ArticlePDF Available

Abstract and Figures

Many proteins of diverse sequence, structure and function self-assemble into morphologically similar fibrillar aggregates known as amyloids. Amyloids are remarkable polymers in several respects. First of all, amyloids can be formed from proteins with very different amino acid sequences; the common denominator is that the individual proteins constituting the amyloid fold predominantly into a β-sheet structure. Secondly, the formation of the fibril occurs through non-covalent interactions between primarily the β-sheets, causing the monomers to stack into fibrils. The fibrils are remarkably robust, considering that the monomers are bound non-covalently. Finally, a common characteristic of fibrils is their unbranched, straight, fiber-like structure arising from the intertwining of the multiple β-sheet filaments. These remarkably ordered and stable nanofibrils can be useful as building blocks for protein-based functional materials, but they are also implicated in severe neurodegenerative diseases. The overall aim of this article is to highlight recent efforts aimed at obtaining insights into amyloid proteins on different length scales. Starting from molecular information on amyloids, single fibril properties and mechanical properties of networks of fibrils are described. Specifically, we focus on the self-assembly of amyloid protein fibrils composed of peptides and denatured model proteins, as well as the influence of inhibitors of fibril formation. Additionally, we will demonstrate how the application of recently developed vibrational spectroscopic techniques has emerged as a powerful approach to gain spatially resolved information on the structure–function relation of amyloids. While spectroscopy provides information on local molecular conformations and protein secondary structure, information on the single fibril level has been developed by diverse microscopic techniques. The approaches to reveal basic mechanical properties of single fibrils like bending rigidity, shear modulus, ultimate tensile strength and fracture behavior are illustrated. Lastly, mechanics of networks of amyloid fibrils, typically forming viscoelastic gels are outlined, with a focus on (micro-) rheological properties. The resulting fundamental insights are essential for the rational design of novel edible and biodegradable protein-based polymers, but also to devise therapeutic strategies to combat amyloid assembly and accumulation during pathogenic disorders.
Content may be subject to copyright.
Feature article
Amyloids: From molecular structure to mechanical properties
Michael Schleeger
a
, Corianne C. vandenAkker
b
, Tanja Deckert-Gaudig
c
, Volker Deckert
c
,
d
,
Krassimir P. Velikov
e
,
f
, Gijsje Koenderink
b
, Mischa Bonn
a
,
*
a
Max Planck Institute for Polymer Research, Department of Molecular Spectroscopy, Ackermannweg 10, 55128 Mainz, Germany
b
FOM Institute AMOLF, Science Park 104, 1098 XG, Amsterdam, The Netherlands
c
Institute of Photonic Technology, Albert-Einstein-Str. 9, 07745 Jena, Germany
d
Institute for Physical Chemistry & Abbe School of Photonics, University of Jena, Helmholtzweg 4, 07743 Jena, Germany
e
Unilever R&D Vlaardingen, Olivier van Noortlaan 120, 3133 AT Vlaardingen, The Netherlands
f
Soft Condensed Matter, Debye Institute for Nanomaterials Science, Department of Physics and Astronomy, Utrecht University, Princetonplein 5, 3584 CC Utrecht, The Netherlands
article info
Article history:
Received 9 November 2012
Received in revised form
21 January 2013
Accepted 11 February 2013
Available online xxx
Keywords:
Amyloids
Vibrational spectroscopy
Biopolymers
abstract
Many proteins of diverse sequence, structure and function self-assemble into morphologically similar
brillar aggregates known as amyloids. Amyloids are remarkable polymers in several respects. First of all,
amyloids can be formed from proteins with very different amino acid sequences; the common de-
nominator is that the individual proteins constituting the amyloid fold predominantly into a
b
-sheet
structure. Secondly, the formation of the bril occurs through non-covalent interactions between pri-
marily the
b
-sheets, causing the monomers to stack into brils. The brils are remarkably robust,
considering that the monomers are bound non-covalently. Finally, a common characteristic of brils is
their unbranched, straight, ber-like structure arising from the intertwining of the multiple
b
-sheet
laments. These remarkably ordered and stable nanobrils can be useful as building blocks for protein-
based functional materials, but they are also implicated in severe neurodegenerative diseases. The overall
aim of this article is to highlight recent efforts aimed at obtaining insights into amyloid proteins on
different length scales. Starting from molecular information on amyloids, single bril properties and
mechanical properties of networks of brils are described. Specically, we focus on the self-assembly of
amyloid protein brils composed of peptides and denatured model proteins, as well as the inuence of
inhibitors of bril formation. Additionally, we will demonstrate how the application of recently devel-
oped vibrational spectroscopic techniques has emerged as a powerful approach to gain spatially resolved
information on the structureefunction relation of amyloids. While spectroscopy provides information on
local molecular conformations and protein secondary structure, information on the single bril level has
been developed by diverse microscopic techniques. The approaches to reveal basic mechanical properties
of single brils like bending rigidity, shear modulus, ultimate tensile strength and fracture behavior are
illustrated. Lastly, mechanics of networks of amyloid brils, typically forming viscoelastic gels are out-
lined, with a focus on (micro-) rheological properties. The resulting fundamental insights are essential for
the rational design of novel edible and biodegradable protein-based polymers, but also to devise ther-
apeutic strategies to combat amyloid assembly and accumulation during pathogenic disorders.
Ó2013 Elsevier Ltd. All rights reserved.
1. Introduction
Amyloids constitute a fascinating class of biopolymers that
consist of non-covalently bound aggregates of misfolded poly-
peptides, resulting in insoluble brous protein aggregates sharing
specic structural traits. Amyloids are set apart from other bio-
polymers such as DNA and polysaccharides, in that the respective
monomers of amyloids are comparatively large polypeptides which
polymerize to form a so-called cross-
b
structure. Moreover, the
bonds between the monomers are non-covalent. As a result of
misfolding, the individual polypeptides constituting an amyloid
bril fold predominantly into a
b
-sheet structure. The strong, but
non-covalent interaction between the
b
-sheets gives rise to stack-
ing of the peptides into protobrils, which can subsequently
assemble into large brils with a diameter of several nms and a
length up to many microns. Many different biological and articial
peptides can form amyloid structures under the right conditions.
In vivo, amyloids are often associated with neurodegenerative dis-
eases like Alzheimer, diabetes and Parkinsons disease. However
both biogenic and articial amyloids are also widely applied, for
instance, in structuring foodstuffs.
*Corresponding author. Tel.: þ49 6131 3791 60; fax: þ49 6131 379 360.
E-mail address: bonn@mpip-mainz.mpg.de (M. Bonn).
Contents lists available at SciVerse ScienceDirect
Polymer
journal homepage: www.elsevier.com/locate/polymer
0032-3861/$ esee front matter Ó2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.polymer.2013.02.029
Polymer xxx (2013) 1e16
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
Amyloid brils can be regarded as biopolymers very similar to
silk [1], exhibiting common key features: they originate primarily
from unstructured precursor proteins and the denition of the
materials occurs through its specic structural and mechanical
properties rather than through its detailed chemical composition.
The cross-
b
sheet core structure of amyloid brils is very rigid and
confers superior mechanical properties. The brils can exhibit a
Youngs modulus similar to that of silk [2] and an ultimate strength
similar to steel [2,3]. As such, amyloid brils constitute promising
building blocks for bio-inspired materials. The brils form spon-
taneously from a wide range of natural proteins. The structural and
mechanical properties are relatively insensitive to the protein
amino acid sequence [4] and to a wide range of chemical and
biochemical modications [5]. The unique structure of amyloid -
brils renders them relatively robust, even under extreme conditions
such as high and low temperature, the presence of proteases, de-
tergents and denaturants, and physical forces [6]. The recent in-
crease in research activity in the area of amyloid-based materials
[7] is therefore not surprising. Amyloid brils have been used as
templates for metallic nanowires that could be used for molecular
electronics [6], have been proven to be efcient drug delivery ve-
hicles [8] and are useful as scaffolds for tissue engineering [9].In
the context of human diseases (see Section 2), their robustness
against chemicals generally impedes effective medical treatment of
amyloid-based diseases.
Despite the broad interest in this relatively novel class of ma-
terials, a detailed understanding of the relation between the mo-
lecular properties of amyloids and their material characteristics like
stiffness and mechanical strength has remained challenging. For
instance, the high
b
-sheet content as the main secondary structure
characteristic of amyloids is related to their structural stability [3].
But amyloids contain additional secondary structures and exhibit a
distinct polymorphism [10e12], seeming to weaken the exclusive
role of
b
-sheets as the determinant of amyloid stiffness and me-
chanical strength. In particular, the role of the amino acid side
chains remains to be elucidated in that context. The lag in our basic
understanding of the relation between amyloid structure and me-
chanical properties may be traced back to the limited number of
techniques that can provide information on the molecular in-
teractions underlying the assembly and material properties of
protein brils. The material properties of amyloids are based on
delicately interconnected effects occurring at a variety of length
scales, as illustrated in Fig. 1. Our article will highlight recent
applications of various techniques aimed at elucidating the mo-
lecular structures and mechanical properties of amyloid-brils and
derived materials.
Specically, this feature article describes recent investigations
into the structural and mechanical properties of amyloids at
different length scales. Nanoscopic level insights are obtained with
recently developed vibrational spectroscopic techniques. The vi-
brations of molecular groups within amyloid structures, and in
particular the amide I mode (with predominant C]O stretch
character), are sensitive reporters of the local environment of those
groups, as well as the protein structure. Hence, vibrational spec-
troscopies are very useful for quantifying secondary structure and
intermolecular interactions in amyloid systems. Moreover, surface-
specic vibrational spectroscopies like vibrational sum frequency
generation (vSFG) can provide information about amyloid forma-
tion at (lipid-) interfaces, which are known to catalyze the poly-
merization of amyloid precursors. The application of tip-enhanced
Raman spectroscopy (TERS) offers the possibility to spatially
resolve structural characteristics of amyloids in the nm regime.
Two-dimensional infrared spectroscopy (2D-IR) reveals amyloid
interactions down to the level of single amino acids, by selective
isotopic substitution.
At the mesoscopic level, we review results from biophysical
assays using transmission electron-, atomic force and uorescence
microscopy to measure single bril mechanical properties. Among
others, several different approaches to determine values for
bending rigidity, shear modulus, ultimate tensile strength and
extensibility are presented. Rheology measurements have been
applied at the microscopic scale to access the mechanical proper-
ties of networks of brils, which constitute promising materials like
viscoelastic gels.
2. Background
Many biological materials rely on brous networks of proteins
for their mechanical strength. Some well-known examples include
tissues such as skin, blood clots, and spider webs. Proteins are
normally folded in a specic geometry dictated by their primary
structure (amino acid sequence). Fibrils are then formed by su-
pramolecular assembly of protein building blocks, which are often
globular (as in the case of actin and microtubules) or rod-like (as in
the case of collagen and brin blood clots).
There is also an alternate path of bril formation: misfolded or
partially unfolded proteins tend to form amyloid brils. This
pathway was originally discovered in the context of several
neurodegenerative diseases (notably Alzheimers, Parkinsons,
Huntingtons, and Prion disease) and late onset diabetes [13,14].
These conformational diseases(or amyloidoses) are characterized
by the deposition of insoluble plaques of aggregated amyloid brils,
which can lead to cell death in specic organs. Surprisingly, several
organisms use amyloids to build natural, protective materials, such
as protective coatings of bacteria and spores and the silkmoth
eggshell [15]. It is increasingly apparent that the amyloid state is an
inherent characteristic of polypeptide molecules under denaturing
conditions, independent of the native structure or primary
sequence. In the context of food, this has been long known, for
instance in the cases of heat-denatured gelation of
b
-lactoglobulin
from milk and lysozyme from egg white [4].
Amyloid brils formed from structurally unrelated peptides and
proteins share a surprisingly similar structure. The change in sec-
ondary structure of the native protein to a
b
-sheet rich secondary
structure represents one of the main determinants of amyloid
formation. Amyloid brils show a characteristic X-ray diffraction
pattern, caused by a so-called cross-
b
core structure. The amyloid
core structure is composed of a stack of
b
-strands perpendicular to
Fig. 1. Hierarchical length scales relevant in amyloid research, starting from the level
of single molecules (left side: hIAPP, PDB 2L86) in the nanoscopic regime. Single or
several peptides and proteins build up secondary structure elements, which are
dominated in amyloids by
b
-sheets (PDB 2KIB). Mature amyloid brils have meso-
scopic dimensions and consist typically of two or more twisted strands, forming thin
and long brils. Networks of brils constitute micro- and macroscopic materials.
Vibrational spectroscopies focus on insights at the molecule and secondary structure
element level, atomic force and electron microscopy (AFM and EM) at the level of
single brils and (micro-) rheology on networks of amyloids.
M. Schleeger et al. / Polymer xxx (2013) 1e162
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
the bril axis separated by 4.8
A and with an intersheet spacing in
the order of 9e11
A[16,17]. The elongated stack is stabilized by a
dense network of hydrogen bonds. The sequence-specic side-
chains tend to affect the propensity to form brils [18]. A model of a
protobril from human amyloid polypeptide hIAPP is shown in
Fig. 2. A single hIAPP molecule forms two
b
-sheets, separated by a
loop region. The views along the entire bril axis in the model
reveal the dense stacking of four
b
-strands from two individual
IAPP molecules per layer. This in-register packing of parallel
b
-sheets is one of the prominent characteristics of amyloids. A view
perpendicular to the bril axis (Fig. 2D) illustrates the 4.8
A spacing
between the stacks of parallel
b
-sheets and their parallel orienta-
tion to the bril axis. Two or more protobrils usually twist
together to form unbranched, elongated mature amyloid brils
which look like twisted rope-like structures or at tapes, depend-
ing on the protein Ref. [12]. The high degree of structural order
within brils leads to very strong interactions between the proto-
brils within a mature amyloid bril (for instance 310 k
B
T/
m
m for
insulin Ref. [19]). The brils are typically about 10 nm in width
(with a range of 5e25 nm) and up to 10
m
m in length.
