.3 Elastic Vibrations in Nanostructured Colloids 83
(b)
In summary, this section reports the first study of localized vibrational exci-
tations in silica-PMMA core–shell particles and PMMA hollow capsules using the
powerful optical technique of BLS, which is applicable for turbid films. The BLS
spectra show up to nine eigenfrequencies of the core–shell particles, which sen-
sitively depend on the particle architecture and the mechanical moduli of the
constituent parts. The values of the Young moduli E and shear moduli G of the
constituent components are computed from the elastic parameters cl and ct and
densities, obtained for each component from the identification of the experimental
modes with the resonance modes appearing at f(n, l) for the n-th order of the l-th
harmonic in the calculated density of states (DOS) spectra. The anticipated
reduction of the moduli of the neat core compared to the bulk material due to its
porosity and subsequent subtle increase above the bulk values upon grafting with
PMMA chains are revealed. The anchoring and confinement of the PMMA
nanoscopic layer impact both types of moduli which were found to exceed the bulk
PMMA values. The observed eigenfrequencies in the hollow capsules exhibit a
peculiar but apparent insensitivity to the variation of the diameter as a result of
antagonistic trends inferred by the DOS calculations, while the PMMA elastic
84 4 The Vibrations of Individual Colloids
constants are not affected by removal of the silica core. In addition to the fun-
damental understanding of localized elastic modes in hybrid particles and the
concurrent determination of the mechanical moduli, the findings of this study will
contribute to a rational design of nanostructured colloids with strong resonances
that can be selectively excited [10].
4.4 Materials
PS Styrene ([99%, Aldrich) was washed three times with a 10% KOH solution
and three times with MilliQ water and then distilled under reduced pressure.
Acrylic acid (AA, 99%, Aldrich) was distilled without washing with KOH.
Sodium 4-vinylbenzenesulfonic acid (NaPSS, 90%, Aldrich) and potassium per-
sulfate (KPS, 99%, Acros) were used as received. By playing around with
monomer, comonomer (KPS, AA), and initiator concentration, M. Retsch could
prepare monodisperse PS spheres with diameter between 180 nm and 1 lm. All
reactions were carried out with 0.1–0.2 g KPS, 0.005–0.06 g NaPSS, and 0–0.3 ml
AA in 250 ml H2O at 70 or 80 C.
PMMA Methyl methacrylate (MMA, 99%, Acros) was washed three times with a
10% KOH solution and three times with MilliQ water and then distilled under
reduced pressure. Sodium 4-vinylbenzenesulfonic acid (NaPSS, 90%, Aldrich),
potassium persulfate (KPS, 99%, Acros), and 2,20 -azobis(2-methyl propionami-
dine) dihydrochloride (ABA, 97%, Aldrich) were used as received.
Cationic initiator (ABA) as well as an anionic system (KPS, NaPSS) have
been used. The particle diameter is found to depend mainly on the initial MMA
concentration.
4.4 Materials 85
The SiO2–PMMA particles with rigid silica cores and softer PMMA shells were
prepared by my coworker Spahn at DKI Darmstadt in a two-step process starting
with the Stöber synthesis of the core followed by emulsion polymerization of the
shell [56]. The silica core, with a diameter of 181 ± 3 nm, was coated with
PMMA spherical shells of three different thicknesses (in average 25, 57, and
112 nm), leading to core–shell particles of final (outer) diameter ranging from
d = 232 to 405 nm.
References
5.1 Introduction
The first papers on photonic effects by Yablonovitch [1] and John [2] in 1987
stimulated over the years much theoretical and experimental work on the
propagation of electromagnetic waves through appropriately structured materials
and subsequently led to the birth of the new research field ‘photonic crystals’ [3].
The tremendous interest in photonic crystals with specially designed periodic
variations in dielectric constant largely originates from their display of propagation
band gaps for light. The appearance of band gaps makes an advanced control over
light propagation possible and permits as well a series of novel optical phenomena
such as slowing and localization of light or negative refraction [3]. Soon after the
discovery of photonic crystals, it was found that in analogy to the electromagnetic
waves, band gaps also exist for the propagation of acoustic waves, and the so-called
phononic crystals [4–10] are the elastic analogue of photonic crystals replacing the
role of the dielectric constant by the elastic parameters and density. Such an analogy
exists as a consequence of the common origin of the band gaps in both cases, i.e., the
destructive interference of Bragg diffracted waves in periodic structures [11, 12],
and hence these gaps are also termed as Bragg gaps. However, the different nature of
electromagnetic and acoustic waves also guarantees the existence of some important
differences between photonic and phononic phenomena. Unlike electromagnetic
radiation that is characterized as transverse waves, acoustic waves in general are full
vector waves with both longitudinal and transverse polarizations, and their
propagation depends additionally on the material density. Even for a homogeneous
and isotropic medium, the acoustic wave propagation is governed by three
parameters, the two Lamé coefficients and density, in contrast to the single
parameter, the dielectric constant, that determines the propagation of light.
Evidently, the phononic phenomena are anticipated to be more complex and rich.
For phononic crystals, the band diagram depends on several parameters such as
the elastic constants and density of the component materials, symmetry of the
lattice, shape of the inclusions, and the filling fraction. The width of the band gap
generally increases with the contrast between the densities and sound phase
velocities of the component materials, and the center of the gap can be tuned by
changing the lattice parameter [9, 10]. The search for phononic structures started
with two theoretical works in 1993, which predicted the existence of phononic
band gaps in periodic two-dimensional (2D) elastic composites of parallel cylin-
ders embedded in a host matrix [9, 10]. The experimental verification of phononic
band gaps followed few years later, realized in metallic macrostructures with gaps
at sonic or ultrasonic frequencies [4, 5, 8, 13]. Further explorations also revealed a
number of peculiar phenomena with potential applications associated with acoustic
wave propagation including tunneling effect [14], negative refraction and focusing
[15], double refraction, etc. [16].
Theoretical calculation of phononic band diagram requires no specification of
the lattice constant of the structure or the corresponding wave frequencies as long
as the crystal is defined by the same set of frequency-independent elastic
parameters. In other words, a fundamental length scale does not exist for phononic
phenomena, which is a direct consequence of the invariance of the wave equation
of elasticity under the simultaneous transformation of space coordinates and
frequency. However, acoustic waves of different frequencies do bear distinct
characteristics, particularly when their applications are concerned, and therefore in
practice the frequency range of the waves of interest constitutes an important
consideration. Recently, growing attention has been paid to hypersonic (GHz)
phononic crystals and the first experimental observations of the hypersonic Bragg
gaps have been lately reported in three-dimensional (3D) colloidal crystals made
by self-assembly [17] and in 2D polymeric porous structures with hexagonal
symmetry created by laser interference lithography [18]. Hypersonic waves, owing
to their high frequencies, display certain unique features that are not possessed by
ordinary acoustic waves such as being thermally excited, acting as the main heat
carrier in dielectrics, and interacting with electrons and photons in a rich manner.
Consequently, hypersonic crystals with the potential to mold the flow of
hypersound may be utilized to achieve high-level control over many important
physical processes involving heat transport and complex phonon–photon or
phonon–electron couplings. A detailed understanding of phonon propagation in
hypersonic crystals thus becomes important.
The fabrication of hypersonic crystals, compared with their sonic and ultrasonic
counterparts, is much more demanding as the dimension typifying the structure has
to be scaled from macroscopic down to sub-micron scale. On the other hand,
advance in relevant nanofabrication, partially driven by the desire of creating
various photonic structures of similar dimensions, has offered some available
means to achieve such a purpose including, for example, holographic interference
lithography [19], direct laser-writing [20], two-photon polymerization [21], or
self-assembly [22, 23]. Colloidal superstructures, self-assembled from colloidal
particles, represent a promising material class for phononic applications. The
maturation of colloidal science enables the preparation of colloidal particles with
well-defined size and shape for a great many of materials ranging from organic to
inorganic, thus providing abundant building blocks with varied elastic properties.
5.1 Introduction 91
The progress in colloidal particle self-assembly allows easy and cheap fabrication
of large area high quality single crystalline 2D and 3D crystals as well as
non-crystalline structures. The appearance of binary [24] and ternary [25] crystals
further enriches the available structure types. By filling the interstitials between the
colloidal particles with different materials, additional freedom in elastic parame-
ters of the system is provided. Moreover, in most cases the colloidal particles
possess a spherical shape due to surface tension effects, thus represent a strong
scattering unit when the elastic contrast between the particle and surrounding
becomes large, which may cause additional gap formation with origin different
from Bragg diffraction as will be encountered later [26].
The experimental exploration of hypersonic crystals, however, faces additional
challenge in monitoring the phonons in such small structures. Apparently, the
commonly used sonic and ultrasonic transmission techniques for macroscopic
sized structures cease to work. It has been demonstrated that Brillouin light
scattering (BLS), which takes advantage of the inelastic scattering of photons by
thermally excited high frequency phonons, represents a powerful tool to record the
phonon dispersion relation in hypersonic crystals.
This chapter deals with the phononic band diagrams of transparent colloidal
systems that can be measured by BLS. A short introduction into the nature of the
bands in the dispersion relation is given as well as into the properties of the
effective medium. Experimental results are discussed treating the effective
medium velocity of infiltrated defect doped colloidal crystals and the nature of
acoustic band gaps in phononic systems, including the first realization of a
hybridization gap.