The unique molecular organization of amyloid brils endows
them with remarkable mechanical properties. Amyloid brils are
among the stiffest biological materials presently known, with a
Youngs modulus on the order of 3e20 GPa [4]. Moreover, amyloid
brils exhibit a high resistance to breakage. Their ultimate strength
was shown to be on the order of 0.6 GPa, comparable to the strength
of silk but also steel [2]. This large fracture strength seems surprising
since amyloid brils are held together by non-covalent interactions.
Recent experimental and theoretical work has demonstrated that
the remarkable rigidity of amyloid brils originates from the regular
network of intermolecular hydrogen bonds in the cross-
b
core [3].
Non-covalent interactions between the variable side-chains (hy-
drophobic or hydrogen-bonding) sometimes further stabilize the
bril and enhance the Youngs modulus [3].
From a materials science perspective, the amyloid pathway to
form protein brils has many advantages. The brils readily self-
Fig. 2. Structural model of an amyloid protobril from human islet amyloid polypeptide (hIAPP) based on the crystal structure from segments of the peptide. AeC View along the
bril axis. A. Two hIAPP molecules, each consisting of a hairpin and a steric zipper interface, tending to the brils axis. The space lling model B. emphasizes the tight steric zipper
interface of the two IAPP molecules. C. Stacking of
b
-sheet segments which are coiled up around the brils axis. D. View perpendicular to the bril axis, revealing the typical 4.8
A
spacing between layers of stacked
b
-sheets. The width of the bril is 64
A, the marked 125
A distance matches a quarter of a full helical turn. Reprinted with permission from Ref.
[17]. Copyright 2008 The Protein Society.
M. Schleeger et al. / Polymer xxx (2013) 1e16 3
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
assemble from natural proteins including edible proteins (such as
b
-lactoglobulin) or pharmaceutically relevant proteins (such as
insulin [19,20]), but they can also be formed from simple de novo
designed peptides [21,22]. Fibrils can also be formed from proteins
or peptides fused with functional proteins or peptides such as en-
zymes or cell adhesion motifs [9,23]. The mechanical properties are
remarkably insensitive to the protein sequence [4], and robust
under extreme conditions, such as high temperature, high con-
centrations of detergents and denaturants, as well as strong phys-
ical forces [6]. All of this makes them attractive candidates for
applications. Indeed, there has recently been a surge of activity
directed at designing materials from amyloids [7]. Amyloid brils
have been coated with metal [6] and conducting polymers [24] to
form conducting nanowires. Since amyloids are biocompatible,
their use as drug delivery vehicles [8] or applications in tissue en-
gineering [9] are also highly promising.
However, the ipside of the remarkable stability of amyloid
aggregates is that such stability is undesirable during the produc-
tion and storage of pharmaceutical and industrial peptides [25].
More importantly, the stability is also particularly undesirable in
the context of conformational diseases, such as Alzheimers and
Parkinsons disease and type II diabetes, where amyloid brils are
the cause of disease symptoms and accumulate in tissues because
they cannot be degraded. More than 20 proteins that are partially
unfolded, misfolded, or aggregated are known to give rise to am-
yloid brils in humans. There has been an intensive search for small
molecules and peptides that can inhibit the formation of brils
[14,26]. Originally, it was believed that the brils were the major
toxic species (the amyloid hypothesis). However, recently it has
become apparent that oligomeric species may be even more toxic
[27]. Thus, small molecule or peptide-based drugs should not only
inhibit brillization, but also divert polypeptide assembly down
alternative, less harmful assembly pathways [28].
Here we focus mainly (but not exclusively) on two exemplary
amyloids, the disease-related human amyloid polypeptide (hIAPP)
or amylin and the primarily food-related
b
-lactoglobulin (
b
-lg).
hIAPP forms extracellular deposits in the pancreas of type 2 dia-
betes mellitus patients [29]. hIAPP is in vivo cosecreted with insulin
by the
b
-islet cells of the pancreas. Amyloid brils and/or oligo-
meric forms of hIAPP are thought to induce cell death of islet cells,
eventually destroying insulin production [29e32].Forin vitro
studies, hIAPP is commonly prepared by solid-state synthesis, since
it is a short peptide of 37 amino acid residues. In contrast,
b
-lg is a
globular protein from milk that is of great interest for industrial
(food) applications [33]. Amyloid brils can be readily formed by
heat-induced denaturation at low pH [34e37]. Both hIAPP and
b
-lg
are examples for well-studied systems in terms of kinetics and
thermodynamics of formation (
b
-lg [36,38]; hIAPP [39]) and bril
structure (
b
-lg [34,37]; hIAPP [17,40e42]).
Recently, plant polyphenols have drawn attention as a possible
candidate to control amyloid formation and stability. For instance,
polyphenols are promising as a drug for prevention or treatment of
Alzheimers disease [4,43,44], and affect the quality of food-related
materials [45]. They are ubiquitous compounds in plant-derived
food and beverages (e.g. green tea) and have recently been
shown to be potential inhibitors of amyloid brillization [43,44,46].
Polyphenols are thought to have many more benecial properties
based on their antimicrobial properties and their antioxidative
capacities, including therapeutic potential for cancer [47] and car-
diovascular diseases [48]. The strong interactions between poly-
phenols and amyloid proteins could potentially be used to combat
amyloidosis, but proteins could also serve as delivery vehicles of
polyphenols in the form of polyphenol-rich food or of drug delivery
agents. Polyphenols are thought to interact with amyloids and
amyloid precursor proteins via aromatic residues [49e51].
3. The nanoscopic level espectroscopic insights into the
molecular structure of amyloids
Vibrational spectroscopy is being widely applied for structural
analysis of biopolymers. In this article, we will refrain from a
description of classical Raman and FT-IR spectroscopy, despite the
signicant importance of these techniques in current amyloid
research [52,53]. We will focus on new techniqueswith only a
very brief and fragmentary description of the techniques them-
selves, rather emphasizing the new insights into amyloid brils
provided by such techniques with the use of a few signicant
examples.
The development of two relatively new vibrational techniques,
coherent two-dimensional infrared spectroscopy (2DIR) and
vibrational sum frequency generation spectroscopy (vSFG) have
depended strongly on the development and common availability of
infrared laser systems with ultra-short pulse duration. Their
application to biopolymers is still at its beginning, not only
providing the general structural sensitivity of vibrational spec-
troscopy but further offering an intrinsic possibility to perform
ultrafast time-resolved experiments down to femtoseconds, the
time domain of formation and breaking of chemical bonds [54].
Tip-enhanced Raman scattering (TERS) provides another route to-
wards the investigation of nanoscale structural variations, utilizing
plasmonic particles for sensitivity enhancement and spatial reso-
lution improvement [55,56]. TERS enables even the investigation of
the polymorphism of single amyloid brils [57]. In the following,
recent results obtained with these three vibrational spectroscopic
approaches are highlighted.
3.1. Vibrational sum frequency generation spectroscopy (vSFG)
Vibrational sum frequency generation spectroscopy (vSFG)
emerged after the development of ultrafast infrared lasers as a
technique mostly applied under ultra-high vacuum (UHV) condi-
tions to characterize the vibrational spectra of molecules adsorbed
onto surfaces. Nowadays vSFG is routinely applied to complex
biologically relevant samples, with its rather unique surface spec-
icity, combined with high sensitivity and time resolution
[54,58,59]. The main strength of vSFG is that it allows one to record
the vibrational spectrum of specically the outermost monolayer of
a bulk material, with high sensitivity.
In a typical vSFG experiment, ultrashort amplied laser light
pulses (e.g. at a wavelength of 800 nm) are split to allow for the
subsequent generation of spectrally narrow visible and broadband
mid-infrared laser pulses. The two coherent pulses are spatially and
temporally overlapped onto the sample surface and interact with
the interface by generating light of the sum frequencyof the visible
and mid-infrared light. This process is enhanced if the infrared light
is in resonance with vibrational modes on the surface. Using
spectrally resolved detection, the frequencies of surface vibrations
are readily extracted from the vSFG spectra by subtracting the
frequency of the narrowband visible pulse. The resulting vSFG
spectrum is usually normalized to the non-resonant vSFG signal of
a reference material or the intensity spectrum of the broadband IR
laser pulse (for a detailed review we refer to [54]). Heterodyne
detection may be applied to obtain not only intensity, but also
amplitude and phase information of the spectra, an approach
which is helpful in particular if vibrational modes strongly overlap.
In addition to its sensitivity, the specicity of vSFG to non-
isotropic samples offers the remarkable possibility to measure
structurally sensitive vibrational spectra exclusively at interfaces.
Isotropic bulk media show no SFG spectra due to the selection rules
of the underlying second-order nonlinear optical process. The
second-order polarizability, the source term of the SFG light, is zero
M. Schleeger et al. / Polymer xxx (2013) 1e164
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
for any medium with inversion symmetry. At interfaces, the sym-
metry of the centrosymmetric bulk phase is broken, allowing for
measurements of vibrational spectra specically at the interface.
The sensitivity of vSFG to surface monolayers enables for
instance the investigation of the very same amyloid samples which
are suitable for atomic force microscopy (AFM) as well. Therefore a
molecular method can be directly aligned to a mesoscopic method.
This approach can be applied to consecutively analyze secondary
structure and e.g. stiffness parameters of amyloids grown at
different conditions. The spatial resolution of a typical SFG setup is
restricted by the size of the two overlapping laser pulses. Typically,
the SFG signal is averaged over an area somewhat less than one
square millimeter. The application of AFM and SFG together allows
therefore the combination of a high spatial resolution technique
with a surface sensitive label-free vibrational spectroscopy.
The combination of vSFG and tapping mode AFM was used to
characterize different morphology forms of amyloids grown from
b
-lactoglobulin [60]. It was shown that
b
-lg forms small, worm-like
amyloids when grown under high concentration (7.5% w/v) con-
ditions and long, straight brils at lower protein concentration
(3% w/v) [60]. The stiffness of the two morphologically different
amyloid forms differs strongly (see below), even though they
display the same bril diameter. vSFG was applied to reveal the
molecular origin of this surprising behavior [60].InFig. 3a, vSFG
spectra in the amide I spectral region of amyloid brils obtained
from different growth conditions are presented. The position of the
vSFG band enables the determination of the secondary structure
distribution of the samples. Fig. 3b depicts the resulting secondary
structure analysis along with two representative AFM images of the
amyloid brils. It could be shown, that the small, worm-like amy-
loids of
b
-lg grown under high concentration (7.5% w/v) conditions,
exhibit a high content of
a
-helical/unordered structures. For the
long brils, formed at lower protein concentration (3% w/v), a
strong content of
b
-sheets could be detected. The bril persistence
length (or bending rigidity) apparently decreases for an increasing
content of
a
-helical/unordered secondary structure. Whereas for
straight brils with a strong
b
-sheet content a persistence length of
3820 160 nm was found, the worm-like brils had a persistence
length of only 92 7nm[60]. While the inuence of secondary
structure on the stiffness of amyloid brils could be unambiguously
demonstrated, the precise spatial distribution of secondary struc-
ture remains unclear. Structure sensitive methods with inherent
high spatial resolution like tip enhanced Raman scattering (TERS,
see also below) may in the future address the question how brils
can be built up with such remarkable differences in their secondary
structure composition, and yet be of the same thickness.
In vivo, the growth of amyloid brils, as well as the cytotoxic
effect of brils or their oligomeric precursors, is thought to take
place on the plasma membrane of cells [30,31]. The possibility of
vSFG to exclusively detect signals from model membranes like the
airewater interface of lipid monolayers provides the opportunity to
measure structural properties of amyloid brils under more bio-
logically relevant conditions than with conventional vibrational
spectroscopy techniques. In their pioneering work on amyloids
forming at lipid monolayer interfaces the Yan group employed
surface-selective SFG to monitor structural changes of islet amyloid
polypeptide during brillation [61,62]. Subsequent they reported
on the orientation of human IAPP amyloids, based on polarization
dependent SFG studies in combination with ab initio calculations
[63]. hIAPP is involved in the death of insulin producing islet cells of
Langerhans [29]. With SFG, the misfolding of hIAPP from an
a
-he-
lical or unordered form to a
b
-sheet rich amyloid structure was
kinetically resolved under the catalyzing action of negatively
charged phospholipids [61].
As mentioned in the background section, polyphenols like
()-epigallocatechin gallate (EGCG) are promising inhibitors of
amyloid formation in the context of various amyloid-related dis-
eases. Applying the experimental conditions employed by the Yan
group [61], we tested the inhibitory action of EGCG using the sur-
face specicity of SFG at the phospholipid interface. Fig. 4 sum-
marizes the experiments following the changes in secondary
structure of human IAPP in the amide I spectral region. Without
inhibitor, a structural transition from an
a
-helical or unordered
secondary structure to a
b
-sheet rich structure was observed by a
shift of the amide I band from 1652 cm
1
e1672 cm
1
[64]. This
secondary structure transition was interpreted as the formation of
amyloid brils (see Fig. 4a), as veried by AFM measurements on
the same samples transferred to a solid substrate. In the presence of
the inhibitor EGCG the structural transition is less pronounced, but
still obvious [64]. This observation is in strong contrast with the
situation in bulk. A standard biophysical technique for detecting
amyloids in solution is the Thioavin T (ThT) uorescence assay.
Upon binding of Thioavin T to
b
-sheet rich amyloids, an enhanced
and red-shifted uorescence is observed, as shown in Fig. 4c for the
Fig. 3. a. SFG spectra of
b
-lactoglobulin amyloids on mica in the amide I spectral region. Amyloid brils were grown for different monomer concentrations, as indicated in % w/v. The
spectra allow for a secondary structure analysis of the samples by tting the vibrational SFG band with two components, corresponding to the contribution of
a
-helices or un-
ordered structures (around 1652 cm
1
) and
b
-sheets (around 1629 cm
1
). b. Quantitative analysis of the SFG spectra shows, that amyloid brils grown under higher concentration
conditions (6 and 7.5% w/v) exhibit higher
a
-helical or unordered structural content, and a less pronounced
b
-sheet structural content. AFM imaging of the
b
-lactoglobulin amyloid
samples on mica shows a morphology change from long, straight brils at low concentration to worm-like ones at higher concentrations (the white bar corresponds to a length of
100 nm). This transition is evidently accompanied by a change in secondary structure. Figures are based with permission on [60]. Copyright 2011 American Chemical Society (Minor
corrections of the peak fractions in b. were performed, which do not affect the general trend of the original gure).
M. Schleeger et al. / Polymer xxx (2013) 1e16 5
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
formation of IAPP amyloid brils in bulk solution (see trace marked
by blue ). In the presence of the inhibitor EGCG, no increase in
uorescence could be detected (red þ). The combined experiments
at the interface and in bulk solution lead to the conclusion, that
whereas EGCG is a potent inhibitor in bulk, its inhibitive effect at
the membrane interface is strongly reduced. The membrane
interface represents the biologically relevant location of the cyto-
toxicity of hIAPP. Even more pronounced is the difference between
bulk and lipid interface for the disaggregation of pre-formed am-
yloids by EGCG. The ability of EGCG to dissolve preformed aggre-
gates in bulk solution could be veried again by a ThT assay as
depicted in Fig. 5a (blue ). The uorescence rapidly decreases after
addition of the inhibitor, indicating a fast disaggregation. SFG is
able to monitor the effect of EGCG exclusively at the lipid interface
(red
D
). Even after incubation times up to 30 h (Fig. 5a) or after
addition of a 1000-fold molar excess of EGCG compared to the
hIAPP concentration (Fig. 5b), the
b
-sheet content of the sample
stays constant. AFM conrmed the presence of brils after treat-
ment with EGCG (see inset in Fig. 5b). In conclusion, at the lipid
interface the inhibitor EGCG is not able to dissolve amyloid brils at
all. The molecular origin of these ndings is the subject of ongoing
research, but the results clearly highlight the importance of struc-
tural information gained selectively at interfaces.
vSFG offers many possibilities for future studies that can address
more complex, biologically relevant problems. For instance, the
inherent fast time-resolution of vSFG may be exploited to reveal
short lived oligomeric states during amyloid formation or disag-
gregation. While applications of whole cells in pharmaceutical
studies of amyloids are well documented [46,65], recently vSFG was
applied on brillar proteins of the extracellular matrix of living cells
as well. In future experiments, the complexity of in vivo tests is
therefore not beyond the means of vSFG [66].
3.2. Tip-enhanced Raman scattering (TERS)
A combination of the excellent spatial resolution of AFM and the
structural sensitivity of vibrational spectroscopy in a single setup is
realized by tip enhanced Raman spectroscopy (TERS). TERS is
becoming an increasingly off-the-shelve tool for label-free struc-
tural investigations of nanoscale phenomena at surfaces. Recent
results have demonstrated a lateral resolution down to the sub-
Fig. 4. a. Fibrillation of human islet amyloid polypeptide (hIAPP) followed by SFG at a phospholipid monolayer-covered water/air interface. At early times, the amide I band is
centered at 1652 cm
1
, corresponding to a mainly
a
-helical or unordered secondary structure. For later times a shift of the amide I band to 1672 cm
1
is observed, denoting a
b
-sheet
rich structure, characteristic for amyloids. b. hIAPP brillation followed in the presence of the inhibitor EGCG. The formation of
b
-sheets is less pronounced, but still evident. Fitting
of three bands (as exemplied for the earliest time by the broken red lines in a and b) to the SFG data allows a quantication of the
b
-sheet contribution to the spectrum. c. A
comparison of the
b
-sheet spectral contribution, revealed by SFG for the lipid interface situation and the uorescence intensity of a Thioavin T assay for bulk solutions. Both
situations are plotted in the absence (blue Bfor SFG and blue for ThT assay) as well as in the presence of the inhibitor EGCG (red
D
for SFG and red þfor ThT assay). The effect of
the inhibitor EGCG is strongly diminished at the interface, compared to the bulk situation. Adapted with permission from Ref. [64]. Copyright 2012 American Chemical Society.
Fig. 5. a. Disaggregation of preformed hIAPP brils, followed by the amount of
b
-sheet secondary structure using SFG at the phospholipid interface (red
D
) and the uorescence
intensity of a ThTassay in bulk solution (blue ). At time zero, EGCG is added, which leads to a rapid bril disaggregation in bulk, whereas no disaggregation takes place at the lipid
interface. b. Even a 1000 fold molar excess (1000
m
M) EGCG does not lead to a signicant decrease in the
b
-sheet content. The presence of brils is veried by AFM imaging
(see inset). Based on [64] with permission. Copyright 2012 American Chemical Society.
M. Schleeger et al. / Polymer xxx (2013) 1e166
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
nanometer range. Clearly, this renders the method a valuable tool
for the direct investigation of single polymer strands.
One of the key components of a TERS setup is a plasmonic tip,
like a sharp edge of gold or an isolated gold or silver nanoparticle.
The external, monochromatic electromagnetic laser eld oscillates
at a given frequency, which drives the electrons in the plasmonic
structure to perform a periodic motion. The resulting oscillating
dipole may increase the strength of the electromagnetic eld in the
immediate vicinity of the plasmonic structure by many orders of
magnitude. Raman scattering can take place in molecules under the
plasmonic structure. Because both the incident and scattered
Raman elds are amplied by the plasmonic effect, the Raman eld
scales linearly with the incident eld, and because the detected
Raman intensity scales with square of the Raman eld, the Raman
intensity enhancement scales with the fourth power of the local
eld enhancement. As a result, extremely low concentrations down
to single molecules can be investigated. The size of the nano-
particles determines the spatial extent of the eld-enhancing re-
gion and as a rule of thumb the lateral resolution in TERS is about
half of the radius of the nanoparticle. In addition further
enhancement mechanisms due to interactions between the metal
particle and the investigated molecule can occur (so called chem-
ical enhancement). Interestingly the depth resolution is even
higher than the lateral resolution. According to simulations the
electromagnetic eld decays very fast and the z-resolution is in the
range of 1e2 nm. Also here, a direct interaction also increases this
depth resolution.
These properties are ideally suited to investigate amyloids with
a thickness of a few nm and length in the micrometer scale. The
correlation of a topographic image of a bril with the spectral map
could possibly reveal the surface structure of complete brils. The
polymorphism of single amyloid brils gets therefore directly
accessible.
An inherent characteristic of TERS is the ability to analyze the
topographic and chemical structure of a specimen in one experi-
ment. Fig. 6 shows the results of such an experiment on a single
segregated hIAPP bril. In Fig. 6a selected TERS spectra show the
typical Raman signals of a protein. The detection sites on the bril
are marked in the topography image in Fig. 6b.
The spectral range in Raman spectroscopy that gives infor-
mation on the secondary structure of a specimen is the amide I
region between 1640 and 1680 cm
1
. The bands highlighted in red
(1660e1680 cm
1
)inFig. 6a are characteristic for
b
-sheet struc-
tures whereas the bands highlighted in green between 1640 and
1655 cm
1
are marker bands for
a
-helix structures. The detection of
both conformations on hIAPP brils indicates that the surface is
composed of a heterogenous secondary structure. This is different
from the core structure, which is known to be pure
b
-sheet. Such
observations have been made previously on insulin [67] brils and
give rise to the assumption that such heterogeneity is a common
feature of amyloid brils. The role of such conformation mixtures is
still under investigation.
The diversity of the spectra in Fig. 6a recorded on positions at
nanometers distance, demonstrates the ability of TERS to differ-
entiate protein structures on the nanometer scale. Besides a char-
acterization of secondary structures on brils this specialized
Raman spectroscopic technique provides information about the
amino acid distribution on the bril, also with nanometer resolu-
tion [57]. It is evident that this approach can be of great assistance
for understanding the brillation process trigger and/or mecha-
nism on the nanometer scale.
In addition to having the ability to address amyloid bril poly-
morphism, TERS holds the possibility to elucidate the role of single
amino acids in determining bril stability and mechanical proper-
ties. This is an important issue, as the side chains of amino acids
have been shown to contribute to the stability of the
b
-cross core of
amyloids, affecting the elastic modulus of brils [3].
3.3. Coherent two-dimensional infrared spectroscopy (2DIR)
2DIR has become a powerful tool for the structural investigation
of proteins. By addressing localized vibrational transitions using a
specic sequence of infrared laser pulses, details of the molecular
response can be obtained, which can be related to structure and
structural evolution. The structural information contained in
respective linear spectra is essentially mapped onto two di-
mensions. The essence of 2DIR spectroscopy is that one specic
vibrational mode is selectively excited, and the effect of that exci-
tation is monitored on both that same mode, and on other vibra-
tional modes. From the study of the mode itself, the distribution of
vibrational frequencies can be obtained, as well as the timescale on
which these frequencies change, as a result of conformational
uctuations, for instance. If the excitation has an effect on a
different mode, this means that the two molecular groups primarily
responsible for the modes are physically close to one another. The
strength and nature of the resulting coupling can be related to the
local structure, and structural rearrangement (basic details can be
obtained from the textbook [68]).
2DIR spectra can be recorded either in the frequency domain or
in the time domain, which both results in basically identical spectra
Fig. 6. a. Four raw TERS spectra obtained from a single IAPP bril (generated at pH 7).
Spectra 1 and 2 were recorded at a relative lateral distance of 1.5 nm at the position
marked by the upper * in panel b. and show characteristic bands of
a
-helix secondary
structures (highlighted in green). Spectra 3 and 4 were recorded at a relative lateral
distance of 1.5 nm at the position marked by the lower * in panel b. and show in
contrast characteristic bands of
b
-sheet structures (highlighted in red). b. Typical AFM
topography of IAPP brils (generated at pH 7). TERS spectra were acquired on the top
bril at two areas (marked with *), each with two positions separated by 1.5 nm,
respectively.
M. Schleeger et al. / Polymer xxx (2013) 1e16 7
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
[69]. Experiments in the frequency domain are realized by the
pump-probe technique. A scanning, spectrally narrow IR pump
pulse excites vibrational transitions and a spectrally broad probe
pulse identies the exited vibrations after a xed time from the
pump pulse. The 2D spectrum is constructed by plotting the
measured probe spectra for each applied pump frequency.
For 2DIR experiments in the time domain, also named echo
spectroscopy, three sequential IR pulses are applied to generate the
echo signal. The latter is recorded with a varying delay between the
two rst (pump) pulses and a xed delay between the second
pump and last (probe) pulses. Heterodyne detection is applied to
extract both amplitude and phase information. The construction of
a 2D spectrum involves a Fourier transformation: if the probe
spectra are plotted against a full set of the time delays between the
two pump pulses, a damped oscillation in the recorded intensity for
each probe frequency is yielded. This is analogous to multidimen-
sional FT-NMR correlation spectroscopy, where a free induction
decay (FID) is measured for a sequence of radio frequency pulses. By
means of Fourier transformation of the time-domain signal a pump
spectrum in the frequency domain is obtained for each measured
probe frequency.
A typical 2DIR spectrum contains valuable information, addi-
tionally to linear spectroscopy. Diagonal elements refer to peaks
near the diagonal of the 2D spectrum. They originate from excited
vibrational states, leading to higher order absorption (e.g. from the
rst to the second excited state), resulting in positive peaks and
emission processes resulting in negative bands slightly shifted with
respect to the positive ones. Due to fast vibrational relaxation
processes occurring between pump and probe pulses, additional
negative bands may arise due to absorption of the probe beam from
vibrational ground states. The main characteristics of 2DIR spec-
troscopy are the cross-peaks, which arise from coupled vibrational
modes. Knowledge on vibrational coupling in terms of distance and
angular dependence can be transferred to structural information. In
complex systems like biomolecules often models based on molec-
ular dynamics simulations are implemented to extract structural
information from vibrational coupling [70]. 2D band shapes include
additional structural information and are inuenced by environ-
ment effects such as hydrogen bonding or solvent effects [71].
Additionally to the structural sensitivity of 2DIR, transient 2DIR
permits probing of the kinetics of structural transitions with sub-
picosecond time resolution, for instance in case of chemical re-
actions [72,73].
A widely used application of 2DIR is the secondary structure
analysis of proteins. 2DIR is regarded as being more accurate than
1D FT-IR because of the increased information content achieved by
having information on both excitation and detection axes, and the
above-mentioned high time-resolution [74,75]. Isotopic labeling of
specic amino acids can be employed to reveal the site-specic,
local secondary structure of proteins [71]. Additionally, challenges
like heterogeneity within a protein, proteineprotein interactions or
ligand binding can be addressed [70]. Although secondary structure
analysis is quite accessible with 2DIR, investigations of the tertiary
structure of proteins are performed in combination with molecular
dynamics calculations to distinguish between different tertiary
structure types. A practical drawback of the technique is the
requirement of high sample concentrations, when compared to
traditional 1D FT-IR or monolayer sensitive vSFG.