(a) (b)
Fig. 5.1 a From (Ref. [27]) calculated acoustic band diagram for lead spheres in a beryllium
matrix forming an fcc lattice with an occupancy of 8.32 vol% the lowest (transverse) branch is
degenerated. The three flat branches around 2.8 xa/c are degenerated, too. There are two
complete band gaps. b Dispersion of the longitudinal waves in a 1D simple crystal (black line)
over several Brillouin zones. The dashed line denotes the dispersionless ‘free particle’ case
with massless elastic springs with spring constant D and equilibrium distance a as
shown in Fig. 5.2 [28, 29]. Let us assume a longitudinal wave, which creates
displacement of the spheres in the direction of the chain. When assuming (in
reasonable approximation) that the elastic forces that determine a vibration have
their origin in the relative displacement only of neighboring particles, symbolized
by the compressed and stretched springs in Fig. 5.2 leading to forces (black
arrows) in the same direction, while the inertia originates from the absolute
displacement of the spheres in space. With u(na) being the absolute displacement
of the nth sphere, the harmonic potential between all neighboring spheres becomes
1 X
U harm ¼ D ðuðnaÞ uð½n þ 1aÞÞ2 : ð5:1Þ
2 n
Since elastic forces and inertia must compensate each other, the equation of
motion becomes
o2 uðnaÞ oU harm
m 2
¼ ¼ D½2uðnaÞ uð½n 1aÞ uð½n þ 1aÞ: ð5:2Þ
ot ouðnaÞ
Assuming a plane longitudinal wave of the form
component supports shear waves and the other component does not, one must not
ignore the conversion between waves of different nature into each other that is
described mathematically by the T-matrix (Eq. 3.28).
Regions in the band diagram where no band exists for all ks are called phononic
band gaps. These are frequency regions where no mechanic wave can propagate.
Since the existence of such gaps is one of the striking features of the band diagram
and promises also some technical importance, the search for and investigation or
tailoring of phononic band gaps is a demanding question in the young field of
phononics.
In the following parts of the introduction, the properties of the effective
medium as well as a short overview over the phononic band gaps is given. The rest
of this chapter is concerned with experimental studies on colloidal phononic
systems, including the investigation of the effective medium velocities in defect
doped opals, the first experimental realization of a hybridization gap and the band
diagram of nanostructured systems.
The slope at the long wavelength limit in the band diagram gives the longitudinal
and transverse sound velocities, cl and ct, respectively. While these velocities are
well known for the most pure materials, their prediction for composite materials of
varying composition and geometric structure is not straight forward and the topic
of the effective medium theory (EMT).
A very general approach to predict the effective longitudinal velocity of a
two components system is given by Wood’s law as the harmonic mean of the
moduli [30]:
1 / ð1 /1 Þ
¼ 1þ ; ð5:8Þ
Meff M1 M2
where M ¼ c2l . is the bulk longitudinal modulus of each component. In principle
this approach can be generalized for n components to
1 / ð/ Þ ð/ Þ
¼ 1 þ 2 þ þ n ð5:9Þ
Meff M1 M2 Mn
P
with n /n ¼ 1: For equal mass densities ð.1 ¼ .2 Þ; the effective longitudinal
velocity would be given as the harmonic mean of the squared sound velocities of
the components (cl ½h1=c2l i1=2 ). For transverse sound velocities, it should
be sufficient to replace the bulk longitudinal moduli in Eq. 5.8 by the share
moduli.
Anyhow, the simple assumption of Wood’s law ignores for example the
interaction of longitudinal and transverse waves between different components as
it (in the longitudinal version) compares only the bulk moduli. However, in certain
5.1 Introduction 95
cases, these interactions may lead to conversion of energy between bulk and shear
waves (T-matrix) and must not be ignored anymore. In the theoretical calculations
mentioned above [27], the authors compare the calculated effective velocities for
several different cases with the arithmetic mean and the harmonic mean of the
sound velocities as well as with the square root of the harmonic mean of the
squared velocities of the components—all weighted with the filling fraction, but
not with the mass densities. This deviation from Wood’s law does not lead to a
reproducible result for all cases, since in some cases the one mean is nearer to the
calculated result while in other cases another mean is nearer to the theoretical
findings. However, even applying Wood’s law, which is not done in the paper,
does not deliver perfect agreements for all cases, although a relatively large
deviation is achieved only in the case of lead inclusions in a beryllium matrix with
a filling fraction of 8.23%. Here, using the parameters given by the authors,
Wood’s law delivers an effective longitudinal sound velocity of 9,250 ms-1, while
the calculated velocities are about 10% higher, between 10,226 and 10,619 ms-1,
depending on the crystallographic direction.
For the special case of spherical inclusions (and the practical application for
approximately spherical inclusions), a more sophisticated but still easy-to-apply
EMT method was developed mainly by Gaunaurd and Wertman, reviewed by the
same in Ref. [31]. In their calculations they take the ratios of longitudinal and
transverse sound velocities for both components into account as well as the ratio of
the mass densities. Furthermore, the effective density is calculated differently for the
two general subcases of matrices that can only support insignificant amounts of shear
(‘fluid matrices’) and ‘elastic matrices’ supporting non-negligible amounts of shear.
When calculating the effective longitudinal sound velocities for the different
cases in Ref. [27] following this approach (utilizing a small C++ program written
by myself), reasonable results are obtained for all subcases (Au/Si, Pb/Si, Pb/Be,
and Au/SiO2), although for some subcases from the theoretical band diagram there
are different effective velocities in different crystallographic directions, which are
described unequally well by this EMT. E.g., for the Pb/Be subcase mentioned
above, the program gives an effective sound velocity of 10,820 ms-1, which is
only about 2% more than predicted for the C–L direction.
In Table 5.1 three general cases of phononic systems with filling fraction / are
compared for Wood’s law and the EMT of Gaunaurd and Wertman: polystyrene
spheres in silicon oil (solid/liquid), alumina spheres in PS (solid/solid), and air
bubbles in silicon oil (gas/liquid).1 For all subcases the values following the
Gaunaurd and Wertman approach are given for fluid (GWf) and elastic matrix
(GWe), although principally the latter should be meaningful only in the alumina/
PS subcase since silicon oil that is matrix in the other two subcases does not
support shear waves.
1
Material parameters: PS: ct: 1,200 ms-1, cl: 2,350 ms-1, .: 1,050 kg m-3; silicon oil: ct: 0 ms-1,
cl: 1,400 ms-1, .: 1,000 kg m-3; Al2O3: ct: 6,345 ms-1, cl: 10,850 ms-1, .: 3,970 kg m-3; air:
ct: 0 ms-1, cl: 343 ms-1, .: 1.2 kg m-3.
96 5 Phononic Behavior of Colloidal Systems
Table 5.1 Comparison between calculated effective longitudinal sound velocities (in ms-1)
using the EMT of Gaunaurd and Wertman (GW) and Wood, respectively, for three cases and five
filling fractions /
/ PS in silicon oil Al2O3 in PS Air in silicon oil
For the first case of PS in silicon oil, it does not make any significant difference
which kind of matrix is taken into account. Both calculations come to practically
the same results, leading to sound velocities lower than those predicted by Wood’s
law. With increasing filling fraction the difference increases up to 12% for the
dense packing case / = 0.74.
In the subcase of a harder solid embedded in a softer one (alumina in PS), GWe
and Wood’s law deliver quite similar results, differing by less than 3% at
/ = 0.74, while the theoretically meaningless GWf delivers significantly higher
values.
The last subcase of air bubbles in silicon oil shows that in this case, where both
components’ transverse sound velocities are set to zero, GWe delivers exactly the
same results as Wood’s law, while the GWf calculations give higher values.
However, even for very low filling fractions all theories give extremely low values.
This means that even a low amount of air bubbles should slow down the effective
sound velocity in a liquid.
Another EMT method is described by Waterman and Truell [32]. Their
approach is in principle a full multiple scattering theory including numerical
calculation with n equations for n scatterers. For every scatterer the exact form of
the multipole coefficient must be known, and so this method is too complicated to
be used for simple calculations of the effective sound velocity. Taking the
effective sound velocity from the theoretical band diagrams calculated by the
multiple scattering method touched on in Sect. 3 in Chap. 3 is very akin to this
approach.
Fig. 5.3 Infiltration of a turbid colloidal crystal with an index matching liquid resulting in a
transparent ‘wet’ crystal (according to [17])
Fig. 5.4 The dispersion relations of a colloidal PS (d = 307 nm) opal infiltrated with silicon oil
(left see also Ref. [17]) and with PDMS (right, see also Fig. 5.11 and Ref. [26]) in the C–M
direction of the fcc crystals are compared. In both spectra there is a clear BG (diagonal pattern).
The dashed lines indicate the C–M distance in the reciprocal space, hence the edge of the first BZ.
The dotted lines denote the frequencies that correspond to qBZ (cf. Sect. 5.1.3). For the opal
infiltrated with PDMS a HG (vertical pattern) appears, too [35]
5.1.3.1 Normalization
In general, the condition that has to be fulfilled to create a Bragg gap is that the
distance between two scattering planes in the direction of the incident wave
matches the wave length of the wave, or, in reciprocal space, that the wave vector
q is half the diameter of the first Brillouin zone in a certain crystallographic
direction. By the method of vertical lifting deposition (Sect. 3.1 in Chap. 3) the
colloidal fcc crystals show a preferred orientation relative to the substrate. When
inserting the sample in the same orientation as prepared into the BLS setup and
utilizing the special transmission geometry, where the wave vector is parallel to
the substrate, the probed direction in the reciprocal lattice is usually the C–M
direction. The first BZ of an fcc lattice is shown in Fig. 5.5. In the C–M direction,
the distance in reciprocal space between C and M is
33=2
p
jCMj ¼ 2pffiffiffi ð5:12Þ
2d
with sphere diameter d. Therefore, the wave vector qBZ in this crystallographic
direction must have the same value to fulfill the Bragg conditions, because |C–M|
is half the diameter of the Brillouin zone DBZ, i.e.,
33=2
DBZ p
qBZ ¼ ¼ 2pffiffiffi : ð5:13Þ
2 2d
If one assumes pure acoustic behavior with an effective sound velocity ceff the
frequency at qBZ would be
33=2
ceff qBZ ceff DBZ ceff
fBZ ¼ ¼ ¼ 2 pffiffiffi ð5:14Þ
2p 4p 2 2d
with f = x/2p = cq/(2p). Now one can normalize the measured values for the
frequency and the wave vector to the values at the edge of the Brillouin zone:
pffiffiffi
f 2f 2d fd
fnorm ¼ ¼ 3=2 1:54 ð5:15Þ
fBZ 3
ceff c eff
2
pffiffiffi
q q 2d
qnorm ¼ ¼ 3=2 0:245qd ð5:16Þ
qBZ 3
p
2
Brillouin light scattering can be used to measure the effective sound velocity in
homogeneous materials as well as in structured materials if the structure size is
smaller than the traveling phonon, or in other words, if 2p/q [ d with structure
spacing d. E.g., for q = 0.01 nm-1 (that corresponds to an angle of 50 in
transmission geometry for the setup described in Sect. 2.3.2 in Chap. 2) the
structure must be clearly smaller than 600 nm.