A recent example of a detailed structural analysis of amyloids
from hIAPP was reported by Wang et al. [73]. The authors com-
bined experimental information from 2DIR spectroscopy with
calculations of 2DIR spectra and MD simulations of hIAPP, allowing
for conclusions on the structure of hIAPP brils beyond secondary
structure analysis. Based on the fact that the diagonal line widths
of the amide I peak of single amino acids differ for residues from a
b
-sheet and C-terminal or turn-structures, the authors deduced
the number of folds of
b
-structures within the peptide. Experi-
mentally, they employed seven samples with (diluted) site-specic
13
C¼
18
O labeled amino acids. It was therefore possible to test the
secondary structure of hIAPP specically on seven distinct loca-
tions along the peptide chain. The diagonal amide I peaks of the
labeled amino acids showed a pronounced down shift compared
to the unlabeled amino acids and could be investigated mostly
undisturbed by the unlabeled part of the protein. Plotting the line
widths of labeled amide I peaks against the residue number, Wang
et al. observed a characteristic W-shapepattern. This pattern
was assigned to the presence of two stable
b
-sheet regions con-
nected by a turn-structure within a single amyloid-forming hIAPP
monomer. The pattern could be compared to calculated 2DIR
spectra based on previously proposed structural models of hIAPP
amyloids relying on solid state NMR and electron microscopy
studies [76,77].
The Zanni group, involved as well in the above mentioned study,
reported on the residue-specic kinetics of hIAPP aggregation [78],
employing isotopic labeled amino acids on six different positions
within the peptide. Based on the time-resolved site-specic for-
mation of
b
-sheets it was possible to conclude on requirements for
a model of the hIAPP aggregation pathway. Such a pathway was
found to be more complex than a simple pathway requiring a
random-coil intermediate, which undergoes a concerted transition
to a ber-like nucleus and subsequent elongation by the addition of
monomers [78].
2DIR contributed further to the elucidation of structural tran-
sitions during the formation of amyloids [78e80], as well as the
effect of inhibitors on amyloid formation [81]. Meng et al. reported
on the effect of an inhibitor based on a sulfated triphenyl methane
derivate on the formation of amyloids from hIAPP. 2DIR revealed an
inuence of the inhibitor on the
b
-sheet content of the samples, as
shown in Fig. 7.InFig. 7a, the 2DIR spectrum of the pure hIAPP
brils shows two diagonal out-of-phase features (doublets). A
strong one at the pump frequency of 1619 cm
1
, corresponding to a
predominant
b
-sheet secondary structure [71], as well as a weaker,
broad feature at 1646 cm
1
, consistent with an unordered structure
[81]. In the presence of inhibitor, a strong decrease in the peak
intensities attributed to the
b
-sheet structure is evident (Fig. 7b). A
smaller effect is an accompanying frequency shift of 3 cm
1
to
higher frequencies compared with pure hIAPP samples, which the
authors attributed to a smaller size of the
b
-sheets.
The interaction of a peptide-based inhibitor of hIAPP bril for-
mation was investigated by Middleton et al. [28]. The authors
applied isotopic labeling for a residue-specic 2DIR analysis of the
inhibitory effect of rat IAPP (rIAPP) on hIAPP brillation. For this
purpose, they incubated the inhibitor and the amyloid peptide for a
short (8 h) and a long time (24 h) after mixing. rIAPP was known
not to form brils under any conditions, but represents a moderate
inhibitor of hIAPP bril formation (for in vitro tests a strong inhi-
bition of bril growth requires approximately a tenfold molar
excess of rIAPP [82]). A rst nding was the identication of the
inhibitor binding side by the fact that the N-terminal
b
-sheet of
hIAPP is not formed within 8 h in the presence of rIAPP (Fig. 8a).
This result explained the structural basis of the slower amyloid
formation in the presence of the inhibitor. After long-time incu-
bation of inhibitor and hIAPP, bril formation was not inhibited.
Surprisingly it was found, that rIAPP and hIAPP form mixed amy-
loidogenic parallel
b
-sheets 24 h after mixing (Fig. 8b). The struc-
tural sensitive 2DIR technique could reveal therefore a potential
additional toxicity of an agent, which was previously thought to
combat amyloid formation. Structure sensitive 2DIR proved
therefore as a very fruitful technique for the investigation of amy-
loid formation and inhibitor interaction, in particular because
M. Schleeger et al. / Polymer xxx (2013) 1e168
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
standard uorescence techniques (Thioavin T assay) suggested
that the system already equilibrated after 8 h [28], contrary to 2DIR.
More complex experiments like 3DIR approaches or additional
pulse sequences [68] may in future further strengthen the appli-
cation of multidimensional IR correlation spectroscopy to struc-
tural characterization of biopolymers.
4. Mesoscopic properties of single amyloid brils
Proteins and peptides form amyloid brils by self-assembling
into protobrils, which have a
b
-sheet core. Usually, between 2
and 4 protobrils subsequently twist together to form a mature
bril [83e85]. Amyloid brils tend to be highly polymorphic,
varying in length and in the number of protobrils. The
b
-sheet
core makes amyloid brils rigid with a Youngs modulus similar to
that of silk and strong with an ultimate tensile strength comparable
to that of steel [2]. Because of their outstanding mechanical prop-
erties, amyloid brils are promising structures for applications in
biomaterials and food technology. However, the mechanical prop-
erties of amyloids are still not fully understood. There are several
theoretical models which assign the large rigidity and strength of
amyloid brils to their
b
-sheet core [3,86]. However, amyloid brils
are polymorphic, and vibrational spectroscopy data indicate that
the
b
-sheet content can be much less than 100%, as shown in Fig. 3.
At least in part, this is a consequence of the presence of side-chains,
whose effect on the mechanical properties of brils is unclear. In
addition, amyloid brils may have imperfect
b
-sheet cores with
interruptions along the long axis. Here we will review recent bio-
physical studies of the mechanical properties of single brils, and
indicate some promising techniques that can be used in future to
better understand the mechanics of amyloid brils and their origin.
4.1. Bending rigidity and shear modulus
A key parameter characterizing the conformation of a
biopolymer is the bending rigidity. The bending rigidity is often
expressed by the length scale beyond which the bril shows sig-
nicant curvature due to thermal forces, quantied by the persis-
tence length P[87,88]. The bending rigidity can be measured by a
variety of methods, which are either based on active deformation of
the brils, or on analysis of the spontaneous, thermal bending
uctuations of the brils.
Active deformation of brils can be achieved by AFM, as
demonstrated in a recent study on insulin amyloid brils composed
of two protobrils [2,3]. Essentially, the experiment was a micro-
scopic equivalent of a traditional three-point bending test: brils
were deposited on a silicon substrate with nanoscale grooves. With
an AFM probe, a controlled load was applied on brils suspended
over a groove, while monitoring the deection of the cantilever that
acts as the force sensor. The Youngs modulus was determined from
the linear (small-force) part of the deection curve, giving an
average Youngs modulus of E¼3.3 0.4 GPa and shear modulus of
G¼0.28 0.2 GPa. This Youngs modulus corresponds to a
persistence length of 42 30
m
m, which is much larger than the
typical bril contour length of 3e6
m
m. An advantage of this
technique is that the measurements are performed on brils in
their natural, hydrated state. However, it is only possible to mea-
sure on the length scale of the tip radius, which is small compared
to the brils. Also, the experiment is technically very challenging
because of the small diameter of the amyloid brils [83].
Technically, it is far easier to measure the bending rigidity based
on the spontaneous, thermal uctuations of brils. The basic idea is
to measure the shape of uctuating brils, either taking snapshots
of a large ensemble of brils at a given moment in time, or taking
time-lapse movies of a single bril. Until now, most studies of
amyloid brils used the rst method, analyzing the shape of a large
ensemble of brils immobilized on a surface. Usually, amyloid -
brils are imaged by AFM [60,89e91], which requires bril deposi-
tion on a mica or glass surface and drying. Alternatively, cryo-
transmission electron microscopy (cryo-TEM) can be used, which
has the benet that brils are preserved (snap-frozen) in a hydrated
state [92]. The persistence length Pcan be calculated by measuring
Fig. 8. Model structures of the co-aggregated hIAPP (grey) and rIAPP (pink) a. 8 h and
b. 24 h after mixing of the peptides. One site-specic isotopic labeling side of hIAPP is
depicted as blue spheres. In a., rIAPP prevents initially the N-terminal formation of
b
-sheets of hIAPP. In b., mixed
b
-sheets from hIAPP and rIAPP are formed. Reused with
permission from Ref. [28]. Copyright 2012 Macmillan Publishers.
Fig. 7. 2DIR spectra in the amide I spectral region of hIAPP after formation of amyloids
in the (a) absence or (b) presence of an inhibitor based on a sulfonated triphenylmethyl
derivate. The same growth conditions were applied. Reprinted with permission from
Ref. [81].
M. Schleeger et al. / Polymer xxx (2013) 1e16 9
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
for each bril the contour length, C, and the end-to-end distance, E
(Fig. 9). According to the worm-like chain model, the mean-squared
end-to-end distance on a two-dimensional surface as a function of
Cdepends on Pas in the following equation:
E
2
2D
¼4PC12P=C1e
C=2P
:
This equation assumes that the brils interact weakly with the
surface and can relax to a two-dimensional equilibrium confor-
mation [87,89,91]. However, if the interaction between the bril
and the surface is much stronger than the thermal energy, the -
brils will be trapped by the surface, leading to more condensed
bril conformations. In this case, the mean-squared end-to-end
distance of the bril amounts to [89]:
E
2
3D
¼4=3PC1P=C1e
C=P
:
An alternative method of determining Pis by analyzing the
correlation of bond angles along the contour:
P¼hli=ð1hcos
4
;
where <l>is the average segment length and
4
is the angle be-
tween segments (Fig. 9)[93].
One of the best-studied types of amyloid brils are those formed
from the model protein
b
-lactoglobulin (
b
-lg). Depending on the
self-assembly conditions, these brils have widely varying persis-
tence lengths and mostly fall into one of two classes: straight brils
or worm-like brils (Fig. 10a, b and Fig. 3b). Straight brils generally
have a persistence length that is comparable to the contour length,
ranging from about 600 nm to several
m
m[37,60,90,91,93e96].
Worm-like brils are much more exible, with a persistence length
of only 10e90 nm (Table 1)[37,60,89,91,93,94,96]. A third type of
morphology, rod-like brils with a persistence length of 135 nm,
was formed from
b
-2-microglobulin (Fig. 10c) [91]. It has to be
noted that amyloid brils are polymorphic; even under a given set
of conditions, brils have different persistence lengths. The
numbers given in Table 1 are average values.
In contrast to nanomechanical manipulation assays, the uctu-
ation analysis approach gives a global measure of the persistence
length of a bril, because uctuations are evaluated on a larger
length scale. However, bril imaging by AFM requires deposition on
a surface and drying, which can potentially lead to artifacts.
One way to overcome this problem is to measure the shape of
freely uctuating brils in solution by uorescence microscopy
(Fig. 10d) [97,98]. Because of the low resolution of optical micro-
scopes in the z-direction, brils are typically conned in a quasi-2D
geometry by sandwiching them between two glass cover slips.
Rather than extracting the persistence length from the static shape
of an ensemble of brils, the persistence length is now calculated by
tracking the bril shape uctuations over time. These experiments
have been performed for yeast prion brils labeled with a
uorescent dye [98]. The persistence length was found to be
3.6 1.1 or 7. 0 2.4
m
m, depending on the assembly conditions. In
addition to the bending rigidity, also the bending dynamics can be
determined from time-lapse movies, as demonstrated for actin
laments and microtubules [88].
A second way to overcome problems associated with drying or
surface immobilization is to perform light scattering of dilute bril
suspensions, again using the worm-like chain model to interpret
the data. The persistence length of
b
-lg amyloid brils formed at
varying ionic strengths was determined by a combination of light
scattering (LS) and small-angle neutron scattering (SANS) [94]. The
persistence length decreased with increasing ionic strength from
600 to 38 nm. These values are in reasonable agreement with the
persistence lengths obtained with imaging techniques (Table 1).
The persistence length of
b
-lg brils also has been estimated
using an adjusted random contact model, based on measurements of
the storage (or elastic) modulus G
0
of bril gels by rheology [37]. The
critical percolation concentration was determined by measuring G
0
as a function of protein concentration. It was assumed that for a
brillar system the percolation mass fraction is described by the
volume of the bril, the number of contacts per rod and the excluded
volume of charged semiexible brils.The volume of the bril can be
determined by the effective diameter, contour length and persis-
tence length. The estimated persistence length was 1.6 0.4
m
m for
b
-lg brils formed at pH ¼2 and 80
C(Table 1, 7th entry).
4.2. Ultimate tensile strength and fracture behavior
The ultimate strength of single brils has been measured by
nanomechanical manipulation with AFM. For insulin brils, the
ultimate strength was measured by actively bending brils sus-
pended over a groove with an AFM tip. The mean ultimate strength
for brils composed of two protobrils was 0.6 0.4 GPa. Strik-
ingly, this is in the same order of magnitude as the strength of steel
(0.6e1.8 GPa) and silk (1.0e1.5 GPa) [2,3]. The corresponding
breakage force was estimated to be in the range of 300e500 pN [2].
Microscopic insight into the molecular mechanisms that deter-
mine the strength of amyloid brils can be obtained by force
spectroscopy, where peptide strands are pulled from brils
immobilized on a surface with a small AFM tip. Such experiments
were reported for brils formed from A
b
(1e40) or A
b
(25e35)
peptides [99]. Strands of more than 100 nm in length could be
pulled out of the brils and stretched. The force-extension behavior
that was observed was tted with a worm-like chain model, from
which a persistence length of about 0.4 nm was calculated. Based
on these results, the authors concluded that the strands were
b
-
sheets that were unzipped from the brils. The unzipping process
was fully reversible. Fibrils formed from A
b
(1e42) peptide showed
a lower unzipping force (w23 pN) than brils formed from A
b
(1e
40) peptide (w33 pN) [99,100]. Unzipping experiments were also
attempted on amyloids formed from Als cell adhesion proteins of a
fungal pathogen, but in this case unzipping of mature brils was
not possible. However, zipper interactions were detected between
monolayers of Als proteins and an Als amyloid sequence attached to
an AFM tip. The characteristic force signatures corresponded to the
mechanical unzipping of
b
-sheet interactions formed between
parallel Als proteins, and was about 30 pN [101].