While in bulk, not too thin films of a single polymer of course the polymer’s
sound velocities (if it support shear waves) are found, in films of thin multilayers
of different components, e.g., for two different polymers, the individual sound
velocities of both components are found when measuring parallel to the interfaces
[41]. If the wave vector is chosen perpendicular to the layers, at least for thin
(\30 nm) layers, only one signal corresponding to the effective sound velocity is
found. Following the effective medium theories, ceff lays in between the velocities
of the individual components [42].
Effective sound velocities are also found in nano- and mesoscopic 3D structures
like the infiltrated colloidal crystals presented in Sect. 5.1.3. In Fig. 5.4 ceff is shown
for PS colloidal crystals (/ = 0.74) infiltrated with silicon oil (left) or PDMS (right).
The effective sound velocities are 1,990 and 1,670 ms-1, respectively. These are
between those of PS (cl = 2,350 m/s, . ¼ 1; 050 kg m-3) and PDMS (cl =
1,050 m/s, . ¼ 965 kg m-3) or silicon oil (cl = 1,400 m/s, . ¼ 1; 000 kg m-3)
[26]. The EMT of Gaunaurd and Wertman [31] delivers effective velocities of 1,714
and 1,488 m/s, for silicon oil and PDMS as infiltration liquid, respectively, using the
numbers given above. Wood’s law, on the other hand, delivers 1,925 and 1,617 m/s,
respectively, i.e., both EMTs seem to underestimate the real effective sound
velocities, however for Wood’s simple law, the error is significantly smaller.
A more systematic study of the influence of the filling fraction on the effective
sound velocity in PS colloidal crystals infiltrated with PDMS and the comparison
with EMT was performed within the framework of this thesis. Therefore, a set of
colloidal crystals consisting of PS colloids (d = 260 nm) and a certain amount of
silica particles with practically the same diameter (d = 255 nm) were synthesized
5.2 Effective Medium Velocity in Defect Doped Opals 101
(i) (j)
Fig. 5.6 SEM pictures: a–d PS:SiO2 colloidal crystals with composition 2:1, 4:1, 8:1, and 12:1;
e–h The corresponding defect doped crystals after etching away the silica beads. Analytics:
i TGA measurement on the composite PS:SiO2 colloidal crystals (decomposition of PS). The
inset shows the loss of water (legend like in j); j UV/VIS spectra of the PS:SiO2 colloidal crystals
shown in a–d. For comparison also the spectra of the pure PS colloids, the pure silica colloids and
of two mixtures with higher amount of silica are shown (according to [43])
102 5 Phononic Behavior of Colloidal Systems
measurements (Fig. 5.6j) were performed and the results were compared with the
counting of the SEM pictures. From the TGA measurements it is easy to calculate
the PS:silica ratio at high temperatures the PS is decomposed and only the silica
spheres remain. By comparing the relative mass loss and taking into account the
different mass densities and the exact radius of the spheres the relative amount of
the polymer is obtained [43]. It should be noted that it is meaningful to compare
the change for the dry samples i.e., to set the starting point of the comparison at
T [ 100 °C (inset in Fig. 5.6i).
The counting of the SEM pictures shows that systematically less silica particles
are built in the crystals than should be expected from the initial dispersion
concentrations. However, the SEM pictures show only the distribution on the
imaged surface and, by what reason ever, the ratio inside the opal could be different.
Anyhow, the comparison between the data obtained by the TGA measurements and
the SEM counting lead to a systematic and consistent result [43]. The amount of
silica is always given by approximately 0.8 times the expected fraction, i.e., the real
ratio of the ‘12:1’ PS:SiO2 sample is &12:0.8, etc., however, for the rest of this
section the names originating from the initial ratio are kept to avoid confusion.
The so found real compositions are verified by the UV/VIS measurements
shown in Fig. 5.6j. From the graph that contains also two additional samples with
initial ratio 1:1 and 0.3:1 it can be seen that there is a clear blue shift of the Bragg
peak for increasing amount of silica, resulting from the change in the effective
refractive index (cf. Sect. 3.3.6 in Chap. 3). The utilization of the modified Bragg
equation (Eq. 3.10)
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
kBragg; theo ¼ 1:63d 0:74vPS n2PS þ 0:74vSiO2 n2SiO2 þ 0:26n2air ð5:17Þ
Table 5.2. First, the velocities are compared to the results of Wood’s simple law,
using the standard elastic parameters for PS and PDMS (given under the table).
Interestingly, cWood seems to capture the experimental values quite well within a
non-systematic error of 1.5% or less, which is in the range of the error of /. Using
the same parameters in the algorithm of Gaunaurd and Wertman (cGW,s) leads to a
significant deviation from the experimental values between 6% and more than 9%,
while the deviation increases systematically for increasing /.
Anyhow, it is possible to bring the results of their EMT in better accordance with
the experimental results when taking cl(PS), ct(PS), and cl(PDMS) as fit parameters
in a mean square fit for all five filling fractions. So the cGW,f are obtained, keeping
the parameters in relatively reasonable ranges. The PDMS is assumed to be about
8% softer (1,109 ms-1 instead of 1,200 ms-1), for the PS the transverse sound
velocity must be chosen smaller (1,100 ms-1 instead of 1,200 ms-1) and the
longitudinal one must be chosen higher (2,470 ms-1 instead of 2,350 ms-1). Doing
so, the error stays within about 1%, however changing systematically from a little to
high values for low / to a little to low values for the undoped crystal. Even better
results can be achieved using really free fit parameters without any fitting ranges.
Table 5.3 Comparison between experimental effective longitudinal sound velocities and results
from Wood’s law and the EMT of Gaunaurd and Wertman using standard and fitted parameters
(in ms-1)
PS:SiO2 / cexp cWood D/% cGW,s D/% cGW,f D/%
2:1 0.529 1,382 1,362 -1.5 1,304 -6.0 1,395 1.0
4:1 0.617 1,464 1,451 0.9 1,372 -6.7 1,473 0.6
8:1 0.673 1,522 1,520 0.1 1,421 -7.1 1,530 0.5
12:1 0.694 1,567 1,548 -1.2 1,441 -8.7 1,553 -0.9
1:0 0.740 1,627 1,617 -0.6 1,488 -9.3 1,609 -1.1
Constant: .(PS) = 1,050 kg m-3, .(PDMS) = 965 kg m-3, ct(PDMS) = 0 ms-1; cWood, cGW,s:
cl(PS) = 2,350 ms-1, ct(PS) = 1,200 ms-1, cl(PDMS) = 1,050 ms-1; cGW,f: cl(PS) = 2,470 ms-1,
ct(PS) = 1,100 ms-1, cl(PDMS) = 1,109 ms-1
However, the best fit is achieved with unreasonable parameters, especially for PS
(cl(PDMS) = 1,069 ms-1, cl(PS) = 2,622 ms-1, and ct(PS) = 1,082 ms-1).
In summary, it can be said that for different filling fractions the investigated
colloidal system shows the expected trend in the effective medium sound velocity.
The increase in ceff with increasing amount of the harder component was predicted
qualitatively by Wood’s law as well as by the more elaborated EMT of Gaunaurd
and Wertman, developed for the special present case of spheres embedded in an
elastic or fluid matrix. However, while the predictions utilizing Wood’s law agree
also quantitatively well with the experimental results, there is a systematic deviation
to the velocities calculated by the other EMT. The reason of the deviation is unclear.
It might originate in a kind of size or frequency effect that leads to deviations for
small spheres (possibly originating from elastic confinement inside the spheres) or
for high frequencies. Anyhow, these assumptions are somehow speculative.
Structured materials with a periodic modulation in the density and elastic coeffi-
cients, so-called phononic crystals [11], can exhibit phonon band gaps at Bragg
frequencies or wavelengths commensurate to their lattice constant. In addition to
Bragg gaps (BG), theory predicts gaps evoked by resonance modes of the
constituent components interacting with the extended acoustic branch of the
composite structure [45]. These gaps prevent elastic waves with certain frequen-
cies to propagate through the crystal at least in certain crystallographic directions.
The width and the position of the BG in general depends on the contrast between
the densities (.), longitudinal and transverse sound velocities (cl and ct) of the
component materials, and on the lattice parameter [9, 10, 33]. Yet, structures with
strong localized resonant elements can shift the gap well below the Bragg
frequency associated to the lattice constant [8, 46]. Soon after the first
5.3 Band Gaps in Polymer Opals and Disordered Systems 105
with a Doppler frequency shift x = ±ck, where c is the speed of sound with
longitudinal (or transverse) polarization in the effective medium (e.g. lowest
spectrum in Fig. 5.10, cf. Sect. 2.3 in Chap. 2).
The precursor dry PS and PMMA opals have a highly ordered crystalline
morphology as indicated by the SEM image in Fig. 5.9. The single crystalline order
extends over a few hundred micrometers. The dry films exhibit strong multiple light
scattering due to the large optical contrast between the particles and the surrounding
air and due to the particles elastic form factor. The wave vector q is, therefore,
ill-defined which precludes the measurement of the dispersion relation x(q).
Nonetheless, the BLS spectrum of the dry opals reveals several localized
(q-independent) modes [54, 55]. As discussed in Chap. 4, these are identified as
Fig. 5.9 SEM top-view images of the PS-307 crystal (left) and the PS-hybrid (right)—both
‘‘dry’’ before infiltration (scale bar is 1 lm). The insets display the Fourier-transform images
computed from the SEM pictures over an area of about 4 by 4 lm [26]
5.3 Band Gaps in Polymer Opals and Disordered Systems 107
Table 5.4 Sound velocities, particle sizes and lowest eigenfrequencies in colloid-based
phononics
Material ct/ms-1 cl/ms-1 d/nm f(1, 2)/GHz c/ms-1
PS 1,200 ± 30 2,350 ± 50 307 3.3a
360 2.9a
PDMS 0 1,050 ± 20a
PS/PDMS 1,670 ± 30a 307 2.2b 1,490b
1,570 ± 30a 360 1.9b 1,490b
Hybrid/PDMS 1,510 ± 20a 300/360 1,400b, c
a
PMMA 1400 ± 40 2,800 ± 50 327 4.0
PMMA/PDMS 1,720 ± 30a 327 2.3b 1,560b
Densities . (kg/m3): PS: 1050, PMMA: 1,190, PDMS: 965
a
Measured by BLS
b
Computed for the wet opals
c
For 65% filling fraction of spheres
vibration eigenmodes of the particles and each mode can be specified by a pair of
indices (n,l) defining the lth spherical harmonic of the nth radial mode. The
frequencies of these sphere eigenmodes depend on the size of the particle, the mass
density and the speed of sound in air and in the particle for both polarizations (i.e.
compression and shear moduli). Consequently, the mechanical properties of the
samples can be reliably determined from the theoretical fit of the experimental
eigenfrequencies. The obtained longitudinal and transverse sound velocities in the
two types of particle materials along with the lowest f(1, 2) eigenfrequency are
given in Table 5.4. For the particles embedded in the infiltrated PDMS, the eigen-
frequencies can be theoretically computed using the elastic parameters of Table 5.4
[51, 56]. Expectedly, f(1, 2) decreases when the spheres are embedded in PDMS.
The infiltration of the thin dry opals by PDMS with a refractive index (n = 1.45)
close to that of the PS particles (n = 1.59) diminishes multiple light scattering and
hence q is well defined. To obtain the desired dispersion relation x(q), the BLS
spectra of the wet opals were recorded along the high-symmetry directions of the
reciprocal space. In the present case of the fcc lattices, the first Brillouin zone (BZ) is
a truncated octahedron (Fig. 5.5). The experimental q is confined in the hexagon
formed by the intersection of the (111) plane with the BZ [17]. The direction of q is
selected along C–M, where M is the edge center of the hexagon and the evolution of
the BLS spectra with q near the BZ for the PS wet opal is shown in Fig. 5.10. The
simple picture of single phonon propagation in the effective medium at low q values
becomes complex as q increases towards the BZ boundary. Up to four Lorentzian
curves are required to represent the experimental BLS spectrum. In contrast to the
PS/silicon oil opal with a single peak splitting across the BZ boundary [17], the same
opal infiltrated in PDMS providing higher elastic contrast displays richer spectral
features and exhibits a second splitting at lower q values within the first BZ. The
experimental dispersion relation is depicted in Fig. 5.11. In the hypersonic PS/
PDMS crystal (d = 307 nm) only one longitudinal phonon branch is observed at
low q values with cl = 1,670 ms-1, intermediate between the longitudinal sound
108 5 Phononic Behavior of Colloidal Systems
velocities in the pure component materials. Transverse phonons are not observed
experimentally, probably PDMS cannot support shear waves and the mechanical
contact between the particles is weakened. The most striking feature of the
dispersion diagram is the simultaneous presence of two band gaps at about 3 and
4 GHz, respectively. The latter is clearly a BG since it occurs at the edge of the BZ,
q ¼ qBZ 0:0133 nm1 ; that matches the distance C–M, i.e. (3/2)3/2p/a, where
pffiffiffi
a ¼ 2d is the lattice constant of the given fcc crystal. An analogue behavior is
observed in the second wet opal of the larger PS spheres (d = 360 nm) for which the
gap positions shift to lower wave vectors and frequencies (open circles in the upper
left diagram of Fig. 5.11).
In order to elucidate further the nature of the two gaps, the effect of the
crystalline order on the experimental band diagram was examined. The formation
of a hybrid colloidal film consisting of a mixture of an equal number of two PS
spheres (d = 300 nm and d = 360 nm) artificially broadens the size distribution
beyond the polydispersity limit of about 5% necessary for crystallization [44].
Indeed, the crystallization is prohibited in this hybrid colloidal film, as indicated
by the lack of a long-range order in the right-hand panel of Fig. 5.8, leading to an
amorphous colloidal glass. The BLS spectra of the infiltrated monodisperse opal
and hybrid films are shown in Fig. 5.12 for two wave vectors. The deletion of one
peak in the spectrum of the hybrid is due to the disappearance of the BG in the
disordered hybrid, as it is clearly shown in the dispersion plot in the upper left
diagram of Fig. 5.11 (solid squares). Apparently, the crystalline order is a
prerequisite for the BG but not for the newly observed lower frequency HG, which
is omnidirectional in the colloidal glass. An identification of the latter as a
theoretically anticipated HG [45], through density-of-states (DOS) calculations
[56], is examined in Fig. 5.11. The right-hand panel of Fig. 5.11 displays several
eigenmodes of the individual PS spheres embedded in the fluid PDMS. The lowest
f(1, 2) appears to compare well with the frequency at the crossing with the acoustic
branch and the opening of the HG. Moreover, the lowest frequency points at the
experimental dispersion at high qs compare very well with the computed f(1, 2).
This and the two higher flat bands of localized modes in Fig. 5.11 cause
sufficiently strong inelastic light scattering at high qs [51] and compare well with
the particle resonances of higher l. Their presence in the BLS spectra obscures the
resolution of the phonons in the second BZ. Interestingly, the bands originating
from particle resonant modes appear to be considerably narrower and occur at
higher frequency than theoretically predicted [45]. This can be ascribed to viscous
losses in the liquid matrix that were not taken into account in the theoretical
calculations, and which weaken the interparticle interactions as a result of the
reduced overlap between the corresponding wave fields. Qualitatively, the opening
of the two gaps is also observed in a third wet opal of PMMA with d = 327 nm
in PDMS (Fig. 5.13). The HG occurs at 2.5 GHz, very close to the f(1, 2) =
Fig. 5.13 Phononic band diagrams for PMMA wet opals in PDMS with d = 327 nm, along the
C–M direction. The change in the DOS induced by the corresponding single PS particle in PDMS
is shown in the right-hand panel [26]
2.35 GHz of this particle in PDMS, whereas the wet opal exhibits a sound velocity
of 1,720 ± 30 ms-1 in the long-wavelength limit (Table 5.4).
It is worth noting that the various effective-medium theories (EMT) yield sound
velocities about 10% smaller than the experiment [31, 37], even if the viscosity
in the liquid matrix is taken into account. On the other hand, if one considers the
colloidal film as a polymer (solid) matrix with fluid inclusions [37], EMT strongly
overestimates the sound velocity. The above, in view also of the fact that measured
sound velocities agree generally well with the results of EMT in non-close-packed
colloidal crystals [58], suggests the existence of consolidation, at least to some
degree that may be different for different samples, in the colloidal films [59].
This also explains why the measured effective sound velocity is different in the
two PS/PDMS crystals (see Table 5.4).
In conclusion, this work presents, for the first time, the discovery of two
phononic band gaps of different nature coexisting at hypersonic frequencies in the
same physical system and elucidated the underlying physical mechanisms. Induced
disorder did not destroy the newly demonstrated HG. This study has been possible
by taking advantage of the opportunities offered by the colloidal science to tailor
the phononic band diagram of nanostructured materials, measured directly by
BLS. Manipulating the flow of phonons may allow heat management, e.g. in
thermoelectrics. Finally, it was pointed out the need of a detailed quantitative
evaluation of the dispersion diagrams of colloid-based phononic structures, by
means of full elastodynamic calculations that take into account consolidation and
soft matter properties (e.g. structural dynamics of the component materials or
interfacial effects), which still remains an open challenging theoretical problem.
5.3 Band Gaps in Polymer Opals and Disordered Systems 111
In Sect. 5.3.1 the realization of a hybridization gap (HG) was presented for
monodisperse colloidal crystals as well as for disordered ‘hybrids’, e.g., for the
1:1 number ratio 300:360 nm PS hybrid in Fig. 5.11. However, when discussing
the HG at that point, the mixing of two different sizes of colloids was only
mentioned in order to introduce structural disorder, i.e., to destroy the colloidal
crystal, but the influence of the composition on the position of the HG was not
elucidated. Indeed, in the top panel of Fig. 5.11 it seems that the HG for the pure
360 nm opal and the 1:1 hybrid come more or less at the same position. In this
section, an experiment is presented in which the number ratio of two kinds of PS
colloids with clearly different size (180 and 360 nm) is varied in a wide range.
The HG for these systems infiltrated with PDMS is observed and the changes in
width and position are discussed; of particular interest is the question if it is
possible to tailor this band gap by simply controlling the size composition of the
particles.
The set of samples has been realized by Retsch mixing distinct amounts of
180 nm PS beads and 360 nm PS beads in aqueous dispersion and preparing the
hybrids by vertical lifting deposition under continuous stirring in order to
guarantee the homogeneous distribution of both particles in the whole sample.
Doing so, hybrids with composition between 1:1 and 40:1 180:360 nm particles
number ratio have been prepared, i.e., the volume ratio varied between 1:8 and 5:1
for the smaller particles compared to the larger ones. After infiltration with PDMS,
BLS was performed in transmission geometry and the dispersion relation was
obtained fitting the spectra with Lorentzians in the q-range between about 0.006
and 0.020 nm-1. The result is shown in Fig. 5.14.
Figure 5.14 contains the experimental band diagrams for the mentioned
systems as well as for bare 180 nm PS particles. The position of the HG is marked
by colored rectangles. Note that for all systems only one HG between &4 and
5.5 GHz was found in the experimentally accessible range, only for the 1:1 hybrid,
a second HG was observed between &2 and 2.5 GHz. By comparison with the HG
from the bare 180 nm PS particles and taking into account the scaling of the HG
with 1/d, it is obvious that this gap originates (mostly) from the 360 nm particles.
The 1:1 hybrid is the only one whose volume fraction of the larger particles
exceeds that of the smaller spheres. Anyhow, in the rest of this section, I want to
focus on the higher frequency HG above 4 GHz.