The weak point of the above mentioned zipping experiments is
that the measurements were based on nonspecic binding be-
tween tip and bril. Therefore, the variability between individual
measurements likely results from random and multiple attachment
sites of the protein to the tip. The nanomechanical properties of
human prion protein amyloid (huPrP90-231) were investigated
using specic binding between the AFM tip and the bril [102]. The
protein was functionalized by replacing one amino acid for a
Fig. 9. Schematic depiction of the end-to-end length E, the segment length l, and the
angle between segments 4, for a general semiexible lament.
M. Schleeger et al. / Polymer xxx (2013) 1e1610
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
cysteine, which is known to bind covalently with a gold-coated
AFM tip. Again, force-extension curves showed elastic stretching
for small retractions, characteristic of entropic stretching of a
disordered peptide chain between the
b
-sheet core and the cova-
lent attachment to the tip. At larger retractions of the tip, rupture
occurred, which likely reects the mechanical extraction of the
prion protein from the core of the amyloid bril. The force at
which rupture took place, was 115 5 pN at a tip retraction rate of
9.4 nN/s, and was dependent on the loading rate.
Even though amyloid brils are generally regarded mechanically
strong, they do break under the inuence of thermal forces. In fact,
spontaneous fracture is thought to be a key feature of the kinetics of
bril growth, since fragmentation increases the number of free
ends, thus enhancing the rate of bril growth [2,103]. It has also
been shown that amyloid brils break easily by elongational ow
[104]. Whey protein isolate (WPI) amyloid brils already fracture at
an elongational ow rate of only 8 s
1
, which is much lower than
the ow rate where DNA strands break, which is close to 10,000 s
1
.
The fracture force estimated from these observations are 0.1 pN
from the extensional ow experiment and 4 pN from the stability of
brils against thermal force [104]. It is still unclear how these low
fracture forces may be reconciled with the much higher fracture
forces of 300e500 pN estimated from active nanomanipulation
experiments of insulin brils [2]. It will be interesting to combine
mechanical measurements with vibrational spectroscopy, to link
the fracture strength to the underlying molecular structure, which
may potentially depend on the peptide used as well as on the as-
sembly conditions.
4.3. Extensibility
It was recently shown that the fundamental structural unit of
several kinds of amyloid brils including A
b
is the
b
-helix structure
[105e107]. The
b
-helix is a protein motif formed by the association
of parallel
b
-strands in a helical pattern with either two or three
faces (Fig. 11)[84]. The mechanical properties of this structure
under tension and compression were studied with molecular dy-
namics (MD) simulations. The calculated maximum tensile force
was 522 pN, while the maximum compressive force was much
higher, namely 3150 pN. The simulations revealed that the
b
-helix
structure is extremely extensible and can sustain tensile strains up
to 800% without rupture of the covalently bonded protein back-
bone [84]. The model that was used only accounts for the core
structure of certain amyloids, not including effects of side chains. It
will be interesting to test by experiments the prediction that am-
yloid brils should be highly extensible.
4.4. Outlook
Although there is quite a large number of studies of the bending
rigidity of single amyloid brils, most of these rely on AFM imaging
Fig. 10. AFM images of long, straight a. and worm-like b.
b
-lg amyloid brils. Scale bars are 500 nm. Reprinted with permission from VandenAkker et al. [60], J. Am. Chem. Soc. 2011,
133. c. Rod-like
b
-2-microglobulin amyloid brils. Image size is 1 1
m
m. Reprinted with permission from Gosal et al. [91], J. Mol. Biol. 2005, 351. d. Fluorescence microscopy image
of single yeast prion amyloid brils labeled with Thioavin T. Scale bar is 2
m
m. Reprinted with permission from Castro et al. [98], Biophys. J. 2011, 101.
Table 1
Persistence lengths of straight and worm-like
b
-lg amyloid brils, measured using
different techniques.
Reference Morphology Technique Persistence length
VandenAkker [60], 2011 Straight AFM 3820 nm 160
Adamcik [90], 2010 Straight AFM 968e3240 nm
Jordens [96], 2011 Straight AFM 2370 nm
Mudgal [93], 2009 Straight TEM 788 nm
Veerman [95], 2002 Straight TEM 1000 nm
Aymard [94], 1999 Straight LS, SANS 600 nm
Sagis [37], 2004 Straight Rheology 1600 nm 400
VandenAkker [60], 2011 Worm-like AFM 92 nm 7
Jordens [96], 2011 Worm-like AFM 29 nm
Mudgal [93], 2009 Worm-like TEM 36 nm 12
Aymard [94], 1999 Worm-like LS, SANS 38 nm
M. Schleeger et al. / Polymer xxx (2013) 1e16 11
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
of brils deposited on a substrate and dried. It is unclear how surface
immobilization and drying may affect the properties of the brils.
Several techniques can be used to circumvent these experimental
limitations. A promising method is to measure the thermal uctu-
ations of freely uctuating amyloid brils with uorescence mi-
croscopy, as was demonstrated for yeast prion amyloids [98].
Several studies have been reported of active mechanical manipu-
lation of amyloid brils. In one study, amyloid brils were bent by an
AFM tip, and in a few studies
b
-sheet segments were unzipped from
brils by AFM. A promising technique to measure both the bending
and stretch rigidity of amyloid brils is by a dual optical tweezers
assay. As demonstrated in a recent study of microtubules and actin
laments, micrometer-sized beads can be attached to the two ends
of a biopolymer and used as handles to stretch the biopolymer by
two laser tweezers [108]. From the force-extension behavior, the
persistence length of single, hydrated bers could be determined
based on well-controlled active experiments. Optical tweezers have
also been used to measure the elastic moduli of individual bers in
networks of the blood clotting protein brin [109]. Micron-sized
beads were attached to bers after clot formation and trapped
with optical tweezers. The bending and stretch rigidity of individual
bers was measured by applying an oscillatory displacement to the
bead either orthogonally or tangentially to the ber.
5. Microscopic properties of amyloid networks
For applications of amyloid brils in food products, tissue en-
gineering, and materials sciences, the mechanical properties of
networks are relevant. Network mechanics is determined by a
combination of bril mechanics and the spatial organization of -
brils and their interactions. The spatial organization of amyloid
bril networks depends on bril rigidity: when the brils are long,
thin, and rigid, they can form liquid crystalline phases or gels
already at low concentrations [110]. The interactions between -
brils are not well-studied, but are thought to be highly dependent
on the side chains on the surface of the brils, which can for
instance confer a pH-dependent electrostatic charge to the brils
[83]. Bulk rheology has been used to probe the mechanical prop-
erties of networks of amyloid brils [93,111,112]. The networks
generally form weak viscoelastic gels. However, quantitative com-
parisons between rheology measurements and theory are still
lacking, since the bril morphology was not well-dened.
5.1. Rheological properties
Rheological properties relate to the ow properties of materials
in response to an applied deformation. The most commonly used
type of deformation is a shear deformation, where a material be-
tween two parallel plates is deformed by moving one of the plates
while the other is kept stationary. Typically, one applies either a
steady shear (constant shear rate) or an oscillatory shear of
controlled strain amplitude and frequency, and measures the stress
response. When the material is a Newtonian uid such as water, the
stress in a steady shear measurement is proportional to the applied
shear rate with a constant of proportionality that is given by the
steady-shear viscosity. The viscosity is independent of the shear
rate. In contrast, polymeric materials are usually viscoelastic and
non-Newtonian, with a viscosity that does depend on shear rate.
Typically, the viscosity is constant at low strain rates, but decreases
when the strain rate is raised, a response that is known as shear
thinning. When the polymer material is elastic, a more suitable
measurement is an oscillatory shear measurement, which mea-
sures the complex shear modulus G*, which is the constant of
proportionality between the stress and the strain. G* is a complex
quantity with a storage modulus G
0
that reects the elastic stress
response that is in-phase with the applied strain, and a loss
modulus G
00
that reects the viscous stress response that is out-of-
phase with the applied strain [110]. Until now, oscillatory mea-
surements of amyloid networks have primarily focused on
resolving the gelation time and critical percolation concentration of
brils formed from food-related proteins including
b
-lg, bovine
serum albumin (BSA) and ovalbumin at pH ¼2[37,112e114]. Steady
shear measurements were used to measure the shear-rate depen-
dence of the viscosity of suspensions of whey protein brils with
lengths of several
m
m[111]. In all cases the suspensions were shear-
thinning [111,113].
Oscillatory measurements have also been used to measure the
inuence of pH and ionic strength on gel strength. At pH ¼3.35, the
whey protein
b
-lg forms worm-like brils with a persistence length
of 35 nm and a diameter of 5 nm [93]. The viscosity of these brillar
networks was observed to increase with protein concentration. In
the presence of NaCl, bril networks remained predominantly
elastic (with G
0
at least 10-fold higher than G
00
) up to ionic strengths
of 100 mM, but became weaker and eventually uid-like above
200 mM NaCl (Fig. 12)[115]. BaCl
2
and MgCl
2
caused a signicantly
Fig. 11. Structures of a. twofold and b. threefold symmetric assemblies of A
b
(1e40) amyloid brils. Reprinted with permission from Xu et al. [86], Biophys. J. 2010, 98.
M. Schleeger et al. / Polymer xxx (2013) 1e1612
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
increased nal viscosity of whey protein (WPI) gels compared to
gels formed without salt [113]. Also monovalent salts increased the
viscosity slightly compared to control brils formed without salt.
All WPI brils formed in the presence of salt, showed a worm-like
morphology, while brils formed without salt are generally long
and straight. Frequency sweep and strain sweep experiments
demonstrated that G
0
was relatively insensitive to added salt, but
the loss modulus G
00
was slightly higher with divalent salts than for
samples without or with monovalent salts [113]. The viscosity de-
pends on different parameters, including the volume fraction of
brils, the bril morphology and the interactions between brils.
Because the effect of mono- and multivalent salts on these pa-
rameters was not determined, the mechanism behind the increase
in viscosity is still not understood.
There are few rheological studies of amyloids other than whey
proteins-related brils. Weak, solid-like behavior was also reported
for hydrogels formed from amyloids of the Alzheimer related A
b
peptide, A
b
(16e20) formed at high concentrations (3% w/v) [112].
An alternative method to measure the rheology of soft poly-
meric materials is by microrheology (MR), which is a collection of
techniques used to determine the local properties of a material
from the motion of embedded, small tracer particles. The most
popular MR technique is video particle tracking, where the ther-
mally induced motions of the tracer particles are observed by op-
tical microscopy [116]. Particle tracking is a convenient method to
obtain a spatial map of variations in the viscoelastic properties of
heterogenous samples [117]. An advantage of MR over bulk
rheology is the possibility to use small sample volumes (down to
5
m
L) and the possibility to determine the rheology over a broad
frequency range. Particle tracking MR has been used to study the
solegel transition in amyloid networks. Gels of
b
-lg brils were
prepared at pH ¼7 and room temperature by adding alcohol [116].
Under these conditions, wormlike brils are formed [118]. Latex
particles with a diameter of 0.5
m
m were used to observe the time
evolution of network formation for
b
-lg concentrations at and
above the critical gelation concentration of 4% w/v. In time, a shift
from purely viscous to viscoelastic behavior was observed, indica-
tive of bril formation. Eventually, a solid gel was formed. Based on
the mean squared displacement (MSD) of the probe particles, the
elastic and loss moduli over a large frequency range could be
Fig. 12. Rheological characterization of 2% w/v
b
-lg protein brils at varying ionic strengths: A. 50 mM, B.100 mM, and C. 200 mM NaCl. D. Storage modulus determined at different
ionic strengths by frequency sweeps. E. Structural recovery behavior of the bril gel prepared at the ionic strength of 50 mM as a function of time after destruction of the gel by a
100% oscillatory shear strain. Reprinted with permission from Ref. [115].
M. Schleeger et al. / Polymer xxx (2013) 1e16 13
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
calculated (Fig. 13). Similar experiments were also performed for
b
-lg amyloid brils formed by decreasing the pH to 2 and heating to
80
C. Under these conditions, the brils that are formed are long
and straight, and the critical concentration for gel formation
was less than 3% w/v. The gels that were formed after about
100 min showed behavior typical of weak gels, with a plateau in G
0
that became more pronounced at longer incubation times up to
200 min [119].
5.2. Outlook
Particle tracking microrheology has a frequency range that is
limited by the camera acquisition rate and also by the inherent
localization error of ca. 20e50 nm in the particle position. These
limitations can be overcome by using laser tweezers combined
with quadrant photodiodes for sub-nanometer localization of
probe particles embedded in a network. By using weak laser beams,
the thermal uctuations of the particles can be detected by laser-
interferometry. As shown for actin networks and worm-like mi-
celles, this method affords a wide frequency window of 0.1e
100 kHz, giving at the same time access to network properties
(below w1 kHz) and single bril properties (above w10 kHz)
[120,121]. Optical tweezers can also be used for active MR, where a
probe particle is actively moved by a laser trap at controlled fre-
quency and amplitude. This method in principle allows measure-
ment of non-linear viscoelastic properties.
6. Conclusions/perspective
We have illustrated here how insights into amyloid chemistry
and physics on different length scales ranging from the molecular
to the macroscopic level can contribute to an improved under-
standing of amyloids and amyloid networks. The overarching goal
of these efforts is to obtain a comprehensive understanding of the
relation between secondary structure, amino acid distribution and
bril assembly on the one hand and morphology and mechanical
properties of amyloids and networks thereof on the other hand.