Indeed, when comparing the HGs of the bare 180 nm particles and of the hybrids,
one finds that the HG for every hybrid is shifted to lower frequencies in a mostly
systematic way. The higher frequency HG of the 1:1 hybrid is distinctly lower and
smaller than that of the opal. With increasing amount of 180 nm particles in the
mixture, the gap widens up and is shifted to higher frequencies. This is true at least
for the 9:1 and the 20:1 hybrid, the 40:1 hybrid, however, seems to contradict the
trend, although its HG is still broader and at clearly higher frequency than that of the
112 5 Phononic Behavior of Colloidal Systems
1:1 hybrid. On the other hand, when ignoring the lowest point in the upper branch of
the 40:1 sample, also this sample would follow the trend.
When comparing the opal with the hybrids, one should note that for an fcc
colloidal crystal of course the filling fraction (/ = 0.74) is different than for a
disordered system. Approximating a uniform filling fraction of / = 0.65 for all
hybrids [60], which is admittedly a bit arguable especially for the extreme ratios,
one can estimate that for the infiltrated opal the effective longitudinal sound
velocity ceff should be higher by about 6.3 or 8.5%, applying the EMT of
Gaunaurd and Wertman [31] or Wood’s law [30], respectively (cf. Sect. 5.2)—in
reasonable agreement with the experimental ceff. Since the position of the band
gaps scales with ceff, it is possible to normalize the frequencies by that. Doing so,
also the HG of the 180 nm opals is shifted towards somehow lower frequencies,
still following the trends mentioned above. In Fig. 5.14 the normalized (relative to
the amorphous systems) HG of the PS 180 nm opal is symbolized by the open
black rectangle.
To rationalize the findings in this experiment one must go back on the nature of
the HG (Sect. 5.3.1). The HG originates from the level repulsion between the
bands of the linear acoustic mode and those from localized modes, i.e., the
eigenvibrations inside the spheres. The generation of a band originating from these
eigenvibrations can be described in analogy to the tight coupling method for the
electronic states in solids, in which the overlap of atomic orbitals is assumed to be
sufficient to require corrections of the picture of the isolated atom, but not so much
to render the atomic description completely irrelevant [29]. This theory predicts
that in a sample of N atoms each electronic state consists of an N-fold degenerated
level. If the atoms come closer together, i.e., if their wave functions overlap
5.3 Band Gaps in Polymer Opals and Disordered Systems 113
significantly, the levels broaden up into bands. In the analogy of the phononic
system, the individual spheres play the role of the atoms, and instead of electronic
wave functions one deals with the acoustic resonance modes of the spheres
(as ‘acoustic wave functions’), which are both described by the same mathematics
(cf. Sect. 2.3.3 in Chap. 2).
This means that in order to obtain a strong band, the spheres must be in close
contact to each other and the acoustic wave function must overlap with that of the
next neighbor. Since the spheres are close-packed (or randomly close-packed) in
all cases, bands can be assumed. The question is how do modes of the same type
but from different particles sizes interact with each other. From symmetry
considerations, there is no reason for them not to interact, i.e., to hybridize.
However, if their energy levels are too different, they should be not able to interact
anymore. Anyhow, in the present experiment the frequency of the gaps is shifted,
which is a hint on a shift of the position of the resonance band [in this case of the
(1, 2)-mode], too. The most reasonable explanation is that, although there is a
significant difference in energy for 180 and 360 nm spheres, the (1, 2) modes of
the 180 nm spheres couple with those of the 360 nm spheres, leading to shift to
lower frequencies with increasing fraction of the 360 nm spheres wave function.
On the other hand, the coupling seems to be weak enough to allow two HGs for
the 1:1 hybrid. Since the higher frequency HG is clearly shifted towards lower
frequency, it is obvious that there cannot be two fully decoupled bands, at least
there must be some influence of the bigger spheres on the localized mode of the
smaller spheres. For lower number ratios of the larger spheres the low frequency
HG completely disappears, which is another hint for the rightness of the
conclusion that this gap must originate from the 360 nm particles, which are too
separated in the other hybrids to interact effectively. The smaller width especially
for the 1:1 hybrid can be rationalized by the larger average sphere to sphere
distance of the small spheres in the sample.
In summary, one can conclude that the experiment can be principally captured
qualitatively by simple hybridization considerations. However, there are still open
questions that could only be solved with additional experiments and—even more
needed—strong theoretical support. From this section, I would like to retain the
message that the distribution of colloids of different size has an influence on the
hybridization gap. Under certain circumstances, the mixing of particles of different
sizes may be a way on which future studies to tailor the band gap could proceed.
In the previous sections of this chapter it was shown that the elastic contrast
between matrix and scatterers is crucial for phononic phenomena. Another
important role is played by the density contrast, which is, however, usually quite
weak for polymer/liquid systems. In this section, experiments on silica colloids in
liquid or elastic matrix are shown, i.e., the elastic contast is relatively high as well
114 5 Phononic Behavior of Colloidal Systems
as the density contrast is. In such systems, additional effects can be found that
cannot be fully captured by theory (and the explanations in the previous chapters),
yet.
(a) (b)
(c)
Fig. 5.15 a Development of the acoustic signal in the BLS spectrum of d = 375 nm silica
spheres in SR256 (/ = 0.74) for 30° B h B 75°. b Exemplary spectra at h = 40° for low
(/ = 0.34) and high (/ = 0.74) filling fraction. c Experimental dispersion relations for dense
packed silica spheres with d = 219 or 275 nm in SR256. For comparison also pure SR256 as well
as silica with d = 375 nm and / = 0.34 are shown
The appearance of the bending requires high filling fraction, since the suspension
with / = 0.34 (black circles) shows only the effective medium acoustic phonon,
which appears at slightly higher frequencies than for the pure matrix. This is
expected by effective medium theory (cf. Sect. 5.1.2) since the density of the silica
inclusions is much higher than that of the matrix. The high density contrast leads to a
strong scattering of the phonons on the hard spheres; i.e., the phonons travel mostly
through the liquid matrix, and hence, ceff is nearly the same in both cases.
Correspondingly, the linewidth at a given q is smaller for lower filling fraction
as is shown for one exemplary angle (h = 40°) in Fig. 5.15b. C is inversely
proportional to the phonon lifetime. If the phonon path is disturbed by many
scatterers, the phonon lifetime decreases drastically. In fact, if C f ; as it is found
at higher qs for / = 0.74, the free phonon pathlength is akin to the phonon
wavelength; i.e., the phonon cannot travel anymore within the medium (the group
velocity is zero), the system is overdamped. However, at higher qs the group
velocity comes back to its value before the damping.
If one wants to summarize the experimental findings, one should start with the
fact that there is no band gap. At a certain q-range the group velocity becomes
116 5 Phononic Behavior of Colloidal Systems
Fig. 5.16 Normalized theoretical band diagrams for fcc opals of SiO2 spheres in SR256 matrix
along two crystallographic directions. The experimental dispersion relations in Fig. 5.15 are
shown as circles and squares
understood as a multiple-scattering effect. Its details, however, are still unclear and
must be developed by theoreticians in the future towards a concrete, descriptive
explanation.
Another realization of silica based phononic materials are melt compressed films
with silica inclusions [22, 67, 68]. The technique of melt compression is presented
in Sect. 3.2 in Chap. 3 When choosing silica–poly(ethyl acrylate) (PEA) core–shell
particles as the starting materials, films can be pressed with different filling
fraction of silica particles in a rubbery PEA matrix, i.e., with different spacing of
the silica spheres in an fcc lattice.
In this section, the results of BLS measurements on several silica–PEA films is
presented, with dSiO2 ¼ 216 or 253 nm: The total diameter of the initial core–shell
particles was chosen between dCSP = 308 and 498 nm, leading to filling fractions of
/ = ðdSiO2 =dCSP Þ3 between / = 0.08 and 0.36 in the pressed films. The crystallo-
graphic orientation of the silica spheres in the pressed film is known to be an fcc
lattice with orientation of lines of next neighbors in the radial direction from the
center of the film [69]. Due to the melt compression, in the film the silica cores come
nearer then in the initial opal, as is schematically shown in Fig. 5.17b. The shortest
intercore distance becomes (0.74)1/3dCSP along the direction marked by the white
118 5 Phononic Behavior of Colloidal Systems
arrow, which is also the main orientation of q in the following experiments. The
pffiffiffi
corresponding lattice constant for an fcc-lattice is a ¼ 2ð0:74Þ1=3 dCSP :
Figure 5.17a shows five exemplary spectra for such a silica–PEA film
(dSiO2 =dCSP ¼ 253=387 nm) at different qs along the direction marked by the white
arrow in panel b. The spectra are represented by a fit of one (lowest q) or two
Lorentzian signals on the Anti-Stokes side. Around q 0:008 nm-1 a band gap
opens up. The corresponding band diagram is shown in Fig. 5.17c, normalized by
the effective medium sound velocity ceff, measured in the long wavelength limit,
(a) (b)
(c)
(d) (e)
Fig. 5.17 a Experimental spectra for a melt compressed film (dSiO2 =dCSP ¼ 253=387 nm) at
different qs, represented by Lorentzian lines. b Scheme of the result of melt compression
(Sect. 3.2 in Chap. 3); the white arrow indicates the orientation of q. c Normalized dispersion
relations for three different films. d Insensitivity of a melt compressed film (217/329 nm) to
stretching e and the orientation of q
5.4 Band Gaps in SiO2 Colloidal Systems 119
5.5 Materials
The synthesis of the polymer and silica colloids is already described in Sect. 4.4 in
Chap. 4. Opals and disordered opals (hybrids) are obtained by vertical lifting
deposition described in Sect. 3.1 in Chap. 3.
For the infiltration of opals shown in Fig. 5.3 silicon oil and PDMS made in
house or purchased from Sigma Aldrich were used with no further purification.
2(2-Ethoxyethoxy) ethyl acrylate (SR256) used in the suspensions was purchased
from Sartomer with no further purification.
5.5 Materials 121
The preparation of the compression molded films is explained in Sect. 3.2 in Chap. 3.
The synthesis of the initial core–shell particles was performed by my coworker
Diana Kiefer at DKI, Darmstadt and is briefly described here [22, 67, 70].
The synthesis starts with the preparation of the silica core following the Stöber
process and subsequent step-by-step growing as described in Sect. 4.4 in Chap. 4.
The silica core’s surface was functionalized by acryl silanes on which a thin
(*5 nm) PMMA interlayer was crafted by emulsion polymerisation of methy
methacrylate (MMA) together with some allyl methacrylate (ALMA) monomer,
using sodium dodecylsulfate (SDS) as an emulsifier and ammonium peroxodisulfate
and sodium dithionite as a redox initiator system. Finally the PEA shell was crafted
onto the allylic double bonds of the ALMA, utilizing again SDS as emulsifier.
References
6.1 Introduction
A wide range of packing structures are available to glasses, with more efficient
packing leading to higher moduli materials [1]. Aging a glass allows for better
packing and a higher modulus, but even long aging times increase the modulus by
only a few percent [2, 3]; preparing high modulus materials in this manner is
impractical. In this chapter it is shown that physical vapor deposition can be used
to circumvent this kinetic limitation and produce glasses whose moduli exceed
those of the ordinary glass by up to 19%. These high modulus glasses resist
thermal treatment and take at least 104 times longer than the structural relaxation
time to transform to the supercooled liquid. The ability to easily produce high
modulus glasses will prove to be useful for fundamental investigations and coating
technologies.
Unlike their crystalline counterparts, glasses have a nearly limitless array of
packing arrangements. As a supercooled liquid is cooled, molecular motions
eventually slow to such an extent that equilibrium cannot be maintained. Below
this transition temperature Tg, a mechanically stable, non-equilibrium glass is
formed. Glasses slowly evolve towards equilibrium (i.e., aging) in a process that
optimizes packing and creates higher moduli materials. The structural relaxation
time sa dictates the rate at which this process takes place, and due to the steep
temperature dependence near Tg, sa is on the order of days only a few degrees
below the glass transition temperature Tg. The aging process thus changes the
moduli so slowly that in practice changes of only a few percent are possible [2, 3].
If high modulus amorphous materials are to be utilized for science and technology,
new preparation techniques are needed which circumvent these kinetic restrictions
and allow for more optimized amorphous packing [4].
This section shows that high modulus glass materials can be made efficiently with
physical vapor deposition. Using this preparation technique, one can avoid the
kinetic limitations of aging and prepare high modulus glasses in a matter of hours.
Enhanced dynamics at the surface of amorphous materials allows for rapid con-
figurational sampling in the top few nanometers [5–9]. Vapor deposition can build
an efficiently packed amorphous material in a layer-by-layer fashion by taking
advantage of the enhanced surface dynamics and thus is not limited by the slow
relaxation dynamics of the bulk [10]. The mechanical properties of vapor-depos-
ited films are determined using Brillouin light scattering (BLS) spectroscopy.
Because of the non-destructive nature of BLS, the moduli of the as-deposited
glass, the supercooled liquid, and ordinary glass (created by cooling the liquid) can
be determined from a single sample.
Physical vapor depositions were performed by my coworker Kenneth Kearns
in Prof. Mark Ediger’s group at University of Wisconsin, Madison, separately on
two organic glass-forming materials: indomethacin (IMC, Tg = 315 K) and tris-
naphthylbenzene (TNB, Tg = 348 K). These two molecules are well-known glass-
formers and their glasses have been previously prepared with physical vapor
deposition [10–12]. During deposition, the temperature of the substrate Tsubstrate
and the deposition rate are the important control parameters. For this work,
Tsubstrate was held near 0.85 Tg, i.e., 265 K for IMC and room temperature
(&295 K) for TNB. The rate of the deposition in all cases was 0.2 ± 0.03 nm/s. It
has previously been shown that these deposition conditions produce glasses with
low enthalpy, high density, and high kinetic stability [10–14]. Tsubstrate was con-
trolled by attaching the SiO2 substrates to a copper temperature stage (Fig. 6.1a).
The deposition rate was controlled by adjusting the temperature of the crucible.
Further details are given in Methods (Chap. 3).
In the BLS experiment, the incident laser polarization was chosen to be per-
pendicular to the scattering plane. Scattered light polarized perpendicular and
parallel to the incident polarization was measured in separate experiments, pro-
viding access to scattering from longitudinal and transverse phonons, respectively.
The scattering from these two polarizations is shown in Fig. 6.1d for vapor-
deposited IMC glass at 298 K (lower panel) and supercooled liquid IMC at 336 K
(upper panel). Stokes and anti-Stokes shifts by the longitudinal (L) and transverse
(T) phonons are observed to the left and right of the Rayleigh line region (shaded
area), respectively. The vertical lines drawn on the Stokes side of the spectrum
illustrate the temperature-independent scattering of the SiO2 substrate and the
temperature-dependent scattering of the IMC. The absence of longitudinal pho-
nons in the transverse spectrum indicates no birefringence in the vapor-deposited
film within the sensitivity.
The peaks in the BLS spectra provide access to the phase velocities cl,t for the L
and T polarizations and through this route the moduli can be obtained. Spectra
similar to those found in Fig. 6.1d were obtained at multiple temperatures for both
6.2 BLS Experiments on IMC 125
(a) (d)
(b) (c)
Fig. 6.1 a Schematic representation of the physical vapor deposition chamber with temperature
controlled substrate. b, c Chemical structures of TNB and IMC, respectively. d Representative
BLS spectra for a stable vapor-deposited IMC glass (lower panel) and supercooled liquid (upper
panel) at the wave vector q = 0.0136 nm-1. (Inset transmission scattering geometry with scat-
tering angle h of 70°.) The Rayleigh line region (shaded area) was removed for clarity. Both the
longitudinal (L, black) and transverse (T, red) spectra are shown. Longitudinal and transverse
phonon scattering for the IMC glass (LIMC and TIMC, respectively) and amorphous SiO2 substrate
(Lsubstrate and Tsubstrate, respectively) are indicated in the lower panel. The vertical lines drawn
near the Stokes scattering peaks illustrate the difference between the supercooled liquid and
as-deposited glass [15]
IMC and TNB. Each peak was fitted with a Lorentzian lineshape yielding the
linewidth CL,T and the peak frequency fl,t. The corresponding phase velocities cl,t
for the L and T polarizations were calculated using cl,t = 2pfl,t/q, and are plotted in
Fig. 6.2. Four experiments are shown in Fig. 6.2; in each case, the as-deposited
stable glass was heated to a temperature above Tg, held isothermally until equi-
librium was attained, then further heated, and finally cooled. During the initial
heating of the stable glass (SG), cl and ct have high values as compared to the
ordinary glass (OG), indicating high moduli. During isothermal annealing above
Tg (arrow 2), significant decreases in the phase velocities are observed. This
change, discussed in detail below, signifies the transformation of the stable glass
into the supercooled liquid (SCL). Once the transformation is complete, the
sample, now a supercooled liquid, is heated still higher and then cooled to 295 K
(arrows 3 and 4). A characteristic kink in the temperature-dependent phase
velocity indicates Tg and the value obtained is in reasonable agreement with the
value obtained from differential scanning calorimetry [11, 12].
From the phase velocities cl,t in Fig. 6.2, the moduli of the stable glass and the
ordinary glass can be calculated. The longitudinal bulk modulus M ¼ .c2l ; the
shear modulus G ¼ .c2t and the commonly used Young’s modulus E = 9MG/
(3M ? G) can all be determined from c. For the stable as-deposited glass of IMC,
M, G, and E moduli were determined to be 8.3, 1.7, and 4.9 GPa at Tg,
126 6 Smaller than Colloids: Characterization of Stable Organic Glass
(a) (b)
Fig. 6.2 a Temperature dependence of phase velocity c for IMC. The longitudinal (squares) and
transverse (circles) c are shown for the stable as-deposited glass (SG), the supercooled liquid
(SCL), and the ordinary glass (OG). Arrows indicate the progression of the heating and cooling
cycles. Arrow 2 indicates the isothermal transformation of the SG to the SCL at Tg + 10 K. b
Temperature dependent longitudinal and transverse c of TNB for the SG, OG and SCL. The
thermal cycle is similar to the one described for a [15]
respectively. The moduli calculated for the ordinary glass of IMC are smaller than
those obtained for the stable glass samples. At 312 K, M for the stable as-deposited
IMC glass is 14% greater than the ordinary glass while G and E are 19% greater.
A similar situation holds for TNB where M, G, and E are 10, 15, and 14% greater,
respectively, for the stable glass samples as compared to the ordinary glass. To
obtain these values, the published densities of 1.31 g/cm3 for IMC [16] and 1.16 g/
cm3 for TNB [17] were used. The stable as-deposited glass was assumed to be
1.5% more dense than the ordinary glass in accordance with previous measure-
ments on TNB vapor-deposited under similar conditions [10, 14]. Notably, the
Poisson ratio r has the same value (0.36 ± 0.01) in both stable and ordinary glass
samples and increases for the supercooled liquid as expected [18].
In contrast to the results presented here, aging experiments on glass-forming
systems show only small improvements in mechanical properties during aging.
BLS experiments were performed on glycerol by quenching from above Tg to
Tg -4.2 K; in this experiment, a 0.6% increase in modulus was achieved over
20 days [3]. For silicone oil, temperature down-jumps of 1.5 K near Tg resulted in
a 2.7% increase in the modulus during the 104 s equilibration time [2]. Based on
calorimetry measurements [10, 12], it was previously estimated that IMC and TNB
samples vapor-deposited under the conditions utilized here have properties similar
to those expected after aging an ordinary glass for more than 300 years. Thus it is
not surprising that aging for hours or days does not produce the large modulus
changes reported here. Pressurization experiments have been shown to change
moduli by up to 20%, but this effect is reversible and upon pressure release, the
system reverts back to the ordinary glass value rendering it of little use for
potential applications [19].