Network mechanics are necessarily a function of bril mechanics,
the architecture of networks, and interactions between brils. Fi-
brils mechanics are determined by the assembly and (core-)
structure of brils, as well as interactions of single amino acids
within the brils. It is evidently challenging to bridge the length
scales from the molecular morphology, through single bril me-
chanical properties, to the macroscopic rheological properties for
these highly complex biopolymers, but some of the examples
presented above demonstrate signicant progress in this eld. The
recent development and application of advanced vibrational
spectroscopic techniques to these complex biological systems show
promise: several examples exist where a combination of tech-
niques focusing on the properties of amyloids at different length
scales was used. For example, the combination of vSFG and AFM
allowed relating the persistence length of different assembly con-
ditions to the global secondary structure composition of amyloids.
The change in morphology and persistence length could be shown
to be accompanied by a change in the secondary structure content
of the amyloid samples. Sophisticated FT-IR and Raman spectros-
copy can provide a comprehensive understanding of the secondary
structure of amyloids [52,53,118], and may be combined with me-
chanical deformation in the same setup, as shown for other bio-
polymers [122]. The advent of several new vibrational spectroscopy
techniques and microrheology approaches offers the possibility to
interrogate macroscopic, mesoscopic and microscopic properties of
amyloid brils.
Acknowledgment
This work is supported by NanoNextNL, a micro and nano-
technology consortium of the Government of the Netherlands and
130 partners, as well as part of the Industrial Partnership Pro-
gramme (IPP) Bio(-Related) Materials (BRM) of the Stichting voor
Fundamenteel Onderzoek der Materie (FOM), which is nancially
supported by the Nederlandse Organisatie voor Wetenschappelijk
Onderzoek (NWO). The IPP BRM is co-nanced by the Top Institute
Food and Nutrition and the Dutch Polymer Institute.
References
[1] Vollrath F, Porter D. Polymer 2009;50(24):5623e32.
[2] Smith JF, Knowles TP, Dobson CM, Macphee CE, Welland ME. Proc Natl Acad
Sci U S A 2006;103(43):15806e11.
[3] Knowles TP, Fitzpatrick AW, Meehan S, Mott HR, Vendruscolo M, Dobson CM,
et al. Science 2007;318(5858):1900e3.
[4] Nyrkova IA, Semenov AN, Aggeli A, Bell M, Boden N, McLeish TCB. Eur Phys J
B 2000;17(3):499e513.
[5] Dinca V, Kasotakis E, Catherine J, Mourka A, Ranella A, Ovsianikov A, et al.
Nano Lett 2008;8(2):538e43.
[6] Scheibel T, Parthasarathy R, Sawicki G, Lin XM, Jaeger H, Lindquist SL. Proc
Natl Acad Sci U S A 2003;100(8):4527e32.
[7] Hamada D, Yanagihara I, Tsumoto K. Trends Biotechnol 2004;22(2):93e7.
[8] Maji SK, Schubert D, Rivier C, Lee S, Rivier JE, Riek R. PLoS Biol 2008;6(2):
240e51.
[9] Gelain F, Bottai D, Vescovi A, Zhang S. PLoS One 2006;1:e119.
[10] Hu KN, McGlinchey RP, Wickner RB, Tycko R. Biophys J 2011;101(9):2242e50.
[11] Berryman JT, Radford SE, Harris SA. Biophys J 2011;100(9):2234e42.
[12] Kodali R, Wetzel R. Curr Opin Struct Biol 2007;17(1):48e57.
[13] Sipe JD, Cohen AS. J Struct Biol 2000;130(2e3):88e98.
[14] Lin JC, Liu HL. Curr Drug Discov Technol 2006;3(2):145e53.
[15] Iconomidou VA, Hamodrakas SJ. Curr Protein Pept Sci 2008;9(3):291e309.
[16] Jahn TR, Makin OS, Morris KL, Marshall KE, Tian P, Sikorski P, et al. J Mol Biol
2010;395(4):717e27.
[17] Wiltzius JJ, Sievers SA, Sawaya MR, Cascio D, Popov D, Riekel C, et al. Protein
Sci 2008;17(9):1467e74.
[18] Wetzel R, Shivaprasad S, Williams AD. Biochemistry 2007;46(1):1e10.
[19] Knowles TP, Smith JF, Craig A, Dobson CM, Welland ME. Phys Rev Lett
2006;96(23):238301.
[20] Guo S, Akhremitchev BB. Biomacromolecules 2006;7(5):1630e6.
[21] Pastor MT, Esteras-Chopo A, Lopez de la Paz M. Curr Opin Struct Biol
2005;15(1):57e63.
[22] Hamley IW. Angew Chem Int Ed Engl 2007;46(43):8128e47.
Fig. 13. Elastic and loss moduli of pre-gel (lower two curves) and post-gel (upper two
curves) systems versus frequency, a and b are shift factors. For clarity, the postgel
curves are shifted upward by a factor of 10
2
, and only every fourth data point is plotted
for each curve. The pregel curve shows viscous behavior at low frequencies (a slope of
1 for G00), while at high frequencies the moduli approach critical behavior and the
power law exponent is 0.59, comparable to the MSD power law exponent at the gel
point. Reprinted with permission from Corrigan et al. [116] Langmuir, 2009, 25.
Copyright 2009 American Chemical Society.
M. Schleeger et al. / Polymer xxx (2013) 1e1614
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
[23] Baldwin AJ, Bader R, Christodoulou J, MacPhee CE, Dobson CM, Barker PD.
J Am Chem Soc 2006;128(7):2162e3.
[24] Hamedi M, Herland A, Karlsson RH, Inganas O. Nano Lett 2008;8(6):1736e40.
[25] Cellmer T, Bratko D, Prausnitz JM, Blanch HW. Trends Biotechnol 2007;25(6):
254e61.
[26] Sinha S, Lieberburg I. Proc Natl Acad Sci U S A 1999;96(20):11049e53.
[27] Lansbury PT, Lashuel HA. Nature 2006;443(7113):774e9.
[28] Middleton CT, Marek P, Cao P, Chiu CC, Singh S, Woys AM, et al. Nat Chem
2012;4(5):355e60.
[29] Höppener JWM, Ahrén B, Lips CJM. N Engl J Med 2000;343(6):411e9.
[30] Kayed R, Head E, Thompson JL, McIntire TM, Milton SC, Cotman CW, et al.
Science 2003;300(5618):486e9.
[31] Engel MFM, Khemtémourian L, Kleijer CC, Meeldijk HJD, Jacobs J, Verkleij AJ,
et al. Proc Natl Acad Sci U S A 2008;105(16):6033e8.
[32] Buttereld SM, Lashuel HA. Angew Chem Int Ed 2010;49(33):5628e54.
[33] Kontopidis G, Holt C, Sawyer L. J Dairy Sci 2004;87(4):785e96.
[34] Gosal WS, Clark AH, Ross-Murphy SB. Biomacromolecules 2004;5(6):2420e9.
[35] Gosal WS, Clark AH, Ross-Murphy SB. Biomacromolecules 2004;5(6):2430e8.
[36] Arnaudov LN, de Vries R. Biomacromolecules 2006;7(12):3490e8.
[37] Sagis LM, Veerman C, van der Linden E. Langmuir 2004;20(3):924e7.
[38] Arnaudov LN, de Vries R. J Chem Phys 2007;126(14):145106.
[39] Padrick SB, Miranker AD. Biochemistry 2002;41(14):4694e703.
[40] Nanga RPR, Brender JR, Vivekanandan S, Ramamoorthy A. Biochim Biophys
Acta Biomembr 2011;1808(10):2337e42.
[41] Nielsen JT, Bjerring M, Jeppesen MD, Pedersen RO, Pedersen JM, Hein KL,
et al. Angew Chem Int Ed Engl 2009;48(12):2118e21.
[42] Kapurniotu A. Biopol Pept Sci 2001;60(6):438e59.
[43] Ono K, Yoshiike Y, Takashima A, Hasegawa K, Naiki H, Yamada M.
J Neurochem 2003;87(1):172e81.
[44] Ehrnhoefer DE, Bieschke J, Boeddrich A, Herbst M, Masino L, Lurz R, et al. Nat
Struct Mol Biol 2008;15(6):558e66.
[45] Cheynier V. Am J Clin Nutr 2005;81(1 Suppl.):223Se9S.
[46] Bieschke J, Russ J, Friedrich RP, Ehrnhoefer DE, Wobst H, Neugebauer K, et al.
Proc Natl Acad Sci U S A 2010;107(17):7710e5.
[47] Fresco P, Borges F, Diniz C, Marques MP. Med Res Rev 2006;26(6):747e66.
[48] Manach C, Scalbert A, Morand C, Remesy C, Jimenez L. Am J Clin Nutr
2004;79(5):727e47.
[49] Shoval H, Lichtenberg D, Gazit E. Amyloid 2007;14(1):73e87.
[50] Azriel R, Gazit E. J Biol Chem 2001;276(36):34156e61.
[51] Necula M, Kayed R, Milton S, Glabe CG. J Biol Chem 2007;282(14):10311e24.
[52] Oboroceanu D, Wang LZ, Brodkorb A, Magner E, Auty MAE. J Agric Food
Chem 2010;58(6):3667e73.
[53] Oboroceanu D, Wang LZ, Kroes-Nijboer A, Brodkorb A, Venema P, Magner E,
et al. Int Dairy J 2011;21(10):823e30.
[54] Arnolds H, Bonn M. Surf Sci Rep 2010;65(2):45e66.
[55] Pettinger B, Schambach P, Villagomez CJ, Scott N. Annu Rev Phys Chem
2012;63:379e99.
[56] Bailo E, Deckert V. Chem Soc Rev 2008;37(5):921e30.
[57] Deckert-Gaudig T, Kammer E, Deckert V. J Biophotonics 2012;5(3):215e9.
[58] Bonn M, Bakker HJ, Ghosh A, Yamamoto S, Sovago M, Campen RK. J Am Chem
Soc 2010;132(42):14971e8.
[59] Sovago M, Vartiainen E, Bonn M. J Chem Phys 2009;131(16):161107.
[60] vandenAkker CC, Engel MFM, Velikov KP, Bonn M, Koenderink GH. J Am
Chem Soc 2011;133(45):18030e3.
[61] Fu L, Ma G, Yan ECY. J Am Chem Soc 2010;132(15):5405e12.
[62] Fu L, Liu J, Yan EC. J Am Chem Soc 2011;133(21):8094e7.
[63] Xiao D, Fu L, Liu J, Batista VS, Yan EC. J Mol Biol 2011.
[64] Engel MF, Vandenakker CC, Schleeger M, Velikov KP, Koenderink GH,
Bonn M. J Am Chem Soc 2012;134(36):14781e8.
[65] Meng F, Abedini A, Plesner A, Verchere CB, Raleigh DP. Biochemistry
2010;49(37):8127e33.
[66] Diesner MO, Welle A, Kazanci M, Kaiser P, Spatz J, Koelsch P. Biointerphases
2011;6(4):171e9.
[67] Kurouski D, Deckert-Gaudig T, Deckert V, Lednev IK. J Am Chem Soc
2012;134(32):13323e9.
[68] Hamm P, Zanni MT. Concepts and methods of 2D infrared spectroscopy 2011.
[69] Cervetto V, Helbing J, Bredenbeck J, Hamm P. J Chem Phys 2004;121(12):
5935e42.
[70] Bredenbeck J, Helbing J, Nienhaus K, Nienhaus GU, Hamm P. Proc Natl Acad
Sci U S A 2007;104(36):14243e8.
[71] Strasfeld DB, Ling YL, Gupta R, Raleigh DP, Zanni MT. J Phys Chem B
2009;113(47):15679e91.
[72] Shim SH, Zanni MT. Phys Chem Chem Phys 2009;11(5):748e61.
[73] Wang L, Middleton CT, Singh S, Reddy AS, Woys AM, Strasfeld DB, et al. J Am
Chem Soc 2011;133(40):16062e71.
[74] Baiz CR, Peng CS, Reppert ME, Jones KC, Tokmakoff A. Analyst 2012;137(8):
1793e9.
[75] Zhuang W, Hayashi T, Mukamel S. Angew Chem Int Ed Engl 2009;48(21):
3750e81.
[76] Luca S, Yau WM, Leapman R, Tycko R. Biochemistry 2007;46(47):13505e22.
[77] Paravastu AK, Leapman RD, Yau WM, Tycko R. Proc Natl Acad Sci U S A
2008;105(47):18349e54.
[78] Shim SH, Gupta R, Ling YL, Strasfeld DB, Raleigh DP, Zanni MT. Proc Natl Acad
Sci U S A 2009;106(16):6614e9.
[79] Strasfeld DB, Ling YL, Shim SH, Zanni MT. J Am Chem Soc 2008;130(21):
6698e9.
[80] Ling YL, Strasfeld DB, Shim SH, Raleigh DP, Zanni MT. J Phys Chem B
2009;113(8):2498e505.
[81] Meng F, Abedini A, Plesner A, Middleton CT, Potter KJ, Zanni MT, et al. J Mol
Biol 2010;400(3):555e66.
[82] Cao P, Meng F, Abedini A, Raleigh DP. Biochemistry 2010;49(5):872e81.
[83] Knowles TP, Buehler MJ. Nat Nanotechnol 2011;6(8):469e79.
[84] KetenS, Buehler MJ. Comput Meth Appl Mech Eng 2008;197(41e42):3203e14.
[85] Jones OG, Mezzenga R. Soft Matter 2012;8(4):876e95.
[86] Xu Z, Paparcone R, Buehler MJ. Biophys J 2010;98(10):2053e62.
[87] Mucke N, Klenin K, Kirmse R, Bussiek M, Herrmann H, Hafner M, et al. PLoS
One 2009;4(11):e7756.
[88] Brangwynne CP, Koenderink GH, Barry E, Dogic Z, MacKintosh FC, Weitz DA.
Biophys J 2007;93(1):346e59.