6.2 BLS Experiments on IMC 127
Notably, the high modulus IMC and TNB glasses produced by vapor-deposition
exhibit remarkable thermal stability. Figure 6.3a and b shows the change in cl
and ct for vapor-deposited IMC during isothermal annealing above Tg. In these
experiments, the temperature of the stable as-deposited glass was increased to
325 K in about 15 min and then held for several hours. Both phase velocities (and
moduli) remain relatively unchanged for the first 120 min of the experiment. After
this induction time, the phase velocities begin to decrease from the high values of
the stable glass to the supercooled liquid values. Approximately 200 min is needed
to complete the transformation to the supercooled liquid. To put this value into
context, the structural relaxation time sa of the IMC supercooled liquid at 325 K
is about 1 s based on dielectric spectroscopy studies [20]; thus the isothermal
transformation of stable vapor-deposited IMC into the supercooled liquid requires
about 104sa. Similar results are obtained for TNB where the stable as-deposited
glasses required 104.8sa to transform into the supercooled liquid at 358.2 K [21].
Conceptually similar experiments have been performed on ordinary glasses that
have stabilized by aging. For example, Kovacs aged poly(vinyl acetate) for
2 months below Tg and then performed dilatometry during isothermal annealing
above Tg. He observed the transformation into the supercooled liquid in a period of
(a) (c)
(b)
Fig. 6.3 BLS experiments on vapor-deposited IMC glasses during isothermal annealing above
Tg (325 K). a Longitudinal, cl, and b transverse, ct, phase velocity changes as a function of time.
Sigmoidal fits are given as guides to the eye. Two different samples (squares and circles) are
shown in a to demonstrate the reproducibility of the transformation. In a and b, the full width at
half height linewidth, C, is shown as crosses. c Spectral line shapes and Lorentzian fits (solid
lines) at three different times during isothermal annealing [15]
128 6 Smaller than Colloids: Characterization of Stable Organic Glass
35sa [22]. Thus, in comparison to glasses prepared by cooling a liquid and then
aging, the vapor-deposited IMC and TNB samples maintain their extraordinarily
high modulus for a remarkably long time.
Changes observed in the BLS spectral linewidth during isothermal annealing
provide insight into the mechanism by which the stable glass transforms into the
supercooled liquid. Figure 6.3a shows that a maximum in C is reached near the
midpoint of the transformation (C is read from the right axis). A similar feature is
observed in Fig. 6.3b although the smaller signal of the transverse scattering
makes this less apparent. The observation of a change in C during the transfor-
mation of a glass to a liquid is unprecedented. The increase in C indicates that the
sample is heterogeneous on the 500 nm (*2p/q) length scale during the trans-
formation. Figure 6.3 shows the BLS spectral lineshapes at three times during
isothermal annealing at 325 K. Before the transition begins (30 min) and after the
transition ends (270 min), narrow lines are observed. In contrast, near the midpoint
of the transformation (180 min), a broad line is observed. The shape of the line
suggests that the sample is a mixture of the supercooled liquid and the stable glass
at this stage. This point is illustrated by the dotted red curve in Fig. 6.3c which is a
linear combination of narrow lineshapes from the 30 min (stable glass) and
270 min (supercooled liquid) data. This conclusion is consistent with a recent
quasi-isothermal differential scanning calorimetry study of these materials [13].
While that work suggested that stable glasses transform to the supercooled liquid
via two-phase intermediate state, the spatial extent of these regions (at least
500 nm) has not been known prior to this BLS study.
Significantly increasing the modulus of amorphous solid materials using con-
ventional aging methods is a time consuming and ineffective process. The work
presented in this chapter shows that physical vapor deposition can be used to
prepare glasses with moduli that are as much as 19% greater than those made by
cooling the liquid. These glasses resist thermal treatment and retain their high
modulus values for at least 104 times longer than the structural relaxation time of
the glass at temperatures significantly above Tg. During this slow transformation, a
mixture of supercooled liquid and glass is observed on length scales greater than
500 nm. High modulus vapor-deposited glasses may prove useful for applications
because of their remarkable mechanical properties but also their extraordinary
thermal stability. Given the known correlation between amorphous packing and
modulus, new insights into this packing could be realized through fundamental
studies of these materials.
IMC and TNB stable glass films were prepared by my coworker Kenneth Kearns at
University of Wisconsin in Madison, WI.
6.3 Materials and Methods 129
IMC was purchased from Sigma Aldrich and was used without further purifi-
cation. TNB was synthesized by McMahon and co-workers using published
methods [23]. SiO2 substrates (5 mm diameter and 0.15 mm thick, Fisher Scien-
tific coverglass) were attached to a copper temperature stage using double-sided
conductive carbon black tape (SPI Supplies). The temperature of the substrate
during deposition was controlled using a Lakeshore 340 temperature controller. To
deposit the IMC and TNB glass films, a quartz crucible containing the crystalline
material was heated inside a vacuum chamber (*10-8 torr) such that the desired
rate of 0.2 nm/s was achieved. The rate of deposition was controlled by main-
taining the temperature of the quartz crucible, and the rate was monitored with
a quartz crystal microbalance (Sycon instruments). Deposition continued until a
10–15 lm film was deposited. After deposition, the substrates were removed from
the deposition chamber, placed in desiccant and stored in dry ice prior to analysis.
BLS was performed in a transmission geometry to obtain the phase velocity and
thus modulus values. The transmission, as opposed to the commonly used
reflective geometry, was used to remove the influence of refractive index on the
measurements, and the details of this geometry are described in detail elsewhere
[24–26]. For all experiments, the sample holder was initially flushed with argon
gas and desiccant was placed in the holder to eliminate any possible plasticization
by atmospheric water [27, 28]. The sample temperature was monitored with a
platinum RTD and controlled to within 0.5 K with a temperature controller which
was built in-house. For each data point in Fig. 6.2, the temperature was allowed to
stabilize for 15 min and then the BLS spectra were acquired for 5 min. The
heating/cooling rate between each temperature measurement was approximately
2–3 K/min. The temperature history establishes an effective heating/cooling rate
of approximately 0.1 K/min. For isothermal experiments, the temperature was
changed from room temperature to the annealing temperature in 15–20 min.
References
7.1 Conclusions
by others, is disputed (Sect. 4.2.1 in Chap. 4). Further, the utilization of densely
packed silica spheres in an index matching liquid led to the first verification of the
theoretical predictions for the q-dependent amplitude of the eigenmode spectrum;
the frequency of these localized modes is expectedly q-independent.
After these fundamental questions concerning the spectrum of the resolved
modes in homogeneous colloidal spheres have been convincingly addressed, the
effect of the elastic constants on the particle vibrational modes was best addressed
in the case of random copolymer (PMMA-PnBA) spherical particles, where
the elastic constants vary with the composition (Sect. 4.2.3 in Chap. 4). The
randomness of the copolymer and the amorphous state of the particle was verified
by the presence of single Tg by DSC, further confirmed by the single a-relaxation
obtained by dielectric spectroscopy, i.e., both kind of segments feel the same energy
landscape. The systematic red shift of the eigenfrequencies with increasing
composition of the softer (PnBA) component as well as the effective medium
velocities measured in bulk copolymer films were theoretically captured when a
higher ‘effective glassy sound velocity’ was assumed for the PnBA in the
copolymer than in its bulk state.
Two additional elaborate BLS experiments on nanostructured core–shell hybrid
systems were presented herein: (1) hard silica colloids surrounded by a soft
polymer shell show eigenmode spectra that are dominated by the scattering from
the soft polymer shell (larger compressibility). Nevertheless, due to the boundary
conditions and displacements in the composite material, it becomes feasible the
extraction of the elastic constants in parts, core and shell at the nano- and
mesoscale. The spatial confinement causes a hardening of the thin polymer
chain (Sect. 4.2.2 in Chap. 4) [5]; (2) in the inverse particle architecture, where the
polymeric core is embraced by a thin silica shell, the former dominates the
eigenmode spectrum. A temperature dependent study has revealed a striking
hardening of the polymeric core as compared to the bare polymer spheres implying
a modified glassy state under confinement. The glass transition temperature Tg of
the core, however, remained virtually constant whereas the thin silica shell assured
a robust shape persistent polymeric core at temperatures well above its Tg (‘nano
armor’, Sect. 4.3.1 in Chap. 4) [6].
Since the hypersonic phonons are in the order of magnitude of few hundreds
nanometers, they do not only probe the effective medium structured in a molecular
scale (like in the copolymers), but also in the mesoscale. By infiltration of the dry
samples with a liquid close to optical matching, it is possible to overcome the
multiple light scattering and, by that, to get access to the collective acoustic behavior
of the colloidal systems. Pure effective medium behavior is observed by BLS for
such systems in the long wavelength limit, i.e., for small q’s, where the phonon
wavelength exceeds the structure length of the colloidal system. Section 5.1.2 in
Chap. 5 presents a short overview over the most important established effective
medium theories, while in Sect. 5.2 in Chap. 5 their applicability on a system of
defect doped liquid infiltrated opals is presented.
At higher q’s, BLS can be used to record directly the dispersion relations that
led to the first demonstration of a distinct hypersonic Bragg gap at the edge of the
7.1 Conclusions 133
first Brillouin zone of a colloidal crystal by Cheng et al. [7]. In this thesis, the
resonant character of the particles was realized to demonstrate, for the first time,
the presence of an additional band gap—the hybridization gap, theoretically
predicted six years earlier. This HG originates from the interaction of the acoustic
band of the effective medium and bands from the multipole modes of
the individual particles (Sect. 5.3.1. in Chap. 5), i.e., the eigenmodes. Hence, its
realization demonstrates the strong correlations between the individual colloid’s
elastic properties (the ‘music’) and the phononic characteristics of the ensemble
(the ‘concert’). For polymer colloids infiltrated with an index matching liquid like
PDMS or silicon oil, the HG is found to open up at frequencies below that of
Bragg gap.
Increase of the mechanical mismatch by going from soft to hard opals based on
silica spheres significantly changes the propagation characteristics (Sect. 5.4 in
Chap. 5). This includes an unseen bending in the band diagram of polycrystalline
silica colloids in a liquid matrix as well as a Bragg gap, which is peculiarly robust
against structural deformation, in silica/rubber phoXonic films.