[89] Relini A, Torrassa S, Ferrando R, Rolandi R, Campioni S, Chiti F, et al. Biophys J
2010;98(7):1277e84.
[90] Adamcik J, Jung JM, Flakowski J, De Los Rios P, Dietler G, Mezzenga R. Nat
Nanotechnol 2010;5(6):423e8.
[91] Gosal WS, Morten IJ, Hewitt EW, Smith DA, Thomson NH, Radford SE. J Mol
Biol 2005;351(4):850e64.
[92] Sachse C, Grigorieff N, Fandrich M. Angew Chem Int Ed 2010;49(7):1321e3.
[93] MudgalP, Daubert CR, Foegeding EA. Food Hydrocolloids 2009;23(7):1762e70.
[94] Aymard P, Nicolai T, Durand D, Clark A. Macromolecules 1999;32(8):2542e52.
[95] Veerman C, Ruis H, Sagis LM, van der Linden E. Biomacromolecules
2002;3(4):869e73.
[96] Jordens S, Adamcik J, Amar-Yuli I, Mezzenga R. Biomacromolecules
2011;12(1):187e93.
[97] Ban T, Hamada D, Hasegawa K, Naiki H, Goto Y. J Biol Chem 2003;278(19):
16462e5.
[98] Castro CE, Dong J, Boyce MC, Lindquist S, Lang MJ. Biophys J 2011;101(2):
439e48.
[99] Kellermayer MS, Grama L, Karsai A, Nagy A, Kahn A, Datki ZL, et al. J Biol
Chem 2005;280(9):8464e70.
[100] Karsai A, Martonfalvi Z, Nagy A, Grama L, Penke B, Kellermayer MS. J Struct
Biol 2006;155(2):316e26.
[101] Alsteens D, Ramsook CB, Lipke PN, Dufrene YF. ACS Nano 2012;6(9):7703e11.
[102] Ganchev DN, Cobb NJ, Surewicz K, Surewicz WK. Biophys J 2008;95(6):
2909e15.
[103] Tanaka M, Collins SR, Toyama BH, Weissman JS. Nature 2006;442(7102):
585e9.
[104] Kroes-Nijboer A, Venema P, Baptist H, van der Linden E. Langmuir
2010;26(16):13097e101.
[105] Govaerts C, Wille H, Prusiner SB, Cohen FE. Proc Natl Acad Sci U S A
2004;101(22):8342e7.
[106] Kishimoto A, Hasegawa K, Suzuki H, Taguchi H, Namba K, Yoshida M.
Biochem Biophys Res Commun 2004;315(3):739e45.
[107] Perutz MF, Finch JT, Berriman J, Lesk A. Proc Natl Acad Sci U S A 2002;99(8):
5591e5.
[108] van Mameren J, Vermeulen KC, Gittes F, Schmidt CF. J Phys Chem B
2009;113(12):3837e44.
[109] Collet JP, Shuman H, Ledger RE, Lee S, Weisel JW. Proc Natl Acad Sci U S A
2005;102(26):9133e7.
[110] Loveday SM, Rao MA, Creamer LK, Singh H. J Food Sci 2009;74(3):R47e55.
[111] Akkermans C, van der Goot AJ, Venema P, van der Linden E, Boom RM. Food
Hydrocolloids 2008;22(7):1315e25.
[112] Krysmann MJ, Castelletto V, Kelarakis A, Hamley IW, Hule RA, Pochan DJ.
Biochemistry 2008;47(16):4597e605.
[113] LovedaySM, Su JH, Rao MA, Anema SG, Singh H. Int Dairy J 2012;26(2):133e40.
[114] Veerman C, de Schiffart G, Sagis LM, van der Linden E. Int J Biol Macromol
2003;33(1e3):121e7.
[115] Bolisetty S, Harnau L, Jung JM, Mezzenga R. Biomacromolecules 2012;13(10):
3241e52.
[116] Corrigan AM, Donald AM. Langmuir 2009;25(15):8599e605.
[117] Valentine MT, Kaplan PD, Thota D, Crocker JC, Gisler T, Prudhomme RK, et al.
Phys Rev E Stat Nonlin Soft Matter Phys 2001;64(6 Pt 1):061506.
[118] GosalWS, Clark AH, Ross-Murphy SB. Biomacromolecules 2004;5(6):2408e19.
[119] Corrigan AM, Donald AM. Eur Phys J E Soft Matter 2009;28(4):457e62.
[120] Koenderink GH, Atakhorrami M, MacKintosh FC, Schmidt CF. Phys Rev Lett
2006;96(13):138307.
[121] Atakhorrami M, Mizuno D, Koenderink GH, Liverpool TB, MacKintosh FC,
Schmidt CF. Phys Rev E Stat Nonlin Soft Matter Phys 2008;77(6 Pt 1):061508.
[122] Boulet-Audet M, Vollrath F, Holland C. Phys Chem Chem Phys 2011;13(9):
3979e84.
M. Schleeger et al. / Polymer xxx (2013) 1e16 15
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
Michael Schleeger received his diploma in chemistry from
the University of Bonn (2004). For his Ph.D. thesis, he
joined the group of Joachim Heberle at Bielefeld Univer-
sity, where he received his Ph.D. degree in 2009. After 2
years of postdoctoral research in the Department of
Physics at the FU Berlin, he joined the group of Prof.
Mischa Bonn at the Max-Planck Institute for Polymer
research in Mainz as a research fellow in 2012. His current
research focus is on the structure and interactions of am-
yloids based on vibrational spectroscopy, in particular
sum frequency generation spectroscopy.
Corianne van den Akker obtained her Masters degree in
Biomedical Engineering from the University of Twente,
The Netherlands in 2010. She is currently studying for a
PhD with Prof. Gijsje Koenderink at FOM Institute AMOLF
in Amsterdam, with research orientated towards the me-
chanical properties and molecular conformation of amy-
loid brils.
Dr. Tanja Deckert-Gaudig works as a postdoc in the
research group of Prof. V. Deckert at the Institute of Pho-
tonic Technology in Jena. She obtained her Ph.D. in Organic
Chemistry in 1997 with Prof. S. Hünig at the University of
Würzburg. After a maternal leave she started working on
TERS with Prof. Deckert at the University of Dresden in
2002, followed by the Institute of Analytical Sciences in
Dortmund. Since 2009 she works in Jena, where her main
interest is the structural elucidation of biopolymers on the
nanoscale by TERS.
Prof. Volker Deckert holds a joint position at the Institute
of Physical Chemistry at the University of Jena, Germany
and the Institute of Photonic Technology, also in Jena. He
obtained his Diploma and Ph.D. from the University of
Würzburg, Germany, working on difference-Raman spec-
troscopy. As a postdoc at the University of Tokyo and the
Kanagawa Academy of Science in Kawasaki, he worked
on non-linear and time-resolved laser spectroscopy of
photo-induced isomerisation reactions . During his habili-
tation at the ETH Zurich he started working on the devel-
opment of high spatial resolution techniques for Raman
spectroscopy. This topic was the basis of his next positions
at the TU Dresden, ISAS Dortmund and now in Jena, where
he in particular explores the possibilities of the technology
to investigate structural changes of bio-related compounds
with nanometer resolution.
Dr. Krassimir Velikov is a Science/Team Leader in Unilever
R&D Vlaardingen. He received his PhD from the University
of Utrecht, The Netherlands. His main research interests
cover topics of soft-condensed ma tter, se lf-as sembly,
colloid and interface science of dispersions (e.g. suspen-
sions, foams, emulsions) and their uses to control product
functionality (e.g. stability, appearance, texture), physical-
chemistry of digestion, and formulation and delivery of
functional ingredients. He is an adjunct assistant professor
in the Soft Condensed Matter group at the Debye Institute
for NanoMaterials Science, Utrecht University and a special
adjunct associate professor in the Department of Chemical
and Biomolecular Engineering in the College of Engineer-
ing, North Carolina State University (USA). He is a Program
Director of Molecular Structure of Food program in
NanoNextNL.
Gijsje Koenderink received her Ph.D. from Utrecht Uni-
versity in 2003 in the area of physical and colloid chem-
istry. She conducted postdoctoral research at the VU
University (Amsterdam) and Harvard University (Cam-
bridge MA) on biophysical properties of cytoskeletal bio-
polymers. She joined the FOM Institute AMOLF in
Amsterdam as a group leader in 2006, where she started
the group Biological Soft Matter. Her research interests
focus on the structural and mechanical properties of bio-
polymers, the non-equilibrium properties of the cellular
cytoskeleton, and cellular mechanosensing.
Mischa Bonn (January 25, 1971, The Netherlands) received
his PhD in 1996 for work resulting from a collaboration
between the Technical University of Eindhoven and the
FOM Institute for Atomic and Molecular Physics, AMOLF.
After post-doctoral work at the Fritz-Haber Institute in
Berlin and Columbia University in New York, Mischa
worked at the chemistry department at Leiden University
(1999e2004), before moving to AMOLF (2004) as head of
the Biosurface Spectroscopygroup. Mischa received
2009 the Gold Medal from the Royal Dutch Chemical Soci-
ety. In 2011 Mischa was appointed as one of the Directors
of the Max Planck Institute for Polymer Research in Mainz.
His research is centered around laser-based vibrational
spectroscopies, specically surface and THz spectroscopies
and CARS microscopy.
M. Schleeger et al. / Polymer xxx (2013) 1e1616
Please cite this article in press as: Schleeger M, et al., Amyloids: From molecular structure to mechanical properties, Polymer (2013), http://
dx.doi.org/10.1016/j.polymer.2013.02.029
... The mechanical behaviour has been widely investigated with a view on potential applications, but also as a way to gain further insight into the amyloid fibril structure and its formation [12][13][14][15]. Atomic force microscopy and molecular dynamics simulations are often the preferred tools to evaluate the mechanical properties [12,14,[16][17][18]. ...
... The mechanical behaviour has been widely investigated with a view on potential applications, but also as a way to gain further insight into the amyloid fibril structure and its formation [12][13][14][15]. Atomic force microscopy and molecular dynamics simulations are often the preferred tools to evaluate the mechanical properties [12,14,[16][17][18]. We have previously studied the mechanical behaviour of TTR 105-115 amyloid fibrils and cellulose using high-pressure X-ray diffraction and Raman spectroscopy [19,20]. ...
... The analysis yields a bulk modulus K of 2.48 ± 0.14 GPa and a pressure-derivative K equal to 8.54 ± 0.80 GPa. This value is similar to that reported for TTR 105-115 fibrils [19] and is of the same order as Young's moduli determined by AFM and computational methods [11,12,15,17]. In fact, if we assume a Poisson's ratio of 0.3 [18], then we can calculate Young's modulus of the SSTSAA crystals to be 3 GPa. ...
Article
Full-text available
Amyloid fibrils have been associated with human disease for many decades, but it has also become apparent that they play a functional, non-disease-related role in e.g. bacteria and mammals. Moreover, they have been shown to possess interesting mechanical properties that can be harnessed for future man-made applications. Here, the mechanical behaviour of SSTSAA microcrystals has been investigated. The SSTSAA peptide organization in these microcrystals has been related to that in the corresponding amyloid fibrils. Using high-pressure X-ray diffraction experiments, the bulk modulus K, which is the reciprocal of the compressibility β, has been calculated to be 2.48 GPa. This indicates that the fibrils are tightly packed, although the packing of most native globular proteins is even better. It is shown that the value of the bulk modulus is mainly determined by the compression along the c-axis, that relates to the inter-sheet distance in the fibrils. These findings corroborate earlier data obtained by AFM and molecular dynamics simulations that showed that mechanical resistance varies according to the direction of the applied strain, which can be related to packing and hydrogen bond contributions. Pressure experiments provide complementary information to these techniques and help to acquire a full mechanical characterization of biomolecular assemblies. This article is part of the theme issue ‘Exploring the length scales, timescales and chemistry of challenging materials (Part 2)’.
... The mechanical properties of amyloid fibrils are also exceptional, with Young's moduli comparable to the highest recorded among other proteinaceous materials such as keratin, collagen, and silk. [34][35][36] These are examples of protein-based materials that are incredibly durable, strong, and versatile while being 100% biobased. Amyloid fibrils are, again due to their extensive intermolecular bonding, incredibly resilient toward chemical denaturation and enzymatic degradation. ...
Article
Full-text available
Functional amyloid (FAs), particularly the bacterial proteins CsgA and FapC, have many useful properties as biomaterials: high stability, efficient, and controllable formation of a single type of amyloid, easy availability as extracellular material in bacterial biofilm and flexible engineering to introduce new properties. CsgA in particular has already demonstrated its worth in hydrogels for stable gastrointestinal colonization and regenerative tissue engineering, cell‐specific drug release, water‐purification filters, and different biosensors. It also holds promise as catalytic amyloid; existing weak and unspecific activity can undoubtedly be improved by targeted engineering and benefit from the repetitive display of active sites on a surface. Unfortunately, FapC remains largely unexplored and no application is described so far. Since FapC shares many common features with CsgA, this opens the window to its development as a functional scaffold. The multiple imperfect repeats in CsgA and FapC form a platform to introduce novel properties, e.g., in connecting linkers of variable lengths. While exploitation of this potential is still at an early stage, particularly for FapC, a thorough understanding of their molecular properties will pave the way for multifunctional fibrils which can contribute toward solving many different societal challenges, ranging from CO2 fixation to hydrolysis of plastic nanoparticles.