7.2 Outlook
Phononic materials might have interesting potential applications starting with the
obvious use as acoustic shields or vibration isolators. However, for most practical
applications, gap frequencies in the range of sonic or ultrasonic seem to be of
higher interest than those in the hypersonic range. On the other hand, if the
blocked phonon’s wavelength is comparable to the structure dimensions, i.e.,
k * d then the structure’s components for ultrasonic devices must be near to the
millimeter-range and even larger to stop sound between kHz and Hz. The ultimate
goal would be to create materials, which can realize gaps at wavelengths much
larger then the inherent length scale of the gap material (k d). Ping Sheng and
coworkers demonstrated an approach going in this direction [8]. They prepared
locally resonant sonic materials by combining a spherical lead core and a silicone
rubber coating in the millimeter size range and building up a crystal out of these.
The resulted superstructure exhibits a band gap around 400 Hz, which corresponds
to a wavelength two orders of magnitude bigger than the diameter of the spheres.
Like HG the explanation is due to localized modes mostly in the soft shells.
Advances in colloidal science can be utilized to provide materials with new
functions. For example core–shell particles with varying composition, hybrid
materials, or materials with even hierarchical order can allow tailoring the
phononic properties of small but smart materials to frequencies, where a broad
range of applications is conceivable.
When we turn our attention back to the manipulation of the phonon flow in the
GHz range, hypersonic phononic materials hold promise to control heat flow in
insulators and semiconductors. The typical phonon wavelength responsible for the
heat transport is even smaller (&10 nm) than the structures investigated in this
134 7 Concluding Remarks
thesis. If we will control the flow of elastic energy by building phonon guiding
devices in full analogy to the lossless light guidance in photonics or by building
devices with a strong anisotropy of phonon transmission, we can also attain new
knowledge in advanced heat management. In this direction, colloidal science could
offer new materials, e.g., multilayer structures with different particle sizes in each
structure that could split and direct the elastic energy. Other analogies to
photonics, e.g., acoustic superlenses already exist [9], but still impose challenges
for high frequency acoustics. It should be mentioned that hypersonic phononics
can simultaneously act as photonics for wavelengths in the visible spectrum
allowing the realization on photonic and phononic band gaps at the same time
(PhoXonics) [10]. The coupling of optical and mechanical degrees of freedom can
be applied to influence the optomechanical dynamics of small systems, utilizing
the radiation pressure of light [11]. The upper frequency limit for such systems is,
so far, in the MHz-range, however, colloidal systems like those presented in this
thesis may expand the investigation of such effects to the GHz-range, giving raise
to novel fascinating phenomena.
In parallel to the development of materials with new phononic functions, there
is also a need for new experimental techniques to measure the band diagrams of
non-transparent structures exhibiting also strong phononic gaps. BLS is a powerful
technique for transparent samples with serious shortcomings in opaque systems
such as dry opals and in measuring hypersonic transmission spectra. The main
problem is the availability of a selective and continuous generation of phonons in
the GHz-range, which will be hopefully overcome in the near future.
References
n0=1
n
ks
ki
2
q
1 qpara
n
n0=1
2pn
qpara ¼ ðsin b þ sin cÞ ð8:7Þ
k
2p 4p h
qpara ¼ ðsin a þ sin ðh aÞÞ ¼ sin : ð8:8Þ
k k 2
In the last equation, valid only in this special case, the refractive index disappears.
c ¼ 180 a ð8:11Þ
¼ 90 d ð8:13Þ
n0=1
n
ks
qperp
q
ki
n
n0=1
11 1 1
c ¼ 180 90 sin sin a þ 90 sin sin ð180 h aÞ
n n
1 1
¼ 180 sin1 sin a sin1 sin ðh þ aÞ ;
n n
ð8:15Þ
qperp ¼ q cos ðb gÞ
4pn 1 1 1 1 1
¼ sin 180 sin sin a sin sin ðh þ aÞ
k 2 n n ð8:19Þ
1 1 1 1 1
cos sin sin a sin sin ðh þ aÞ :
2 n n
By comparison with Eq. 8.16 it is clear that q ¼ qperp if the last term on the right
side becomes one. Due to the periodicities of the inverse sine function, this is the
case for
h 180
a¼ : ð8:20Þ
2
i.e., when chosing this special geometry in the BLS experiment, the wave vector
lies fully perpendicular to the substrates surface. However, the careful comparison
with Fig. 2.6 in Chap. 2 shows that the fulfillment of this geometry would lead to a
reflection of the laser directly into the Fabry–Pérot. Therefore, only geometries
very near to that can be chosen in practice, in order not to destroy the detector, and
the qs probed in reflection geometry are only approximately equal to the qperp s.
Curriculum Vitae
Tim Still
Date of Birth: April 3, 1982
Place of Birth: Neuwied, Germany
Nationality: German
ResearcherID: A-3286-2009
Publications
2010
139
140 Curriculum Vitae
2009
2008
2009
2008
A E
Anderson localization, 116 Effective medium, 7, 66, 69, 91, 94, 98, 100ff,
Anisotropy, 134 114ff, 132
Eigenmode, 15, 29ff, 47, 53ff, 96, 98, 107,
116, 120, 131ff
B Electron microscopy, 38f, 56f, 61ff, 72ff, 101f
Band diagram, 7, 43, 45ff, 89, 91ff, 108ff,
133f
Band gap, 1ff, 36, 45, 54, 89ff, 104ff, 133 F
Bragg gap, 1ff, 42, 45, 89, 97ff, 104, 116, 119, Fabry–Pérot, 6, 20f, 27ff, 138
132 Finesse, 26ff
Bulk modulus, 13ff, 125 Finite element modelling, 44, 51, 54ff, 63, 76,
131
Free spectral range, 26ff, 56, 70
C
Calculation methods, 43ff G
Centrifugation, 5, 36, 69f, 84, 86 Glass/glass transition, 4, 6, 7, 25, 30, 35, 37,
Colloidal mixture, 42, 111ff, 131 57, 63ff, 69ff, 105, 109, 123ff, 131ff
Compressibility, 23f, 68, 72
Core–shell particle, 5, 37, 48, 55f, 63, 71ff,
77ff, 85f, 117ff, 121, 132f H
Correlation, 38f, 78, 128, 133 Hybrid, 5, 32, 37, 55, 61ff, 72, 77, 81, 84, 91,
94, 98, 105ff, 111ff, 132f
Hybridization gap, 7, 54, 91, 94, 98, 105ff,
D 111, 113, 119, 133
Dense packing, 37, 62, 67, 96, 114
Density, 2, 3, 14, 21, 23, 33, 38, 40f, 53ff, 64f,
71, 74, 77, 80ff, 89, 94, 101ff, I
113ff, 124, 131 IMC, see IMC
Density of states, 54, 80, 83, 109, 131 Indomethacin, 124
Differential thermo analysis, 40
Differential scanning calorimetry, 40, 64, 73f,
125ff L
Disorder, 61, 63, 104ff, 131 Lamé coefficients, 11, 13, 30, 46, 89
Dispersion relation, 6, 20f, 29, 32, 43, 51, 61ff, Laser, 5f, 20f, 25, 29, 30, 32, 38, 53, 57, 58,
70, 91ff, 105ff, 132 90, 105
143
144 Index
Layer multiple scattering, 44, 50, 116 Resonance vibration, 20, 45, see eigenmode
Linewidth, 26, 73, 81, 114f, 120 Rigidity, 33, 55, 63ff, 72, 73f, 77
Rubber, 37, 67ff, 73ff, 117, 120, 133
M
Melt compression, 37, 117f S
Molecular weight, 43, 64, 84 Scattering angle, 4, 18ff, 23, 29, 32, 38, 41, 65,
Multiple scattering, 7, 32, 55, 57, 58, 65, 69, 69ff, 105, 125, 135
79, 96, 116, 117, 131 Scattering cross section, 43, 44ff, 56, 58, 60,
Music, see eigenmodes 63, 67
Scattering geometry, 30ff, 39, 56, 65, 99f, 105,
111, 125, 129, 135ff
N Setup (BLS), 25ff, 29
Nanostructure, 5, 21, 53, 55, 71ff, 94, 110, Shear modulus, 11, 13, 15, 57, 66, 68, 73f, 80f,
123, 131f 125
Silica, 4, 6, 7, 20, 37, 55f, 60, 62, 66ff, 77ff,
85f, 100ff, 113ff, 132f
O Spherical waves, 9, 15, 33, 104
Opal, 4, 20, 35, 37, 54ff, 94ff, 133f Strain, 9ff, 119f
Order, 104ff, see disorder Stress, 9, 11ff, 32f, 54
Organic glass, 123ff Supercooled liquid, 123ff
Suspension, 5, 55, 62, 69ff, 78, 98, 105,
114ff
P Synthesis, 4, 63, 73, 77, 84ff, 120f, 128f
Phononic crystal, 2ff, 46, 49, 51ff, 90, 97, 98,
104
Photonic crystal, 1ff, 7, 38, 89 T
PhoXonics, 4, 6, 117, 133f T-Matrix, 48f, 94f
Plane wave, 17f, 34, 43ff, 81f, 91 TNB, see TNB
Poisson’s ratio, 13, 14, 57, 65, 68, 69, 80, 120, tris-Naphtylbenzene, 164ff
126
Poly methyl methacrylate, 55, 63ff, 72, 77ff,
97f, 105ff, 121, 132 U
Poly n-butyl acrylate, 63ff UV/Vis spectroscopy, 40ff, 101f
Polydispersity, 61ff, 82, 85, 109, 131
Polystyrene, 4, 31, 37, 45, 55f, 67ff, 72, 79, 84,
95ff, 105ff V
Pump-probe, 79 Vertical lifting deposition, 35ff, 65, 78, 97, 99,
101, 105, 111, 120, 131
R
Raman scattering, 5, 16, 25, 32, 53ff W
Reciprocal space, 1, 41, 46, 93, 97ff, 107 Wide angle X-ray scattering, 41, 64
Reflectivity, 5, 25ff, 53
Refractive index, 1ff, 18, 21, 25, 31, 42, 57,
69, 96, 102, 107, 129, 136 Y
Relaxation, 64, 123ff, 132 Young’s modulus, 30ff, 65, 125ff