Chapter
Protein-based nanoencapsulation possesses a higher compound/drug loading capacity than other nanostructures. It improves the absorption and bioavailability of the encapsulated compounds (Abaee et al. 2017; Chen et al. 2006). These nanostructures are prepared by the hydrophobic/hydrophilic interaction of bioactive compounds with the encapsulation matrices. Protein-based nanostructures are responsive to changes in the environment, such as pH change, temperature, enzymatic conditions, and ionic strength, making them suitable candidates for the targeted delivery of bioactive compounds to specified sites (Fang et al. 2014). Several types of proteins, such as whey, zein, and collagens, are used to form these nanocarriers. The release of these encapsulated compounds depends on their interaction with the encapsulation matrix; hydrophilic compounds are dispersed by diffusion, whereas hydrophobic compounds are released through enzymatic degradation of the protein matrix in the gastrointestinal tract (GIT). Additionally, these structures possess several limitations, such as disruption by the presence of protease enzymes in the GIT, making it a challenge to deliver bioactive compounds encapsulated with protein matrices (Bourbon et al. 2011; Donato-Capel et al. 2014). Nevertheless, there are different types of protein-based nanostructures, such as nanoparticles, nanohydrogels, nanotubes, hollow nanoparticles, nanofibrillar aggregates, electrospun nanofibers, and native state proteins as natural nanocarriers cited in the literature (Fig. 3.1) (Mohammadian et al. 2020).
Article
Proteinaceous amyloid fibrils are one of the stiffest biopolymers due to their extensive cross-β-sheet quaternary structure, whereas cellulose nanofibrils (CNFs) exhibit interesting properties associated with their nanoscale size, morphology, large surface area, and biodegradability. Herein, CNFs were supplemented with amyloid fibrils assembled from the Curli-specific gene A (CsgA) protein, the main component of bacterial biofilms. The resulting composites showed superior mechanical properties, up to a 7-fold increase compared to unmodified CNF films. Wettability and thermogravimetric analyses demonstrated high surface hydrophobicity and robust thermal tolerance. Bulk spectroscopic characterization of CNF-CsgA films revealed key insights into the molecular organization within the bionanocomposites. Atomic force microscopy and photoinduced force microscopy revealed the high-resolution location of curli assemblies into the CNF films. This novel sustainable and cost-effective CNF-based bionanocomposites supplemented with intertwined bacterial amyloid fibrils opens novel directions for environmentally friendly applications demanding high mechanical, water-repelling properties, and thermal resistance.
Article
Full-text available
To address the increasing environmental footprint of the fast-growing textile industry, self-repairing textile composites have been developed to allow torn or damaged textiles to restore their morphological, mechanical, and functional features. A sustainable way to create these textile composites is to introduce a coating material that is biologically derived, biodegradable, and can be produced through scalable processes. Here, we fabricated self-repairing textile composites by integrating the biofilms of Escherichia coli (E. coli) bacteria into conventional knitted textiles. The major structural protein component in E. coli biofilm is a matrix of curli fibers, which has demonstrated extraordinary abilities to self-assemble into mechanically strong macroscopic structures and self-heal upon contact with water. We demonstrated the integration of biofilm through three simple, fast, and scalable methods: adsorption, doctor blading, and vacuum filtration. We confirmed that the composites were breathable and mechanically strong after the integration, with improved Young’s moduli or elongation at break depending on the fabrication method used. Through patching and welding, we showed that after rehydration, the composites made with all three methods effectively healed centimeter-scale defects. Upon observing that the biofilm strongly attached to the textiles by covering the extruding textile fibers from the self-repair failures, we proposed that the strength of the self-repairs relied on both the biofilm’s cohesion and the biofilm-textile adhesion. Considering that curli fibers are genetically-tunable, the fabrication of self-repairing curli-expressing biofilm-textile composites opens new venues for industrially manufacturing affordable, durable, and sustainable functional textiles.
Article
Protein-based biomaterials, particularly amyloids, have sparked considerable scientific interest in recent years due to their exceptional mechanical strength, excellent biocompatibility and bioactivity. In this work, we have synthesized a novel amyloid-based composite hydrogel consisting of bovine serum albumin (BSA) and aloe vera (AV) gel to utilize the medicinal properties of the AV gel and circumvent its mechanical frangibility. The synthesized composite hydrogel demonstrated an excellent porous structure, self-fluorescence, non-toxicity, and controlled rheological properties. Moreover, this hydrogel possesses inherent antioxidant and antibacterial properties, which accelerate the rapid healing of wounds. The in vitro wound healing capabilities of the synthesized composite hydrogel were evaluated using 3T3 fibroblast cells. Moreover, the efficacy of the hydrogel in accelerating chronic wound healing via collagen crosslinking was investigated through in vivo experiments using a diabetic mouse skin model. The findings indicate that the composite hydrogel, when applied, promotes wound healing by inducing collagen deposition and upregulating the expression of vascular endothelial growth factor (VEGF) and its receptors. We also demonstrate the feasibility of the 3D printing of the BSA-AV hydrogel, which can be tailored to treat various types of wound. The 3D printed hydrogel exhibits excellent shape fidelity and mechanical properties that can be utilized for personalized treatment and rapid chronic wound healing. Taken together, the BSA-AV hydrogel has great potential as a bio-ink in tissue engineering as a dermal substitute for customizable skin regeneration.
Article
Full-text available
Silks have a great potential as sustainable, ecologically benign commercial polymers. Here we discuss this fascinating bio-material by merging the biologist's with the polymer scientist's views i.e. combine insights into the characterisation and understanding of evolved structure, property and function in natural silk proteins with the broad scope of applied disciplines ranging from molecular modelling to rheology and mechanical testing. We conclude that silk cannot be defined simply by only its origin or material composition but any meaningful designation must include the key feature of formation by extrusion spinning. We further conclude that silk ‘spinning’ largely depends on a highly specific denaturation process dependent on competing molecular-level interactions of hydrogen bonding between water and main chain amide groups in the silk protein chains. Finally we conclude that silks have a bright future not only as archetype models to guide our understanding of highly adapted and energy efficient bio-polymers but also as prototype models to guide the design of totally novel polymer systems.
Article
Full-text available
Self-assembly of whey proteins into amyloid-like fibrils during heating at pH 2 and low ionic strength is sensitive to the presence of NaCl and CaCl2. Our earlier work established that at 10e120 mM of these salts speeds up self-assembly and favours short, flexible fibrils over long semiflexible fibrils in a way that depends on cation concentration and cation type. Here we explored how other monovalent and divalent salts affected fibril morphology and the rheology of fibril dispersions. Divalent salts (MgCl2, CaCl2, BaCl2) had much stronger effects than monovalent salts (LiCl, NaCl, KCl) on gelation kinetics, and differences between salts of the same type were not large. No marked effects of salt type on fibril morphology were evident, but there were subtle differences in the size and extent of fibril networks with monovalent versus divalent salts, which may explain differences in bulk rheology.
Article
2D infrared (IR) spectroscopy is a cutting-edge technique, with applications in subjects as diverse as the energy sciences, biophysics and physical chemistry. This book introduces the essential concepts of 2D IR spectroscopy step-by-step to build an intuitive and in-depth understanding of the method. This unique book introduces the mathematical formalism in a simple manner, examines the design considerations for implementing the methods in the laboratory, and contains working computer code to simulate 2D IR spectra and exercises to illustrate involved concepts. Readers will learn how to accurately interpret 2D IR spectra, design their own spectrometer and invent their own pulse sequences. It is an excellent starting point for graduate students and researchers new to this exciting field. Computer codes and answers to the exercises can be downloaded from the authors' website, available at www.cambridge.org/9781107000056.
Article
Proteinfibrils are relevant not only in medicine and amyloid-related neurodegenerative diseases, but also as functional structures in material science or biology. The assembly of protein into fibrils can be promoted or inhibited based on the chosen environmental conditions and interaction with suitable components. We review here the key strategies for promotion and inhibition of protein fibrillation in both physiological and non-physiological conditions in order to create functional designs. The major variables discussed are solvent conditions, metals/ions, biopolymers, aromatic compounds, and surface active components. Due to bias in research directions, deeper investigation has traditionally been carried out for inhibition of fibrillation, but focus has recently shifted. Thus, while various strategies are presented on the breakdown of mature proteinfibrils, emphasis is given to the approaches leading to increased rigidity and length of resultant fibrils. We highlight important areas in this field that require further development and promising lines of future experiments.
Article
The structure and internal dynamics of β-lactoglobulin aggregates formed after heat-induced denaturation at pH 2 and different ionic strengths were investigated using light, neutron, and X-ray scattering. Polydisperse aggregates are formed with a rigid rodlike local structure with mass per unit length close to that of a string of β-lactoglobulin monomers but with a somewhat larger diameter. The persistence length decreases with increasing ionic strength from more than 600 nm at 0.013 M to 38 nm at 0.1 M. At ionic strengths of 0.1 and 0.2 M, a self-similar structure with fractal dimensions of 1.8 and 2.0 is seen by using light scattering. The concentration dependence of the static structure factor and the internal dynamics are close to those of flexible linear chains. In contrast, a rigid behavior is observed at lower ionic strength (0.03 and 0.013 M). The persistence length of aggregates formed at 0.013 M is reduced after dilution in 0.1 and 0.2 M ionic strength solvents but remains larger than that of aggregates formed and diluted in 0.1 and 0.2 M. The ionic strength of formation is thus a determining factor for the structure. At pH 2, there is no evidence for a two-step aggregation process as was observed at pH 7.
Article
The effect of high pressure microfluidization on native β-lactoglobulin (β-lg) or whey protein isolate (WPI), both before and after heat-induced protein fibril formation at pH 2.0, was investigated using atomic force microscopy (AFM), shear birefringence, reversed phase-high pressure liquid chromatography, attenuated total reflectance–Fourier transform infrared spectroscopy and fluorescence spectroscopy. The morphology and length distribution of the fibrils were determined using AFM and flow-induced birefringence, respectively. High pressure (≥50 MPa) microfluidization treatment of β-lg induced ∼30% protein denaturation, accompanied by changes in secondary structure. Fibrils formed from high pressure treated β-lg or WPI were similar in length to fibrils formed from non-pressure treated proteins. High pressure (≥50 MPa) microfluidization of fibrils formed from β-lg or WPI resulted in their breakup into more uniformly sized and much shorter fibrils. Microfluidization pressures of up to 170 MPa resulted in slightly shorter fibrils but did not completely dissociate them.
Article
We have investigated the thermodynamic and dynamic behavior of multistranded β-lactoglobulin protein fibrils in water, by combining static, dynamic, and depolarized dynamic light scattering (SLS, DLS, DDLS), small angle neutron scattering (SANS), rheology, and cryogenic transmission electron microscopy (cryo-TEM). We focus on the region of the phase diagram at which ionic strength and concentration changes induce transitions in gelation and lyotropic liquid crystalline behavior. An increase in ionic strength, induced by NaCl salt, progressively causes the phase transitions from nematic (N) to gel (G) phases; a further increase causes the transition to a translucent phase and to a macroscopic phase separation, respectively. An increase in fibril concentration induces first a phase transition from an isotropic (I) to a nematic phase (N); a further increase induces the formation of a gel phase. The protein gel strength is investigated by rheology measurements. SANS and osmotic compressibility calculated by SLS measurements clearly capture the main features of the IN transition of β-lactoglobulin protein fibrils. The form and structure factors measured by scattering experiments are analyzed by the polymer reference interaction site model (PRISM). Dynamics of the protein fibrils at different concentrations, measured by polarized and depolarized dynamic light scattering, show both individual and collective diffusion after the isotropic-nematic transition. Above this transition, cryo-TEM images further demonstrate the alignment of the protein fibrils, which is quantified by a 2D order parameter. This work discusses comprehensively, both experimentally and theoretically, the thermodynamics and dynamic features of β-lactoglobulin amyloid fibrils in a vast region of the concentration-ionic strength phase diagram.
Article
Bacterial and fungal species produce some of the best-characterized functional amyloids, that is, extracellular fibres that play key roles in mediating adhesion and biofilm formation. Yet, the molecular details underlying their mechanical strength remain poorly understood. Here, we use single-molecule atomic force microscopy to measure the mechanical properties of amyloids formed by Als cell adhesion proteins from the pathogen Candida albicans. We show that stretching Als proteins through their amyloid sequence yields characteristic force signatures corresponding to the mechanical unzipping of β-sheet interactions formed between surface-arrayed Als proteins. The unzipping probability increases with contact time, reflecting the time necessary for optimal inter β-strand associations. These results demonstrate that amyloid interactions provide cohesive strength to a major adhesion protein from a microbial pathogen, thereby strengthening cell adhesion. We suggest that such functional amyloids may represent a generic mechanism for providing mechanical strength to cell adhesion proteins. In nanotechnology, these single-molecule manipulation experiments provide new opportunities to understand the molecular mechanisms driving the cohesion of functional amyloid-based nanostructures.
Article
Amyloid fibrils associated with many neurodegenerative diseases are the most intriguing targets of modern structural biology. Significant knowledge has been accumulated about the morphology and fibril-core structure recently. However, no conventional methods could probe the fibril surface despite its significant role in the biological activity. Tip-enhanced Raman spectroscopy (TERS) offers a unique opportunity to characterize the surface structure of an individual fibril due to a high depth and lateral spatial resolution of the method in the nanometer range. Herein, TERS is utilized for characterizing the secondary structure and amino acid residue composition of the surface of insulin fibrils. It was found that the surface is strongly heterogeneous and consists of clusters with various protein conformations. More than 30% of the fibril surface is dominated by β-sheet secondary structure, further developing Dobson's model of amyloid fibrils (Jimenez et al. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 9196-9201). The propensity of various amino acids to be on the fibril surface and specific surface secondary structure elements were evaluated. β-sheet areas are rich in cysteine and aromatic amino acids, such as phenylalanine and tyrosine, whereas proline was found only in α-helical and unordered protein clusters. In addition, we showed that carboxyl, amino, and imino groups are nearly equally distributed over β-sheet and α-helix/unordered regions. Overall, this study provides valuable new information about the structure and composition of the insulin fibril surface and demonstrates the power of TERS for fibril characterization.